You are on page 1of 11

Chemical Engineering Journal 228 (2013) 1110–1120

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Enhanced photocatalytic activities of BiOI/ZnSn(OH)6 composites


towards the degradation of phenol and photocatalytic H2 production
Huiquan Li a,b, Yumin Cui a,⇑, Wenshan Hong a, Bolian Xu b,⇑
a
School of Chemistry and Chemical Engineering, Fuyang Normal College, Fuyang 236037, People’s Republic of China
b
Key Laboratory of Mesoscopic Chemistry of MOE, Jiangsu Provincial Key Laboratory of Nanotechnology, School of Chemistry and Chemical Engineering, Nanjing University,
Nanjing 210093, People’s Republic of 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
Under UV or visible light irradiation with energy (hv) greater than the band-gap energy of BiOI and/or
BiOI/ZnSn(OH)6 catalysts were ZnSn(OH)6, when ZnSn(OH)6 and BiOI were brought in contact, the energy levels difference between
prepared by a facile and economical ZnSn(OH)6 and BiOI caused that the electrons flowed from one phase to another. In this way, the photo-
route.19.7% BiOI/ZnSn(OH)6 showed generated electron–hole pairs BiOI/ZnSn(OH)6 composites were effectively separated.
the best photocatalytic activity.The
photocatalytic mechanism of BiOI/
ZnSn(OH)6 was discussed.A transfer
process of photogenerated carriers was
proposed. O +
2 and h were the main
reactive species for the degradation of
phenol.

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

Article history: BiOI/ZnSn(OH)6 composite photocatalysts with different weight percents of BiOI were successfully syn-
Received 8 December 2012 thesized by a facile and economical deposition method at 80 °C and characterized by X-ray diffraction
Received in revised form 21 May 2013 (XRD), high resolution transmission electron microscopy (HR-TEM), selected area electron diffraction
Accepted 22 May 2013
(SAED), X-ray photoelectron spectroscopy (XPS), photoluminescence (PL) spectroscopy, UV–vis diffuse
Available online 3 June 2013
reflection spectroscopy (UV–vis DRS) and Brunauer–Emmett–Teller (BET) surface area measurements.
Photocatalytic reforming of ethanol for H2 production and photocatalytic degradation of phenol over
Keywords:
BiOI/ZnSn(OH)6 samples were investigated. The results showed that the BiOI/ZnSn(OH)6 photocatalysts
Deposition method
H2 production
exhibited a coexistence of both tetragonal BiOI and cubic perovskite ZnSn(OH)6 phases. With increasing
ZnSn(OH)6 BiOI content, the absorption intensity of BiOI/ZnSn(OH)6 increased in the 280–550 nm region and the
BiOI absorption edge shifted significantly to longer wavelengths compared with that of pure ZnSn(OH)6.
Photocatalysis The 19.7% BiOI/ZnSn(OH)6 catalyst exhibited obviously higher photocatalytic activity as compared to
either pure ZnSn(OH)6 or BiOI, which could be primarily attributed to the presence of intimate contacts
between ZnSn(OH)6 and BiOI resulted in easier charge transfer and more efficient separation of electron–
hole pairs. The photocatalytic mechanism of BiOI/ZnSn(OH)6 was discussed.
Ó 2013 Elsevier B.V. All rights reserved.

1. Introduction of photogenerated electron–hole pairs and the low conversion effi-


ciency of solar energy. The two main drawbacks of TiO2 limit its
TiO2 has been a very attractive research topic for its outstanding industrial application. In order to overcome its drawbacks, one of
bulk and surface physicochemical properties as well as its potential approaches is the modification of TiO2 with noble metals, such as
applications in solar energy conversion, photocatalytic hydrogen Pt [7], and Pd [8]. Nevertheless the stability of these modified
evolution and photocatalytic degradation of environmental pollu- TiO2 has been concerned about owing to the oxidation of noble
tants [1–6]. However, TiO2 suffers from the high recombination metal nanoparticles on the TiO2 surface [9]. Furthermore, the
high cost of noble metal further hinders its application.
⇑ Corresponding authors. Tel.: +86 558 2596249; fax: +86 558 2596703. Therefore, the development of new efficient photocatalysts is still
E-mail addresses: cuiyumin0908@163.com (Y. Cui), xubolian@nju.edu.cn (B. Xu). in progress.

1385-8947/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2013.05.086
H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120 1111

In recent years, some new type semiconductor photocatalysts dissolved in 20 mL ethylene glycol to obtain a clear solution I.
such as a-Ga2O3 [10], Bi2WO6 [11–13], BiOCl [14], BiOBr [15], BiOI ZnSn(OH)6 (1.0 g) was ultrasonically dispersed into deionized
[16], In(OH)3 [17,18], InOOH [19], ZnSn(OH)6 [20] and so on have water to form a homogeneous mixture II. Then the solution I was
been developed. ZnSn(OH)6 has been widely used in high effective added dropwise into the mixture II under strong stirring. After
flame retardants due to its nontoxicity and safety [21,22] and it is a being stirred for 1.0 h at room temperature, the resulting mixtures
perovskite-structured hydroxide, and its surface is full of OH were heated at 80 °C for 2.0 h in a water bath. Finally, the precip-
groups which can accept photogenerated holes to form hydroxyl itates were collected, washed thoroughly with deionized water
radicals (OH), the principal reactive oxidant in photocatalytic reac- and ethanol, and dried at 50 °C in air. The final samples with
tion [23]. In 2012, Chen et al. [20] reported that ZnSn(OH)6 had weight percents of BiOI of 11.0, 19.7 and 32.9 wt%, were denoted
high photocatalytic activity under UV light irradiation for the deg- as 11.0%, 19.7% and 32.9% BiOI/ZnSn(OH)6, respectively. For com-
radation of benzene. In addition, Fu et al. [23] reported that parison, pure BiOI powder was prepared by adopting the method
ZnSn(OH)6, even without the loading of noble metals, showed high mentioned above in the absence of ZnSn(OH)6, and the synthesis
photocatalytic activity toward degradation of benzene and no process of pure BiOI was referred to the literature [27].
deactivation of the catalyst was observed under UV light irradia-
tion for 50 h. But ZnSn(OH)6 has a wide band-gap energy of 2.2. Catalyst characterization
4.0 eV [20], therefore it can only use a small UV fraction of solar
light. In order to expand the optical response of ZnSn(OH)6 from X-ray diffraction (XRD) were performed on a Philips X’ pert dif-
the UV to the visible light region and improve the separation effi- fractometer equipped with Ni-filtered Cu Ka radiation source
ciency of photogenerated electron–hole pairs in the ZnSn(OH)6, (k = 0.15418 nm). X-ray photoelectron spectra (XPS) measurements
one of the common approaches is to combine it with narrow band were carried out using Multilab 2000 XPS system with a monochro-
gap semiconductor. matic Mg Ka source and a charge neutralizer (Multilab 2000 XPS,
Recently, BiOI (bismuth oxyiodide), as a promising photocata- Thermo Scientific, America). The Brunauer–Emmett–Teller (BET)
lyst, has drawn extensive interests of researchers owing to its po- surface areas of samples were determined from N2 adsorption iso-
tential applications in the photocatalytic degradation of organic therms at 77 K using a Micromeritics ASAP 2020 instrument with a
pollutants [24–26]. BiOI has a small band-gap energy of 1.8 eV computer-controlled measurement system. High resolution trans-
and it has strong absorption in the visible light region. As a typical mission electron microscopy (HR-TEM) images and selected area
p-type semiconductor, it may be a potential substitute for nobel electron diffraction (SAED) were taken using a JEM-2100 electron
metal cocatalysts. In 2011, Jiang et al. [27] reported that the ZnO/ microscope. UV–vis diffuse reflection spectroscopy (UV–vis DRS)
BiOI composite photocatalyst had higher photocatalytic activity of the samples was determined with a Shimadzu UV-3600 spectro-
than pure ZnO and BiOI, and the enhanced photocatalytic activity photometer (Japan) using BaSO4 as a reference. The photolumines-
could be primarily attributed to the fact that the heterojunction cence (PL) spectroscopy, obtained at room temperature with an
at the interface between BiOI and ZnO improved the separation excitation wavelength of 280 nm, were recorded on a CARY Eclipse
efficiency of photogenerated electron–hole pairs in the ZnO/BiOI (America) fluorescence spectrophotometer.
phtocatalyst. Thus, BiOI/ZnSn(OH)6 composites may be an ideal
system to expand the optical response to the visible light region 2.3. Photocatalytic reaction
and improve the separation efficiency of photogenerated charge
carriers, and then achieve a high photocatalytic activity. Moreover, 2.3.1. Photocatalytic reforming of ethanol to H2
to the best of our knowledge, there was no report on the prepara- Photocatalytic reforming of ethanol for H2 production reactions
tion and photocatalytic performance in the BiOI/ZnSn(OH)6 was performed under N2 (>99.99%) atmosphere in a top-irradiation
system. quartz reaction cell (18-cm-long, 7-cm-diameter), which was con-
In this work, BiOI/ZnSn(OH)6 composite photocatalysts with dif- nected to a closed gas-recirculation tube system. In a typical exper-
ferent weight percents of BiOI were successfully synthesized by a iment, BiOI/ZnSn(OH)6 photocatalysts (0.05 g) were first suspended
facile and economical deposition method at 80 °C, and the effect in 170 mL ethanol aqueous solution (containing 5.0 mL ethanol).
of BiOI contents on the structure and photocatalytic activity of After stirring for 40 min, the suspension was illuminated by a
BiOI/ZnSn(OH)6 catalysts was investigated in detail. Photocatalytic high-pressure mercury lamp (kmax = 254 nm, 125 W) during UV
reforming of ethanol for H2 production and photocatalytic degrada- light experiments or by a 400 W metal halide lamp (kmax = 588 nm)
tion of phenol over BiOI/ZnSn(OH)6 samples were greatly enhanced with the combination of a cut-off filter (k > 400 nm) to eliminate UV
as compared to either pure ZnSn(OH)6 or BiOI. The photocatalytic radiation during visible light experiments. The solution tempera-
mechanism of BiOI/ZnSn(OH)6 was also investigated. ture was maintained at about 30 °C by circulating water through
a jacket around the reactor. The generated gas of H2 was deter-
mined by an online gas chromatograph (GC, model DS 6200; Donam
2. Experimental Instruments Inc., Gyeonngi-do, Korea).

