You are on page 1of 7

RSC Advances

View Article Online


PAPER View Journal | View Issue
Published on 08 July 2014. Downloaded by State University of New York at Stony Brook on 16/09/2014 11:30:11.

Synthesis, characterization, and visible-light


photocatalytic activity of BiOI hierarchical flower-
Cite this: RSC Adv., 2014, 4, 31393
like microspheres†
Jingyi Liu,a Haiping Li,b Na Du,c Shue Songc and Wanguo Hou*c

BiOI hierarchical flower-like microspheres were hydrothermally prepared, using tetrabutylammonium


iodide as an iodine source and template. Many BiOI hierarchical structures have been synthesized using
KI, NaI, HI or ionic liquids as the iodine source, but the use of alkyl ammonium iodide as the iodine
source was not reported in the literature. The so-obtained BiOI samples were characterized by X-ray
diffraction, scanning electron microscopy, transmission electron microscopy, nitrogen sorption
measurements, and ultraviolet-visible diffuse reflectance spectroscopy. The effects of hydrothermal
temperature and time on the BiOI crystal structure and morphology were investigated. The BiOI
microspheres were composed of BiOI nanosheets. The photocatalytic performance of the BiOI samples
was determined from the degradation of Rhodamine B, under visible light irradiation. BiOI microspheres
prepared at 160  C over 30 h exhibited excellent photodegradation efficiency, which was more than five
and seven times higher than those of BiOI nanoplates and N-doped TiO2, respectively. The high
photocatalytic performance was attributed to the high specific surface area and low nanosheet
Received 22nd May 2014
Accepted 7th July 2014
thickness. A morphologic factor was proposed to represent the ratio of specific surface area to
nanosheet thickness, and correlated to the photodegradation efficiency of the BiOI samples. The
DOI: 10.1039/c4ra04829f
photocatalytic efficiency increased with increasing morphologic factor. The BiOI photocatalyst exhibited
www.rsc.org/advances excellent stability and reusability, and has potential in environment remediation.

structure, characterized by [Bi2O2]2+ slabs interleaved by double


1. Introduction slabs of iodine atoms.17–19 Several BiOI nanostructures have
Environmental pollution currently receives global attention,1,2 been reported, including nanoplates20–22 and three-dimensional
and new technologies and materials are continually being (3D) hierarchical microspheres.23–27 Zhang et al. synthesized
developed to assist with this problem. The photodegradation of ower-like BiOI nanoplates via an ethylene glycol (EG)-assisted
pollutants by semiconductor photocatalysts is a promising solvothermal process at 170  C.25 Xia et al. fabricated BiOI
approach, which has evoked great interest among scientists.3–6 porous microspheres by a reactable ionic liquid (1-butyl-3-
TiO2 has received tremendous attention throughout the last two methylimidazolium iodine) assisted solvothermal method.26
decades, because of its wide availability and low toxicity.7–9 TiO2 Ren et al. prepared hollow ower-like microspheres with the
has a wide band gap (3.2 eV), so is responsive to ultraviolet assistance of polyvinylpyrrolidone and citric acid through a
wavelengths. It exhibits low photocatalytic activity under solvothermal process at 70  C.27 These BiOI hierarchical struc-
sunlight conditions, which limits its use.10–12 Developing effi- tures were synthesized using KI, NaI, HI or ionic liquids as the
cient visible light photocatalysts is of great importance. iodine source,25–34 but the use of alkyl ammonium iodide as the
BiOI has a narrow band gap (1.77–1.92 eV), and has attracted iodine source was not reported in the literature.
much interest as a visible light photocatalyst for pollutant Herein, we report the hydrothermal synthesis of BiOI ower-
degradation.13–16 It is a ternary compound with a layered like microspheres, in which tetrabutylammonium iodide (TBAI)
was used as an iodine source and template. To the best of our
knowledge, this is the rst report of a BiOI hierarchical struc-
a
Environment Research Institute, Shandong University, Jinan 250100, P.R. China ture synthesized using an alkyl ammonium iodide as the iodine
b
National Engineering Technology Research Center for Colloidal Materials, Shandong source. The effects of hydrothermal reaction temperature (TH)
University, Jinan 250100, P.R. China and time (tH) on the crystal structure, morphology, and photo-
c
Key Laboratory of Colloid and Interface Chemistry (Ministry of Education), Shandong catalytic performance of the BiOI samples were investigated.
University, Jinan 250100, P.R. China. E-mail: wghou@sdu.edu.cn; Fax: +86-531-
The BiOI ower-like microspheres exhibited higher photo-
88364750; Tel: +86-531-88365460
† Electronic supplementary information (ESI) available. See DOI:
catalytic activity than BiOI nanoplates and N-doped TiO2 (N-
10.1039/c4ra04829f TiO2), and have potential in environment remediation.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 31393–31399 | 31393
View Article Online