2.1. Catalyst preparation 2.3.2. Photocatalytic degradation of phenol


Photoacatalytic degradation of phenol was performed in an
The ZnSn(OH)6 sample was prepared by a homogeneous precip- aqueous solution. The UV light was obtained by four 4 W fluores-
itation method. In a common preparation, equal volume of cent UV bulbs (TUV 4 W/G4 T5, Philips, kmax = 254 nm). The visible
0.6 mol L1 NaOH and 0.1 mol L1 ZnCl2 aqueous solution was light source was a 400 W metal halide lamp (kmax = 588 nm) with
added into subsequently 0.1 mol L1 SnCl4 solution with magnetic the combination of a cut-off filter (k > 400 nm) to eliminate UV
stirring at room temperature. The above mixed aqueous solution radiation during visible light experiments. For each UV light test,
was stirred for 40 min to obtain a slurry. Then the slurry was 150 mL phenol aqueous solution (2.12  104 mol L1) and 0.08 g
washed repeatedly with deionized water and separated by filtra- catalyst samples were used in an internal-irradiation quartz reac-
tion. The product was finally dried at 80 °C in air. tion cell (250 mL), while for each visible light test, 40 mL phenol
The BiOI/ZnSn(OH)6 samples with different BiOI contents were aqueous solution (1.06  104 mol L1) and 0.05 g catalyst samples
prepared by a deposition method. In a typical experiment, different were used in an internal-irradiation quartz reaction cell (50 mL). A
stoichiometric amounts of Bi(NO3)35H2O and 0.05 g KI were general procedure was carried out as follows. First, phenol aqueous
1112 H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120

solution was placed into a water-jacketed reactor maintained at Table 1


25 °C, and then the catalyst samples were suspended in the solu- Atomic composition, average crystallite size (L) and BET surface areas (SBET) of BiOI,
ZnSn(OH)6, 11.0%, 19.7% and 32.9% BiOI/ZnSn(OH)6 samples.
tion. The suspension was stirred vigorously for 1.0 h in the dark
to establish the adsorption–desorption equilibrium of phenol, then Sample Elements (atomic L (nm) SBET (m2 g1)
irradiated under UV or visible light. About 2.5 mL solution was composition %)

withdrawn from the reactor periodically and centrifuged and ana- Zn Sn O Bi I


lyzed for the degradation of phenol by using a TU-1901 ZnSn(OH)6 5.1 7.4 42.7 – – 46.5 25.7
spectrophotometer. 11.0% BiOI/ZnSn(OH)6 4.2 6.1 38.9 3.8 2.2 43.7 34.2
In order to investigate the effect of relevant reactive species, a 19.7% BiOI/ZnSn(OH)6 3.4 5.0 37.2 6.4 5.6 41.4 37.8
32.9% BiOI/ZnSn(OH)6 2.9 3.1 35.2 8.3 7.1 38.3 40.4
quantity of different appropriate species quenchers were intro-
BiOI – – 20.8 18.5 16.1 34.8 23.4
duced into the photocatalytic degradation process of phenol in a
manner similar to the photodegradation experiment. The dosages
of these species quenchers were referred to the literatures [28,29].

2.4. Reuse of photocatalyst

The 19.7% BiOI/ZnSn(OH)6 photocatalyst was immersed in eth-


anol for 5.0 h and rinsed with deionized water, and then dried at
80 °C. After this, the 19.7% BiOI/ZnSn(OH)6 photocatalyst was re-
used for the degradation of phenol, and the reuse experiment
was done for five times.

3. Results and discussion

3.1. Catalyst structure

Fig. 1 shows the XRD patterns of BiOI/ZnSn(OH)6 samples with


different BiOI contents. It can be seen that 19.7% and 32.9% BiOI/
Fig. 2. XRD patterns of the 19.7% BiOI/ZnSn(OH)6 catalyst before and after 5.0 h UV
ZnSn(OH)6 samples exhibit a coexistence of both tetragonal BiOI and visible light irradiation.
(JCPDS card No. 73-2062) and cubic perovskite ZnSn(OH)6 (JCPDS
card No. 20-1455) phases, the peaks around 2h of 29.7°, 31.7°,
and 45.6° were indexed to those of tetragonal BiOI and correspond
BiOI/ZnSn(OH)6 samples were calculated and summarized in Ta-
to (0 1 2), (1 1 0), and (0 2 0), respectively. When the amount of BiOI
ble 1, respectively. It can be seen that the BET surface areas of
is lower than 19.7%, no significant diffraction peak of BiOI can be
BiOI/ZnSn(OH)6 composites are obviously higher than those of
detected, which can be ascribed to its lower content and high dis-
pure BiOI and ZnSn(OH)6. In general, a greater BET surface area
persion on the surface of ZnSn(OH)6 particles. The average crystal-
of photocatalysts can supply more surface active sites, leading to
lite sizes of ZnSn(OH)6 in the BiOI/ZnSn(OH)6 composites were
an enhancement of the photocatalytic performance [31]. However,
calculated from the (200) peaks, according to the Scherrer formula
the corresponding BET surface areas of 11.0%, 19.7% and 32.9%
[30]: L = 0.90k/bcos h, and the results were listed in Table 1.
BiOI/ZnSn(OH)6 samples did not have obvious difference in Table 1,
Fig. 2 exhibits the XRD patterns of the 19.7% BiOI/ZnSn(OH)6
which demonstrates that the BET surface area is not the main influ-
catalyst before and after 5.0 h irradiation of UV and visible light
encing factor for the photocatalytic performance of different BiOI/
for the degradation of phenol. It can be clearly observed that the
ZnSn(OH)6 samples.
phase and structure of the 19.7% BiOI/ZnSn(OH)6 catalyst remained
The chemical composition and surface chemical states of the
unchanged after reaction, which suggests that the catalyst is stable
BiOI/ZnSn(OH)6 samples were further investigated by X-ray photo-
in the photocatalytic degradation process of phenol.
electron spectroscopy (XPS). The peak positions in all of the XPS
Based on the nitrogen adsorption–desorption isotherms, the
spectra were calibrated with C 1s at 284.6 eV. The typical survey
BET surface areas of ZnSn(OH)6, BiOI, 11.0%, 19.7% and 32.9%
XPS spectrum of the 19.7% BiOI/ZnSn(OH)6 sample indicates that
the catalyst consists of Sn, Zn, Bi, O and I elements (Fig. 3A). As dis-
played in Fig. 3B and C, the binding energy of Bi 4f and I 3d of 19.7%
BiOI/ZnSn(OH)6 shifts to the higher values compared with that of
Bi 4f and I 3d of pure BiOI at 158.4 eV (Bi 4f7/2) and 163.7 eV (Bi
4f5/2) [26,32], 618.1 eV (I 3d5/2) and 629.6 eV (I 3d3/2) [26], respec-
tively. In Fig. 3D, the binding energy of Zn 2p3/2 of 19.7% BiOI/
ZnSn(OH)6 also shifts to a higher value compared with that of Zn
2p3/2 of pure ZnSn(OH)6 at 1022.0 eV [33]. Above results probably
suggest that there is a strong interaction between BiOI and
ZnSn(OH)6 nanoparticles. The photoelectron peaks at 495.0 and
486.6 eV of ZnSn(OH)6 and 19.7% BiOI/ZnSn(OH)6 samples shown
in Fig. 2E correspond to the binding energies of Sn 3d3/2 and Sn
3d5/2, respectively, which is in accordance with those in the litera-
ture [34]. Because the combination with Bi or I cannot cause obvi-
ous change in the binding energy of Sn 3d, it was hard to say
whether there were chemical bonds existing between Bi or I and
Fig. 1. XRD patterns of BiOI/ZnSn(OH)6 samples with different BiOI contents (wt%): Sn simply, but Bi or I may have a strong interaction with other ele-
(a) 0.00, (b) 11.0, (c) 19.7, (d) 32.9 and (e) 100. ments except Sn in BiOI/ZnSn(OH)6 from XPS characterization [35].
H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120 1113

Fig. 3. XPS survey spectra of the 19.7% BiOI/ZnSn(OH)6 sample (A) and XPS spectra Bi 4f (B), I 3d (C), Zn 2p (D) and Sn 3d (E) of different samples.

The above XPS analysis results at least confirmed that BiOI/ TEM image reveals the highly crystalline nature of the sample.
ZnSn(OH)6 was indeed a composite of BiOI and ZnSn(OH)6. The Clear fringe with an interval of 0.302 nm could be indexed to
XPS results further confirm the coexistence of ZnSn(OH)6 and BiOI (0 1 2) lattice plane of tetragonal BiOI and that of 0.385 nm agreed
in the BiOI/ZnSn(OH)6 samples. The atomic composition of BiOI, with the (2 0 0) lattice plane of cubic perovskite ZnSn(OH)6, which
ZnSn(OH)6, 11.0%, 19.7% and 32.9% BiOI/ZnSn(OH)6 samples was demonstrated that the heterojunction might be formed between
also analyzed by X-ray photoelectron spectroscopy (XPS) study, BiOI and ZnSn(OH)6. The results were consistent with what we ob-
and the results were listed in Table 1. In 11.0% BiOI/ZnSn(OH)6 tained from the XRD patterns in Fig. 1.
sample, BiOI nanoparticles are detected on the ZnSn(OH)6 surface As shown in the corresponding SAED pattern (Fig. 4B), diffrac-
and atomic concentration of Bi and I are only 3.8% and 2.2%. With tion spots with d values of 0.388 and 0.303 nm are observed, which
the increase of the loaded BiOI contents, atomic concentration of Bi correspond to the (2 0 0) lattice plane of cubic perovskite
and I increases to 18.5% and 16.1% in 32.9% BiOI/ZnSn(OH)6, and ZnSn(OH)6 and (0 1 2) lattice plane of tetragonal BiOI, respectively.
there is a similar trend between at% from XPS and the nominal This can also be regarded as a proof for the coexistence of BiOI and
BiOI%, which implies that the ZnSn(OH)6 may be covered by BiOI ZnSn(OH)6.
nanoparticles [36,37]. The optical properties of the BiOI/ZnSn(OH)6 samples with dif-
Fig. 4A presents the HR-TEM image of the 19.7% BiOI/ZnSn(OH)6 ferent BiOI contents were investigated by UV–vis diffuse reflec-
sample in which we can observe different lattice fringes. The HR- tances spectroscopy. As revealed from Fig. 5A, with increasing
1114 H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120

Fig. 4. HR-TEM image (A) and SAED pattern (B) of the 19.7% BiOI/ZnSn(OH)6 sample.

Fig. 5. UV–vis diffuse reflectances spectra (A) of BiOI/ZnSn(OH)6 samples with different BiOI contents (wt%): (a) 0.00, (b) 11.0, (c) 19.7, (d) 32.9 and (e) 100, and the plotting of
(ahv)1/2 versus photon energy (B) and the plotting of (ahv)2 versus photon energy (C).

BiOI contents, the absorption intensity of BiOI/ZnSn(OH)6 samples pure ZnSn(OH)6 sample can be thus estimated from a plot of
increased in the 280–550 nm light region and the absorption edge (ahv)2 versus photon energy (hv), as shown in Fig. 5C. The intercept
shifted significantly to longer wavelengths as compared to pure of the tangent to the x-axis will give a good approximation of the
ZnSn(OH)6 sample, clearly revealing that the absorption edges of band-gap energies for the pure BiOI and ZnSn(OH)6 samples,
BiOI/ZnSn(OH)6 samples shifted to the lower energy region. It is repectively.
well known that the optical absorption near the band edge of a The valence band (VB) edge position and the conduction band
crystalline semiconductor follows the formula: ahv = A(hv  Eg)n/ (CB) edge position of BiOI and ZnSn(OH)6 at the point of zero charge
2
, where a, A, v and Eg are the absorption coefficient, a constant, were calculated by the following empirical formulas [32,39]:
light frequency and band-gap energy, respectively [38]. Among
them, n depends on the characteristics of the transition in a semi- EVB ¼ X  Ee þ 0:5Eg ð1Þ
conductor (direct transition: n = 1; indirect transition: n = 4). For
pure BiOI, the value of n is 4 [39], and for pure ZnSn(OH)6, the value ECB ¼ EVB  Eg ð2Þ
of n is 1 [40]. The band-gap energy (Eg value) of pure BiOI sample
can be thus estimated from a plot of (ahv)1/2 versus photon energy where EVB was the valence band (VB) potential, and ECB was the
(hv), as shown in Fig. 5B, and the band-gap energy (Eg value) of conduction band (CB) potential. X was the electronegativity of the
H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120 1115