RSC Advances Paper

20  C. In a typical process, 0.01 g of BiOI was dispersed in 50 mL


2. Experimental
of aqueous solution containing 10 mg L1 of RhB. Prior to
2.1. Materials irradiation, the suspension was magnetically stirred, with a
Bi(NO3)3$5H2O, KI, urea, and ethanol were of AR grade and stirring rate of 1000 rpm, in dark for 30 min to achieve sorption
purchased from Aladdin (P. R. China). Commercial P25 TiO2 equilibrium of RhB on the catalyst surface. Approximately four
was purchased from Degussa (Germany). TBAI was of AR grade microliters aliquots of suspension were removed from the
Published on 08 July 2014. Downloaded by State University of New York at Stony Brook on 16/09/2014 11:30:11.

and purchased from Sinopharm (P. R. China). All chemicals reaction tube every 10 min, and centrifuged (7100  g) for 5 min
were used as received. Ultrapure water obtained from a Hitech- to remove particles. The pH values of suspensions were 7.3 
Kow water purication system (Hitech, P. R. China) was used 0.2 at the beginning and increased to 7.7  0.1 aer reactions.
throughout experiments. The RhB concentration was determined using a SP-1105 visible
spectrophotometer (Spectrum, P. R. China), by measuring the
2.2. Preparation of BiOI ower-like microspheres absorbance at 554 nm. The ratio of the RhB concentration to its
initial concentration (C/C0) was obtained by calculating the
0.1 g of Bi(NO3)3$5H2O in 20 mL of ethanol was stirred for ratio of the corresponding absorbances.
15 min, and the pH of the Bi(NO3)3 precursor was 1.42. 0.076 g
of TBAI in 20 mL of water was added dropwise under stirring.
The color of the suspension gradually changed from yellow to 3. Results and discussion
red, and the pH of the system changed to 2.13. Aer stirring for 3.1. XRD analysis
20 min, the mixture was transferred to an 80 mL polytetra-
Fig. 1 shows XRD patterns of the BiOI samples prepared at
uoroethylene-lined stainless steel autoclave, and heated to TH
different TH and tH. Retaining TH at 160  C while varying tH from
for different tH. The product was collected by ltration, washed
6 to 48 h, resulted in all samples exhibiting the characteristic
thoroughly with water, and dried at 60  C for 12 h. The prepared
peaks of highly crystalline BiOI (Fig. 1a). Peaks at 19.3, 29.6,
products were denoted as Bi-TH/tH. In addition, BiOI nanoplates
31.7, 37.1, 39.3, 45.5, 51.4, 55.2, 60.2, and 66.3 corresponded to
(denoted Bi-p) were prepared with KI as an iodine source via a
the (002), (012), (110), (013), (004), (014), (114), (212), (115), and
similar process (S1, ESI†). N-TiO2 was synthesized according to
(220) planes, respectively, and were indexed to the tetragonal
the literature35 (S2, ESI†).
phase of BiOI (JCPDS Card no. 10-0445). No peaks of impurities
were observed. Among these BiOI samples, Bi-160/30 exhibited
2.3. Characterization the highest intensity XRD peaks, suggesting it possessed the
Powder X-ray diffraction (XRD) patterns were recorded using a highest crystallinity. Bi-p exhibited a similar XRD pattern
D8 Advance diffractometer (Bruker, Germany), with Cu Ka (Fig. S1, ESI†). The effect of TH was investigated at a constant tH
radiation (l ¼ 1.54184 Å). Field-emission scanning electron
microscopy (SEM) images were collected using a Supra55
microscope (ZEISS, Germany). Transmission electron micros-
copy (TEM) and high-resolution TEM (HRTEM) images were
collected using a Jeol JEM-2100F microscope (Japan). Specic
surface areas (SBET) were determined by measuring volumetric
N2 adsorption–desorption isotherms at liquid nitrogen
temperature, using an ASAP 2020 HD88 instrument (Micro-
meritics, USA). Ultraviolet-visible diffuse reectance spectra
(DRS) were recorded on a Cary 100 spectrophotometer (Agilent,
USA) with a BaSO4 reference. Dynamic light scattering (DLS)
analyses were carried out at 25  C, using a Brookhaven Model
(BI-200SM) instrument equipped with an Ar laser lamp (300 mW
at 532 nm).