semiconductor, and Ee was the energy of free electrons on the equilibrium adsorption capacity (qe) was usually estimated by the
hydrogen scale (4.5 eV). Eg was the band-gap energy of the semi- pseudo-first-order and pseudo-second-order model, and the correct
conductor and it was shown in Fig. 5B and C. Accordingly, the EVB of values were chosen according to the higher correlation coefficient
BiOI and ZnSn(OH)6 were estimated to be 2.31 and 4.20 eV, respec- R2 [43,44]. Herein, the equilibrium adsorption capacity (qe) of BiOI,
tively. The ECB of BiOI and ZnSn(OH)6 were 0.57 and 0.49 eV, ZnSn(OH)6 and 11.0%, 19.7%, 32.9% BiOI/ZnSn(OH)6 catalysts is 7.43,
respectively. 7.86, 10.02, 10.75 and 11:27 mg g1catalyst , respectively. Obviously, the
adsorption ability of the ZnSn(OH)6 was enhanced by introduction
3.2. Adsorption studies of BiOI, which is a prequisite for good photocatalytic activity [45],
but the absorbance ability of 11.0%, 19.7% and 32.9% BiOI/
It is well known that the activity of photocatalyst is related to ZnSn(OH)6 samples did not have obvious difference, demonstrating
its adsorbability [41,42]. Therefore, a research of phenol adsorption that the absorbance ability was also not the main influencing factor
was performed on BiOI, ZnSn(OH)6, 11.0%, 19.7% and 32.9% BiOI/ for the photocatalytic activities of different BiOI/ZnSn(OH)6
ZnSn(OH)6 photocatalysts by using 15 mg of photocatalyst at room samples.
temperature in the dark. The adsorption capacity of all samples
was presented in Fig. 6, in which the adsorption behaviors of phe-
nol on all samples followed the Langmuir model. It can be seen that 3.3. Photocatalytic activity
phenol uptake capacity moderately increases in the presence of the
11.0%, 19.7% and 32.9% BiOI/ZnSn(OH)6 photocatalyst compared 3.3.1. Photocatalytic degradation of phenol
with that of the pure ZnSn(OH)6 or BiOI photocatalyst. The ad- The effect of BiOI contents on the photocatalytic activity of BiOI/
sorbed quantities of phenol, qt (mg g1 catalyst ), at time t were calcu-
ZnSn(OH)6 photocatalysts has been investigated by phenol degra-
lated according to Eq. (3): dation in an aqueous solution under UV and visible light irradiation
(Fig. 7). It can be seen that under UV and visible light irradiation
qt ¼ ½ðC 0  C t Þ  1000  MW  V 0 =wcatalyst ð3Þ the self-degradation of phenol is negligible in the absence of pho-
where C0 and Ct (mol L1) were the initial concentrations and con- tocatalyst, indicating that the photolysis can be ignored. While in
centration at time t of phenol, respectively; MW was the molecular the presence of photocatalysts, the irradiation of UV and visible
weight of phenol (g mol1); V0 was the volume of phenol aqueous light can result in the obvious degradation of phenol, and the BiOI
solution (L); and wcatalyst was the mass of photocatalyst (g). The content in the ZnSn(OH)6 exerts great influences on the photocat-
alytic activity of BiOI/ZnSn(OH)6 photocatalysts. With BiOI content
increasing, the photocatalytic activities of BiOI/ZnSn(OH)6 photo-
catalysts under UV and visible light irradiation first increase,
reaching the maximums at BiOI content of 19.7%, respectively,
and then decrease with further increasing BiOI content. A similar
effect of BiOI content on photocatalytic activity was also reported
in the BiOI/BiOBr [39] and BiOI/Bi2WO6 [45] systems.
In generally, the photocatalytic activity of composite photocat-
alyst is related to its BET surface areas, absorbance abilities, com-
ponent contents, band structure matching and heterojunction
interface, etc. But the corresponding BET surface areas (Table 1)
and the absorbance abilities (Fig. 6) of 11.0%, 19.7%, 32.9% BiOI/
ZnSn(OH)6 samples did not have obvious difference, which demon-
strates that the main influencing factor for the enhanced photocat-
alytic performance of BiOI/ZnSn(OH)6 samples is more likely to be
the heterojunction structure rather than both BET surface areas
and absorbance abilities. With the increase of BiOI content in
BiOI/ZnSn(OH)6 from 11.0% to 19.7%, the more contact areas be-
Fig. 6. Phenol adsorptivity plot for BiOI/ZnSn(OH)6 samples with different BiOI
tween BiOI and ZnSn(OH)6 will be formed, which facilitates the
contents (wt%): (a) 0.00, (b) 11.0, (c) 19.7, (d) 32.9 and (e) 100.
efficient separation of photoinduced electrons and holes, leading

Fig. 7. Effects of BiOI content (wt%) in the BiOI/ZnSn(OH)6 catalysts on the degradation of phenol under UV (A) (kmax = 254 nm) and visible (B) (kmax = 588 nm) light
irradiation: (a) 0.00, (b) 11.0, (c) 19.7, (c’) 19.6, (d) 32.9 and (e) 100 (Deposition method (a–e), Mechanically mixed (c0 )).
1116 H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120

to the enhanced photocatalytic performance. However, If the con- that that there were strong interactions between different phases
tent of BiOI increases continuously to excess, more BiOI and BiOI in the BiOI/ZnSn(OH)6 prepared by deposition method.
contacts were established, which dominated the BiOI and The photocatalytic degradation of organic pollutants ususally
ZnSn(OH)6 contacts. Since BiOI and BiOI contacts were not as effec- follows first-order kinetics [46]. The linear relationship between
tive as BiOI and ZnSn(OH)6 contacts for an effective charge separa- ln(C0/C) and t shown in Fig. 8 confirmed that the photocatalytic
tion. Moreover, the excessive BiOI with narrow band gap could be degradation process of phenol followed the apparent pseudo-
acted as the recombination center of electrons and holes [39], first-order model expressed as Eq. (4) [47–49]:
impeding the photocatalytic activity on the contrary. Thus, 19.7%
t
BiOI/ZnSn(OH)6 with an appropriate BiOI content exhibited the lnðC 0 =CÞ ¼ kapp ð4Þ
highest photocatalytic activity.
In order to confirm the advantage of intimate contacts between where kapp was the apparent pseudo-first-order rate constant
BiOI and ZnSn(OH)6, 19.7% BiOI/ZnSn(OH)6 was compared with the (min1), C was phenol concentration in aqueous solution at time t
mechanical mixture of BiOI and ZnSn(OH)6 with the similar com- (mol L1), C0 was initial phenol concentration (mol L1). Thus, the
position under identical photocatalytic degradation conditions, as kapp of as-prepared samples were calculated and shown in Fig. 9,
shown in Fig. 7. Obviously, the UV and visible light photocatalytic respectively. It can be seen that the kapp for phenol photodegrada-
activity of mechanically mixed 19.6% BiOI/ZnSn(OH)6 were lower tion over the 19.7% BiOI/ZnSn(OH)6 catalyst under UV and visible
than that of 19.7% BiOI/ZnSn(OH)6 prepared by deposition method. light irradiation was about 3.7 and 13.5 times, 7.9 and 2.8 times
This was mainly because BiOI and ZnSn(OH)6 behaved as indepen- higher than that over pure ZnSn(OH)6 and BiOI, respectively. Other
dent photocatalysts rather than a coupled system in the mechani- BiOI/ZnSn(OH)6 catalysts also has much higher kapp values than
cal mixture of BiOI and ZnSn(OH)6, which was unfavorable to the pure ZnSn(OH)6 and BiOI, but lower than 19.7% BiOI/ZnSn(OH)6
transfer of charge carriers from one phase to another. However, photocatalyst. The result suggests that BiOI/ZnSn(OH)6 displayed
the photocatalytic activity of mechanical mixture of BiOI and the enhanced photocatalytic activities than the single ZnSn(OH)6
ZnSn(OH)6 was enhanced compared with that of pure BiOI and and BiOI while 19.7% BiOI/ZnSn(OH)6 had the best photocatalytic
ZnSn(OH)6, even the mathematical sum of them. This might be ow- activity compared to 11.0% BiOI/ZnSn(OH)6 and 32.9% BiOI/
ing to the inevitable partial contact between BiOI and ZnSn(OH)6 in ZnSn(OH)6.
the mechanical mixture of BiOI and ZnSn(OH)6, which was in favor To test the stability of the 19.7% BiOI/ZnSn(OH)6 sample for the
of the photocatalytic activity to some extent but was still far infe- photocatalytic reaction of phenol under UV and visible light irradi-
rior to the effect of BiOI/ZnSn(OH)6 composites. The results showed ation, the catalyst was reused for photocatalytic reaction 5 times