2.4. Photocatalytic activity measurements


The photocatalytic performance of the as-prepared catalysts
under visible light irradiation was evaluated by degrading RhB
at room temperature, using a XPA-7 photocatalytic reaction
apparatus (Xujiang Electromechanical Plant, P. R. China). A 400
W metal halide lamp was used as the light source, and was
equipped with an ultraviolet cutoff lter (l $ 420 nm) to provide
visible light. The distance between the lamp and glass tubes
containing dye solutions was 10 cm. A water lter around the
lamp was used to absorb infrared light. The reaction tubes Fig. 1 XRD patterns of BiOI samples prepared at (a) 160  C for tH and
around the lamp were soaked in a bath with temperature of (b) different TH for 30 h.

31394 | RSC Adv., 2014, 4, 31393–31399 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

of 30 h (Fig. 1b). Decreasing TH to 120  C or increasing to 180  C


(from the initial TH of 160  C) resulted in decreased crystallinity,
as indicated by decreased XRD peak intensities. Thus, the
optimal TH and tH for preparing highly crystalline BiOI were 160

C and 30 h, respectively.
In addition, the average crystallite sizes of the BiOI samples
Published on 08 July 2014. Downloaded by State University of New York at Stony Brook on 16/09/2014 11:30:11.

were calculated from the (012) peaks of the XRD patterns using
the Scherrer formula, and were in the range of 17.9–29.5 nm
(Fig. S2, ESI†). No obvious trend in average crystallite sizes was
observed with either TH or tH.

3.2. SEM and TEM analyses


Fig. 2 shows SEM and TEM images of the BiOI samples. Non-
uniform nanoplates were observed for Bi-p (Fig. 2a and c),
which had an average size and thickness of about 1–2 mm and
65.7 nm, respectively. Bi-160/30 exhibited a morphology con-
sisting of three-dimensional, hierarchical, ower-like micro-
spheres, with diameters of 2 mm (Fig. 2b and d). These ower-
like BiOI microspheres were constructed from nanosheets of
average thickness 26.4 nm (Fig. 2b). Insets in Fig. 2c and d
showed HRTEM images of Bi-p and Bi-160/30, respectively. The
observed lattice fringe spacing of 0.282 nm was corresponded to
(110) facets.
To investigate intermediates and the formation mechanism
of the ower-like microspheres, a time-evolution experiment
was conducted at TH of 160  C, to track the morphologic change
of BiOI crystals. The reagents and precursors were rst char-
acterized, so that the change in morphology of the products
could be better understood. The average hydrodynamic diam-
eter of TBAI micelles (or aggregates) in solution was 102 nm,
as measured by DLS (Fig. S3, ESI†). Bi(NO3)3 in ethanol

Fig. 3 SEM images of (a) Bi(NO3)3, (b) BiONO3, (c) Bi-25/0, (d) Bi-160/
1, (e) Bi-160/6, (f) Bi-160/12, (g) Bi-160/18, (h) Bi-160/24, (i) Bi-160/36,
(j) Bi-160/42, (k) Bi-160/48, (l) Bi-120/30, (m) Bi-140/30, and (n) Bi-
180/30.

Fig. 2 (a and b) SEM and (c and d) TEM images of (a and c) Bi-p and (b
and d) Bi-160/30. Insets show HRTEM images.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 31393–31399 | 31395
View Article Online

RSC Advances Paper

exhibited irregular ake particles (Fig. 3a). Mixing Bi(NO3)3 with


H2O resulted in a rapid reaction producing BiONO3.36 BiONO3
was prepared by mixing the ethanol dispersion of Bi(NO3)3 and
water, and exhibited disc-shaped particles of size 1.5 mm,
which appeared to consist of akes (Fig. 3b). Both the ethanol
dispersion of Bi(NO3)3 and the TBAI micelle solution were
Published on 08 July 2014. Downloaded by State University of New York at Stony Brook on 16/09/2014 11:30:11.

colorless. Mixing these two systems immediately produced a


brown dispersion. The SEM image of the resulting BiOI (Bi-25/0)
showed nanosheets of thickness 73.9 nm (Fig. 3c). This was
signicantly different from the morphologies of BiONO3 and
Bi(NO3)3 in ethanol (Fig. 3a and b). This suggested that the
formation of BiOI was very quick, which was supported by XRD
measurements (Fig. S4, ESI†). No characteristic reections of
BiONO3 were observed in the XRD patterns of any BiOI sample. Fig. 4 Average nanosheet thicknesses of BiOI samples.
Aer hydrothermal treatment at 160  C for 1 h, only irregular
BiOI nanosheets were observed (Fig. 3d). This suggested that no
obvious change in BiOI morphology occurred. When the tH was where larger sheets continued to grow and smaller ones dis-
increased to 6 h, rudiments of the ower-like microspheres were appeared. This may have been responsible for the partial
formed by the nanosheets (Fig. 3e). These ower-like rudiments destruction of the microsphere structure.
gradually grew with increasing tH (Fig. 3f). When tH was >18 h, All BiOI samples obtained at different TH and tH exhibited
integral ower-like microspheres were formed (Fig. 2b and 3g– hierarchical structures constructed from many nanosheets
i). Increasing tH to above 36 h caused some degree of destruc- (Fig. 2 and 3). The average thicknesses of a nanosheet in each
tion to the microsphere structure (Fig. 3j and k), which resulted BiOI sample (Hns) was measured from >50 nanosheets for each
in loose aggregates. The inuence of TH on the morphology of sample, and the results are shown in Fig. 4. Increasing tH at a
BiOI was also investigated at tH of 30 h. No integral ower-like constant TH of 160  C resulted in the Hns of the Bi-160/tH
microspheres were obtained at TH of 120, 140, and 180  C, as samples to gradually decrease and then increase. Bi-160/30
shown in Fig. 3l–n, respectively. This suggested that a TH of exhibited the lowest Hns. Samples prepared at higher and lower
160  C was suitable for forming the ower-like microspheres. TH exhibited increased Hns. The reason for the change in Hns
Based on the above discussion, a possible formation mech- with TH and tH has not yet been resolved.
anism of the ower-like microspheres is proposed, as illustrated
in Scheme 1. Mixing the ethanol dispersion of Bi(NO3)3 and
TBAI micelle solution caused the TBAI micelles to act as reac- 3.3. N2 sorption isotherms
tion centers. The reaction between I and Bi2O22+ formed BiOI N2 adsorption–desorption isotherms were measured to deter-
nuclei at the micelle surface. These nuclei gradually grew into mine the SBET of the BiOI-TH/tH and Bi-p samples, and the
nanosheets embedded in the micelles, to form aggregates. results are shown in Fig. 5. All isotherms were of type IV with H3
Dissolution-recrystallization occurred during hydrothermal hysteresis loops,37,38 which were characteristic of mesoporous
treatment, causing the nanosheets to gradually bond, forming materials. SBET (m2 g1) values were analyzed by the Brunauer–
ower-like rudiments and then compact microspheres. The Emmett–Teller (BET) method, from the adsorption branch of
ower-like BiOI microspheres were retained aer washing out the isotherms, and the results are also shown in Fig. 5. No
the TBAI cations. In this process, TBAI functioned as an iodine obvious trend in SBET was observed with increasing TH at
source and template. Excess tH resulted in Ostwald ripening, constant tH of 30 h, or with increasing tH at constant TH of

Scheme 1 Proposed formation mechanism of the BiOI flower-like microspheres.