Fig. 8. Linear transform ln(C0/C) = f(t) of the kinetic curves of phenol degradation over BiOI/ZnSn(OH)6 catalysts with different BiOI contents (wt%) under UV (A)
(kmax = 254 nm) and visible (B) (kmax = 588 nm) light irradiation: (a) 0.00, (b) 11.0, (c)19.7, (c0 ) 19.6, (d) 32.9 and (e)100 (Deposition method (a–e), Mechanically mixed (c0 )).

Fig. 9. Photocatalytic degradation rate constant k of phenol as a function of BiOI content (wt%) under UV (A) (kmax = 254 nm) and visible (B) (kmax = 588 nm) light irradiation:
(a) 0.00, (b) 11.0, (c) 19.7, (c0 ) 19.6, (d) 32.9 and (e) 100 (Deposition method (a–e), Mechanically mixed (c0 )).
H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120 1117

Fig. 10. Reuse in the photocatalytic degradation of phenol in the presence of 19.7% BiOI/ZnSn(OH)6 catalyst under UV (A) (kmax = 254 nm) and visible (B) (kmax = 588 nm) light
irradiation.

Fig. 11. The amounts of H2 generated (A and B) and the H2 evolution rate obtained (C and D) over BiOI/ZnSn(OH)6 catalysts with different BiOI contents (wt%) under UV (A, C)
and visible (B, D) light irradiation: (a) 0.00, (b) 11.0, (c) 19.7, (d) 32.9 and (e) 100.

under the same conditions and the result is shown in Fig. 10. It can alysts under UV and visible light irradiation were plotted as a
be seen that the photocatalytic activity does not exhibit any signif- function of time in Fig. 11A and B, respectively. It can be seen that
icant loss after five recycles for the photodegradation of phenol, the generation of H2 almost linearly increases with irradiation
which indicate that the catalyst is stable for the photocatalysis of time. The corresponding H2 evolution rates were showed in
phenol molecules. Fig. 11C and D, respectively. Clearly, pure BiOI shows very poor
activity for hydrogen generation under UV light irradiation
3.3.2. Photocatalytic reforming of ethanol to H2 (Fig. 11C), which is mainly because pure BiOI with narrow band
The effect of BiOI contents on the photocatalytic reforming of gap could be easily acted as the recombination center of electrons
ethanol to H2 over BiOI/ZnSn(OH)6 photocatalysts under UV and and holes. Fig. 11D shows that the ZnSn(OH)6 alone has no photo-
visible light irradiation was also investigated. Blank experiments catalytic activity for H2 production under visible light irradiation,
showed that without photocatalysts or irradiation, no H2 was pro- which can be attributed to the fact that pure ZnSn(OH)6 with a
duced. The amounts of H2 produced over BiOI/ZnSn(OH)6 photocat- large band-gap energy (3.71 eV) cannot be excited by visible light.
1118 H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120

The photocatalytic activities of BiOI/ZnSn(OH)6 composite photo- 3.4.2. Photocatalytic activity enhancement mechanism of BiOI/
catalysts under UV and visible light irradiation first increase with ZnSn(OH)6
the increment of BiOI component, reaching a maximum at BiOI On the basis of the above experimental studies, a schematic dia-
content of 19.7%, whereas a further increase in the amount of BiOI gram of the band levels of the BiOI/ZnSn(OH)6 composites and the
component leads to a decrease in H2 evolution rates. possible reaction mechanism of the photocatalytic procedure are
The enhanced photocatalytic activities of BiOI/ZnSn(OH)6 com- proposed and illustrated in Fig. 13. It is well known that using
posites are attributed to the synergic effect between BiOI and two semiconductors in contact with different redox energy levels
ZnSn(OH)6, which plays a key role to the effective separation of of conduction band (CB) and valence band (VB) can be used to im-
photogenerated electron–hole pairs. At lower BiOI content, a good prove photogenerated carriers separation and enhance the effi-
dispersion of BiOI on the ZnSn(OH)6 surface could be achieved, ciency of the interfacial charge transfer [27]. The relative
which is supported by the XRD analysis (Fig. 1). In this case, positions of energy bands of ZnSn(OH) (EVB = 4.20 eV, ECB = 0.49 -
the increase of BiOI content would lead to more BiOI generated eV) and BiOI (EVB = 2.31 eV, ECB = 0.57 eV) are known and shown
on the ZnSn(OH)6 surface, which creates larger BiOI-ZnSn(OH)6 in Fig. 13. It can be seen that both ZnSn(OH)6 and BiOI can be
contact areas and allows more efficient charge transfer between simultaneously excited to form electron–hole pairs under UV light
ZnSn(OH)6 and BiOI. However, excess BiOI loading, the fine BiOI irradiation, then the photoinduced electrons can transfer easily
nanoparticles might be agglomerated, leading to a relatively poor from the CB bottom of ZnSn(OH)6 to that of BiOI. When ZnSn(OH)6
dispersion, and the excessive BiOI with narrow band gap could be and BiOI semiconductors are in contact, the conduction band (CB)
also acted as the recombination center of electrons and holes potential of ZnSn(OH)6 is more negative than that of BiOI. There is
[39], which is unfavorable for high photocatalytic efficiency of hence diffusion of electrons from ZnSn(OH)6 to BiOI, resulting in
the BiOI/ZnSn(OH)6, as a result, the hydrogen generation gradu- accumulation of negative charges in BiOI close to the junction. At
ally becomes worse. Therefore, there is an optimal loading of the same time, the holes transfer from BiOI to ZnSn(OH)6, leaving
19.7% BiOI on ZnSn(OH)6, which showed the best photocatalytic a positive section in ZnSn(OH)6 near the junction. With equilibra-
activity in our studies. tion of BiOI and ZnSn(OH)6 Fermi levels, an internal electric field
directed from ZnSn(OH)6 to BiOI is simultaneously built to stop
3.4. Photocatalytic mechanism the charge diffusion from ZnSn(OH)6 into BiOI. Meanwhile, the en-
ergy bands of ZnSn(OH)6 shifted downward along with the Fermi
3.4.1. Effects of reactive species level, whereas the energy bands of BiOI shift upward in this pro-
The effect of various radical scavengers on the degradation of cess [27,32,39,52–54] (Fig. 13B). Under the irradiation of visible
phenol over the 19.7% BiOI/ZnSn(OH)6 sample under UV and visi- light, BiOI is excited and there is the generation of electron–hole
ble light irradiation was performed to study the underlying photo- pairs. This new CB edge potential (about 0.65 eV) [32,39,52] of
degradation mechanism. Benzoquinone (BQ), as an  O 2 scavenger, BiOI is more negative than that of ZnSn(OH)6, resulting in the pho-
was introduced to the reaction system [28,50]. As the scavenger toinduced electrons at the surface of BiOI can easily migrate to the
of  OH, isopropanol (IPA), was added to the reaction system CB energy level of ZnSn(OH)6, while the excited holes remain on
[28,51]. In order to investigate the role of h+ and H2O2 radical spe- the VB of BiOI. Furthermore, the migration of photogenerated elec-
cies, ammonium oxalate (AO) [29] and catalase (CAT) [28] were trons and holes can be promoted by the internal electric field. Thus,
also introduced to the reaction system, respectively. The results the recombination of photogenerated electron–hole pairs could be
were shown in Fig. 12. Once the radical species played a major role effectively inhibited and the corresponding UV and visible light
in the degradation of phenol, the degradation ratio was expected to photocatalytic activities would be greatly improved, as clarified
be decreased remarkablely. by following equations:
As shown in Fig. 12, the addition of IPA had a negligible effect on þ
BiOI þ hv ðUV or visibleÞ ! ecb þ hvb ð5Þ
the degradation ratio of phenol compared with no scavenger under
otherwise identical conditions, indicating that  OH was not the þ
main reactive species. However, the degradation ratio of phenol ZnSnðOHÞ6 þ hv ðUVÞ ! ecb þ hvb ð6Þ
decreased significantly after the addition of BQ, AO and CAT,
respectively. That is to say,  O + + ecb þ O2 !  O2 ð7Þ
2 , h and H2O2, especially h and
 