31396 | RSC Adv., 2014, 4, 31393–31399 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances


Published on 08 July 2014. Downloaded by State University of New York at Stony Brook on 16/09/2014 11:30:11.

Fig. 5 N2 adsorption–desorption isotherms and SBET values of BiOI


samples.

Fig. 7 (a and b) Photocatalytic degradation of RhB, (c) pseudo first-


order kinetics fitting plots and (d) apparent first-order rate constants
for various photocatalysts.

compared with that of N-TiO2 and P25 TiO2. The results are
shown in Fig. 7a and b. No obvious degradation of RhB was
observed in the absence of a photocatalyst. Bi-160/30 exhibited
the highest photocatalytic efficiency among the tested catalysts.
Approximately 88% of RhB was degraded by Bi-160/30 aer
50 min.
The pseudo-rst-order kinetics model was used to t the
Fig. 6 Ultraviolet-visible diffuse reflectance spectra of BiOI samples.
photodegradation data, the equation for which is:
Insert shows (ahn)1/2 versus hn curves.
ln(Ct/C0) ¼ kt, (2)

160  C. The SBET of Bi-160/30 was 14.4 m2 g1, which was twice where k and t are the apparent rst-order rate constant and
more than that of Bi-p (5.5 m2 g1). irradiation time, respectively. The model tting plots and cor-
responding k values are shown in Fig. 7c and d, respectively. Bi-
3.4. Optical absorption properties 160/30 exhibited the highest k, which was 5, 7, and 20 times
The optical properties of the BiOI samples were investigated by higher than those of Bi-p, N-TiO2 and P25 TiO2, respectively.
ultraviolet-visible DRS, and are shown in Fig. 6. Bi-p and Bi-160/ The effect of pH on the photodegradation activity of Bi-160/
30 exhibited absorption edges up to 650 nm, which indicated 30 was studied (Fig. S5, ESI†). The degradation rate almost kept
a small band gap. An absorption peak at 450 nm was also constant at pH 3–7.5, but dramatically decreased at pH > 8. The
exhibited by both samples. The band gaps of crystalline semi- isoelectric point of the BiOI sample is pH 7.72. The decrease of
conductors can be calculated from their absorption spectra degradation rate at pH > 8 was probably ascribed to the
according to: decreased RhB adsorption on the surface of Bi-160/30 owing to
the high electrostatic repulsion between RhB and Bi-160/30.
ahv ¼ K(hv  Eg)n/2, (1)

where K, a, v, and Eg are the proportionality constant, absorp-


tion coefficient, light frequency, and band gap, respectively. n
depends on the transition characteristics of the semiconductor,
and values of 1 and 4 correspond to direct and indirect band-
gap semiconductors, respectively. BiOI has been shown to have
an n value of 4.39 Thus, the band gaps of Bi-p and Bi-160/30 were
determined to be about 1.81 eV (Fig. 6, inset), which are similar
to reported values.13

3.5. Photocatalytic activity


The photocatalytic performance of the BiOI samples was eval-
uated by degrading RhB under visible light irradiation, and was Fig. 8 Three consecutive RhB degradation cycles by Bi-160/30.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 31393–31399 | 31397
View Article Online

RSC Advances Paper


Published on 08 July 2014. Downloaded by State University of New York at Stony Brook on 16/09/2014 11:30:11.

Fig. 9 (a) XRD patterns and (b) SEM image of Bi-160/30 after three
consecutive photocatalytic degradations of RhB.