O2 , jointly dominated the photodegradation process of phenol
while  OH could be ignored. ecb þ  O2 þ 2Hþ ! H2 O2 ð8Þ

Fig. 12. The effect of different scavengers on the degradation of phenol over the 19.7% BiOI/ZnSn(OH)6 sample under UV (A) and visible (B) light irradiation.
H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120 1119

Fig. 13. Diagrams of energy position and photogenerated electron–hole pairs transfer of BiOI and ZnSn(OH)6 under UV (A) and visible (B) light irradiation.

2ecb þ 2Hþ ! H2 ð9Þ greatly inhibited by the BiOI introduction. In other words, an
appropriate amount of BiOI in the ZnSn(OH)6 was helpful to sepa-
phenol þ  O2 ! products ð10Þ rate the photogenerated charge carriers and increase the lifetime of
charge carriers, and then improve the photocatalytic perfomance
phenol þ H2 O2 ! products ð11Þ of BiOI/ZnSn(OH)6 photocatalysts.

þ
phenol þ hvb ! products ð12Þ 4. Conclusions

When irradiation with energy (hv) greater than the band-gap en- In summary, the BiOI/ZnSn(OH)6 composite photocatalysts with
þ
ergy of BiOI and ZnSn(OH)6, electron–hole pairs (ecb  =hvb ) are gen- different BiOI contents were successfully synthesized by a facile
þ
erated in BiOI and ZnSn(OH)6 (Eqs. (5) and (6)). Then, the (ecb  =hvb and economical deposition method at 80 °C. With increasing BiOI
pairs migrated to the surface of the catalysts and reacted with the content, the absorption intensity of BiOI/ZnSn(OH)6 samples in-
species adsorbed on the surface (e.g. Eqs. (7)–(12)). These reactions creased in the 280–550 nm light region and the absorption edge
þ
prevented the e cb =hvb pairs from combining. The better separation shifted significantly to longer wavelengths as compared to pure
of electrons and holes in the BiOI/ZnSn(OH)6 photocatalysts was ZnSn(OH)6 sample. Photocatalytic reforming of ethanol for H2 pro-
confirmed by photoluminescence (PL) emission spectra of pure duction and photocatalytic degradation of phenol over the 19.7%
ZnSn(OH)6 and 19.7% BiOI/ZnSn(OH)6 photocatalysts in Fig. 14. BiOI/ZnSn(OH)6 sample were greatly enhanced as compared to
It is well known that PL emission spectra have been a useful either pure ZnSn(OH)6 or BiOI sample, which could be primarily
technique to investigate the separation efficiency of photogenerat- attributed to the presence of intimate contacts between ZnSn(OH)6
ed charge carriers in a semiconductor photocatalyst [55–59]. The and BiOI resulted in easier charge transfer and more efficient sep-
comparison of PL emission spectra (excited at 280 nm) of pure aration of electron–hole pairs. The tests of radical scavengers con-
ZnSn(OH)6 and 19.7% BiOI/ZnSn(OH)6 photocatalysts at room tem- firmed that  O +
and H2O2, especially h+ and  O
2, h 2 , jointly
perature was shown in Fig. 14. It can be seen that pure ZnSn(OH)6 dominated the photodegradation process of phenol while  OH
and 19.7% BiOI/ZnSn(OH)6 photocatalysts have a broad emission could be ignored. This work provides a simple and economical
peak in the wavelength range of 370–430 nm. The strongest emit- route to couple two different semiconductors with different phys-
ting peaks around 385 nm are similar while PL emission intensity icochemical properties for developing novel high-performance
of the 19.7% BiOI/ZnSn(OH)6 photocatalyst is dramatically weak- photocatalysts.
ened compared with that of pure ZnSn(OH)6 photocatalyst, indicat-
ing that the recombination of photogenerated charge carriers was
Acknowledgments

This work was supported by the National Natural Science Foun-


dation of China (21201037) and Natural Science Foundation of
Higher Education Institutions in Anhui Province (KJ2012A217)
and Anhui Provincical Key Laboratory for Degradation and Moni-
toring of Pollution of the Environment.

References

[1] X.B. Chen, L. Liu, P.Y. Yu, S.S. Mao, Increasing solar absorption for
photocatalysis with black hydrogenated titanium dioxide nanocrystals,
Science 331 (2011) 746–750.
[2] G. Palmisano, E. García-López, G. Marcí, V. Loddo, S. Yurdakal, V. Augugliaro, L.
Palmisano, Advances in selective conversions by heterogeneous
photocatalysis, Chem. Commun. 46 (2010) 7074–7089.
[3] Y.Z. Li, H. Zhang, Z.M. Guo, J.J. Han, X.J. Zhao, Q.N. Zhao, S.J. Kim, Highly efficient
visible-light-induced photocatalytic activity of manostructured AgI/TiO2
Photocatalyst, Langmuir 24 (2008) 8351–8357.
[4] P. Oancea, T. Oncescu, Photocatalytic degradation of dichlorvos in aqueous
TiO2 suspensions, J. Photochem. Photobiol. A 199 (2008) 8–13.
Fig. 14. Photoluminescence (PL) spectra of ZnSn(OH)6 and 19.7% BiOI/ZnSn(OH)6 [5] J. Velásquez, S. Valencia, L. Rios, G. Restrepo, J. Marín, Characterization and
samples recorded at room temperature with the excitation wavelength of 280 nm. photocatalytic evaluation of polypropylene and polyethylene pellets coated
1120 H. Li et al. / Chemical Engineering Journal 228 (2013) 1110–1120

with P25 TiO2 using the controlled-temperature embedding method, Chem. [32] X. Zhang, L. Zhang, T.F. Xie, D.J. Wang, Low-temperature synthesis and high
Eng. J. 203 (2012) 398–405. visible-light-induced photocatalytic activity of BiOI/TiO2 heterostructures, J.
[6] Y.Y. Lv, L.S. Yu, X.L. Zhang, J.Y. Yao, R.Y. Zou, Z. Dai, P-doped TiO2 nanoparticles Phys. Chem. C 113 (2009) 7371–7378.
film coated on ground glass substrate and therepeated photodegradation of [33] L. Zou, X. Xiang, M. Wei, F. Li, D.G. Evans, Single-crystalline ZnGa2O4 spinel
dye under solar light irradiation, Appl. Surf. Sci. 257 (2011) 5715–5719. phosphor via a single-source inorganic precursor route, Inorg. Chem. 47 (2008)
[7] M. Maicu, M.C. Hidalgo, G. Colón, J.A. Navío, Comparative study of the 1361–1369.
photodeposition of Pt, Au and Pd on pre-sulphated TiO2 for the photocatalytic [34] D. Deng, J.Y. Lee, Hollow core–shell mesospheres of crystalline SnO2
decomposition of phenol, J. Photochem. Photobiol. A: Chem. 217 (2011) 275– nanoparticle aggregates for high capacity Li+ ion storage, Chem. Mater. 20
283. (2008) 1841–1846.
[8] J.B. Zhong, Y. Lua, W.D. Jiang, Q.M. Meng, X.Y. He, J.Z. Li, Y.Q. Chen, [35] Y.C. Zhang, Z.N. Du, K.W. Li, M. Zhang, D.D. Dionysiou, High-performance
Characterization and photocatalytic property of Pd/TiO2 with the oxidation visible-light-driven SnS2/SnO2 nanocomposite photocatalyst prepared via
of gaseous benzene, J. Hazard. Mater. 168 (2009) 1632–1635. in situ hydrothermal oxidation of SnS2 nanoparticles, ACS Appl. Mater.
[9] H. Einaga, T. Ibusuki, S. Futamura, Improvement of catalyst durability by Interfaces 3 (2011) 1528–1537.
deposition of Rh on TiO2 in photooxidation of aromatic compounds, Environ. [36] C.Z. Sun, L.C. Liu, L. Qi, H. Li, H.L. Zhang, C.S. Li, F. Gao, L. Dong, Efficient
Sci. Technol. 38 (2004) 285–289. fabrication of ZrO2-doped TiO2 hollow nanospheres with enhanced
[10] Y.D. Hou, X.C. Wang, L. Wu, Z.X. Ding, X.Z. Fu, Efficient decomposition of photocatalytic activity of rhodamine B degradation, J. Colloid Interf. Sci. 364
benzene over a beta-Ga2O3 photocatalyst under ambient conditions, Environ. (2011) 288–297.
Sci. Technol. 40 (2006) 5799–5803. [37] X. Li, J.T. Chen, H.L. Li, J.T. Li, Y.T. Xu, Y.J. Liu, J.R. Zhou, Photoreduction of CO2 to
[11] A.K.P. Mann, S.E. Skrabalak, Synthesis of single-crystalline nanoplates by spray methanol over Bi2S3/CdS photocatalyst under visible light irradiation, J. Nat.
pyrolysis: a metathesis route to Bi2WO6, Chem. Mater. 23 (2011) 1017–1022. Gas Chem. 20 (2011) 413–417.
[12] D.X. Wu, H.T. Zhu, C.Y. Zhang, L. Chen, Novel synthesis of bismuth tungstate [38] X. Zhang, Z.H. Ai, F.L. Jia, L.Z. Zhang, Generalized one-pot synthesis,
hollow nanospheres in water–ethanol mixed solvent, Chem. Commun. 46 characterization and photocatalytic activity of hierarchical BiOX (X = Cl, Br, I)
(2010) 7250–7252. nanoplate microspheres, J. Phys. Chem. C 112 (2008) 747–753.
[13] F. Amano, K. Nogami, M. Tanaka, B. Ohtani, Correlation between surface area [39] J. Cao, B.Y. Xu, H.L. Lin, B.D. Luo, S.F. Chen, Chemical etching preparation of
and photocatalytic activity for acetalde-hyde decomposition over bismuth BiOI/BiOBr heterostructures with enhanced photocatalytic properties for
tungstate, Langmuir 26 (2010) 7174–7180. organic dye removal, Chem. Eng. J. 185 (186) (2012) 91–99.
[14] F. Dong, Y.J. Sun, M. Fu, Z.B. Wu, S.C. Lee, Room temperature synthesis and [40] L.L. Wang, K.B. Tang, Z.P. Liu, D.K. Wang, J. Sheng, W. Cheng, Single-crystalline
highly enhanced visible light photocatalytic activity of porous BiOI/BiOCl ZnSn(OH)6 hollow cubes via self-templated synthesis at room temperature
composites nanoplates microflowers, J. Hazard. Mater. 219–220 (2012) 26–34. and their photocatalytic properties, J. Mater. Chem. 21 (2011) 4352–4357.
[15] L. Kong, Z. Jiang, H.H. Lai, R.J. Nicholls, T.C. Xiao, M.O. Jones, P.P. Edwardsa, [41] X. Chen, S.S. Mao, Titanium dioxide nanomaterials: synthesis, properties,
Unusual reactivity of visible-light-responsive AgBr–BiOBr heterojunction modifications, and applications, Chem. Rev. 107 (2007) 2891–2959.
photocatalysts, J. Catal. 293 (2012) 116–125. [42] S.F. Chen, Y.Z. Liu, Study on the photocatalytic degradation of glyphosate by
[16] H. Liu, W.R. Cao, Y. Su, Y. Wang, X.H. Wang, Synthesis, characterization and TiO2 photocatalyst, Chemosphere 67 (2007) 1010–1017.
photocatalytic performance of novel visible-light-induced Ag/BiOI, Appl. Catal. [43] W. Kangwansupamonkon, W. Jitbunpot, S. Kiatkamjornwong, Photocatalytic
B: Environ. 111–112 (2012) 271–279. efficiency of TiO2/poly[acrylamide-co-(acrylic acid)] composite for textile dye
[17] T.J. Yan, J.L. Long, Y.S. Chen, X.X. Wang, D.Z. Li, X.Z. Fu, Indium hydroxide: a degradation, Polym. Degrad. Stabil. 95 (2010) 1894–1902.
highly active and low deactivated catalyst for photoinduced oxidation of [44] Y.S. Ho, G. McKay, A comparison of chemisorption kinetic models applied to
benzene, Comptes Rendus Chim. 11 (2008) 101–106. pollutant removal on various sorbents, Trans. Inst. Chem. Eng. 76B (1998)
[18] T.J. Yan, X.X. Wang, J.L. Long, P. Liu, X.L. Fu, G.Y. Zhang, X.Z. Fu, Urea-based 332–340.
hydrothermal growth, optical and photocatalytic properties of single- [45] H.Q. Li, Y.M. Cui, W.S. Hong, High photocatalytic performance of BiOI/Bi2WO6
crystalline In(OH)3 nanocubes, J. Colloid. Interface Sci. 325 (2008) 425–431. toward toluene and Reactive Brilliant Red, Appl. Surf. Sci. 264 (2013) 581–588.
[19] Z.H. Li, Z.P. Xie, Y.F. Zhang, L. Wu, X.X. Wang, X.Z. Fu, Wide band gap p-block [46] S.S. Hong, M.S. Lee, S.S. Park, G.D. Lee, Synthesis of nanosized TiO2/SiO2
metal oxyhydroxide InOOH: a new durable photocatalyst for benzene particles in the microemulsion and their photocatalytic activity on the
degradation, J. Phys. Chem. C 111 (2007) 18348–18352. decomposition of p-nitrophenol, Catal. Today 87 (2003) 99–105.
[20] Y.B. Chen, D.Z. Li, M. He, Y. Hu, H. Ruan, Y.M. Lin, J.H. Hu, Y. Zheng, Y. Shao, [47] Y. Li, X. Li, J. Li, J. Yin, Photocatalytic degradation of methyl orange by TiO2-
High photocatalytic performance of zinc hydroxystannate toward benzene and coated activated carbon and kinetic study, Water Res. 40 (2006) 1119–1126.
methyl orange, Appl. Catal. B: Environ. 113–114 (2012) 134–140. [48] J.H. Sun, X.L. Wang, J.Y. Sun, R.X. Sun, S.P. Sun, L.P. Qiao, Photocatalytic
[21] J.Z. Xu, C.Y. Zhang, H.Q. Qu, C.M. Tian, Zinc hydroxystannate and zinc stannate degradation and kinetics of Orange G using nano-sized Sn(IV)/TiO2/
as flame-retardant agents for flexible poly (vinyl chloride), J. Appl. Polym. Sci. ACphotocatalyst, J. Mol. Catal. A: Chem. 260 (2006) 241–246.
98 (2005) 1469–1475. [49] C.H. Wu, H.W. Chang, J.M. Chen, Basic dye decomposition kinetics in
[22] F. Andre, P.A. Cusack, A.W. Monk, R. Seangprasertkij, The effect of zinc photocatalytic slurry reactor, J. Hazard. Mater. 137 (2006) 336–343.
hydroxystannate and zinc stannate on the fire properties of polyester resins [50] M.C. Yin, Z.S. Li, J.H. Kou, Z.G. Zou, Mechanism investigation of visible light-
containing additive-type halogenated flame retaredants, Polym. Degrad. Stab. induced degradation in a heterogeneous TiO2/Eosin Y/Rhodamine B System,
40 (1993) 267–273. Environ. Sci. Technol. 43 (2009) 8361–8366.
[23] X.L. Fu, X.X. Wang, Z.X. Ding, D.Y.C. Leung, Z.Z. Zhang, J.L. Long, Hydroxide [51] L.S. Zhang, K.H. Wong, H.Y. Yip, C. Hu, J.C. Yu, C.Y. Chan, P.K. Wong, Effective
ZnSn(OH)6: a promising new photocatalyst for benzene degradation, Appl. photocatalytic disinfection of E. coli K-12 using AgBr–Ag–Bi2WO6
Catal. B: Environ. 91 (2009) 67–72. nanojunction system irradiated by visible light: the role of diffusing
[24] J.X. Xia, S. Yin, H.M. Li, H. Xu, L. Xua, Q. Zhang, Enhanced photocatalytic activity hydroxyl radicals, Environ. Sci. Technol. 44 (2010) 1392–1398.
of bismuth oxyiodine (BiOI) porous microspheres synthesized via reactable [52] Y.Y. Li, J.S. Wang, H.C. Yao, L.Y. Dang, Z.J. Li, Chemical etching preparation of
ionic liquid-assisted solvothermal method, Colloid Surf. A: Physicochem. Eng. BiOI/Bi2O3 heterostructures with enhanced photocatalytic activities, Catal.
Asp. 387 (2011) 23–28. Commun. 12 (2011) 660–664.
[25] Y.Y. Li, J.S. Wang, H.C. Yao, L.Y. Dang, Z.J. Li, Efficient decomposition of organic [53] Y.Y. Li, J.S. Wang, B. Liu, L.Y. Dang, H.C. Yao, Z.J. Li, BiOI-sensitized TiO2 in
compounds and reaction mechanism with BiOI photocatalyst under visible phenol degradation: a novel efficient semiconductor sensitizer, Chem. Phys.
light irradiation, J. Mol. Catal. A: Chem. 334 (2011) 116–122. Lett. 508 (2011) 102–106.
[26] Y.N. Wang, K.J. Deng, L.Z. Zhang, Visible light photocatalysis of BiOI and its [54] L. Chen, S.F. Yin, S.L. Luo, R. Huang, Q. Zhang, T. Hong, P.C.T. Au, Bi2O2CO3/BiOI
photocatalytic activity enhancement by in situ ionic liquid modification, J. photocatalysts with heterojunctions highly efficient for visible-light treatment
Phys. Chem. C 115 (2011) 14300–14308. of dye-containing wastewater, Ind. Eng. Chem. Res. 51 (2012) 6760–6768.
[27] J. Jiang, X. Zhang, P.B. Sun, L.Z. Zhang, ZnO/BiOI Heterostructures: [55] J.W. Tang, Z.G. Zou, J.H. Ye, Photophysical and photocatalytic properties of
photoinduced charge-transfer property and enhanced visible-light AgInW2O8, J. Phys. Chem. B 107 (2003) 14265–14269.
photocatalytic activity, J. Phys. Chem. C 115 (2011) 20555–20564. [56] K. Fujihara, S. Izumi, T. Ohno, M.J. Matsumura, Time-resolved
[28] G.T. Li, K.H. Wong, X.W. Zhang, C. Hu, J.C. Yu, R.C.Y. Chan, P.K. Wong, photoluminescence of particulate TiO2 photocatalysts suspended in aqueous
Degradation of acid orange 7 using magnetic AgBr under visible light: the roles solutions, Photochem. Photobiol. A: Chem. 132 (2000) 99–104.
of oxidizing species, Chemosphere 76 (2009) 1185–1191. [57] J.G. Yu, H.G. Yu, B. Cheng, X.J. Zhao, J.C. Yu, W.K. Ho, The Effect of calcination
[29] N. Zhang, S.Q. Liu, X.Z. Fu, Y.J. Xu, Synthesis of M@TiO2 (M = Au, Pd, Pt) core– temperature on the surface microstructure and photocatalytic activity of TiO2
shell nanocomposites with tunable photoreactivity, J. Phys. Chem. C 115 thin films prepared by liquid phase deposition, J. Phys. Chem. B 107 (50)
(2011) 9136–9145. (2003) 13871–13879.
[30] M. Galceran, M.C. Pujol, C. Zaldo, F. Díaz, M. Aguiló, Synthesis, structural, and [58] X.Z. Li, F.B. Li, C.L. Yang, W.K. Ge, Photocatalytic activity of WOx–TiO2 under
optical properties in monoclinic Er:KYb(WO4)2 nanocrystals, J. Phys. Chem. C visible light irradiation, J. Photochem. Photobiol. A 141 (2001) 209–217.
113 (2009) 15497–15506. [59] L.Q. Jing, Y.C. Qu, B.Q. Wang, S.D. Li, B.J. Jiang, L.B. Yang, W. Fu, H.G. Fu, J.Z. Sun,
[31] H.Q. Li, Y.M. Cui, X.C. Wu, W.S. Hong, H. Lin, The Effect of La contents on the Review of photoluminescenceperformance of nano-sized semiconductor
structure and photocatalytic activity of La-SrTiO3 catalysts, Chin. J. Inorg. materials and its relationships with photocatalytic activity, Sol. Energy. Mat.
Chem. 28 (2012) 2597–2604. Sol. Cells 90 (2006) 1773–1787.

You might also like