Fig. 11 Photodegradation efficiency of RhB over Bi-160/30, in the


Stability and reusability are important considerations for the presence of different scavengers: 20 mM t-BuOH, 0.2 mM sodium
use of photocatalysts in practical application. Fig. 8 shows oxalate, 10 mM K2Cr2O7, and 0.1 mM BQ.
cycling experiments for RhB degradation by Bi-160/30. Aer
three consecutive runs, the photocatalytic efficiency of Bi-160/
30 declined by <6%. The XRD patterns and SEM images showed thickness, and was referred to as the morphologic factor (Fm).
that the structure and morphology of Bi-160/30 remained Fm reected the effects of SBET and Hns on the photocatalytic
largely constant throughout the three cycles (Fig. 9). Thus, Bi- performance of the photocatalysts.
160/30 exhibited high stability and reusability during the pho- The photodegradation of organic pollutants may involve one
tocatalytic reactions. or more of the active species cOH, h+, cO2 and e.45 To investi-
Understanding the effect of a photocatalysts morphology on gate the contributions of these active species, different scaven-
its photocatalytic performance is of theoretical and practical gers were added to the photocatalytic degradation reaction, with
importance. Large specic surface areas and low lamella Bi-160/30 used as the model photocatalyst. Tert-butyl alcohol (t-
thicknesses for aky photocatalysts generally enhance photo- BuOH) was used as the scavenger for cOH,46 sodium oxalate for
catalytic performance.40,41 Large specic surface areas favor the h+,47 benzoquinone (BQ) for cO2,46 and potassium dichromate
adsorption of more organic pollutants, and provide more for e.48 Fig. 11 shows the effects of the scavengers on the pho-
reactive sites on the photocatalyst surface.42 Low lamella tocatalytic efficiency. In the presence of t-BuOH or K2Cr2O7, the
thicknesses can allow photogenerated electrons and holes to photocatalytic efficiency of Bi-160/30 was slightly reduced, indi-
transfer to the particle surface in a shorter distance, and thus cating that cOH and e were not signicant in the photo-
decrease the recombination of electrons and holes.43,44 The degradation process. When BQ or sodium oxalate was added to
amount of adsorbed RhB on the BiOI samples increased with the dispersion, the photocatalytic activity of Bi-160/30 decreased
increasing SBET (Fig. S6, ESI†). However, no obvious trend in k signicantly, indicating that cO2 and h+ played important roles.
was observed with either SBET or Hns (Fig. S7, ESI†). This was Therefore, cO2 and h+ were the main active species for RhB
because SBET and Hns were not constant for the BiOI samples. k degradation by BiOI, in agreement with previous reports.49
was found to correlate well with SBET/Hns, as shown in Fig. 10.
With increasing SBET/Hns, k gradually increased, and then
dramatically increased at SBET/Hns > 0.53 m2 (g1 nm1). SBET/
4. Conclusions
Hns represented the ratio of specic surface area to lamella Three-dimensional, hierarchical, ower-like BiOI microspheres
were hydrothermally prepared, using TBAI as an iodine source
and template for the rst time. The BiOI microspheres had an
average diameter of 2 mm, and consisted of BiOI nanosheets.
The photocatalytic performance of the BiOI samples was inu-
enced by their specic surface area and nanosheet thickness. Fm
was used to describe the photodegradation performance of the
BOI samples, with photocatalytic efficiency increasing with
increasing Fm. Bi-160/30 exhibited the highest photodegradation
efficiency for RhB under visible light irradiation, which was ve
and seven times more than those of Bi-p and N-TiO2, respectively.
Bi-160/30 exhibited good stability and reusability.

Acknowledgements
Fig. 10 Dependence of rate constant and degraded RhB percentage The present work is supported nancially by the National
on Fm. Natural Science Foundation of China (no. 21173135) and the

31398 | RSC Adv., 2014, 4, 31393–31399 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

Specialized Research Fund for the Doctoral Program of Higher 24 X. Xiao and W. D. Zhang, J. Mater. Chem., 2010, 20, 5866–
Education of China (no. 20110131130008). 5870.
25 B. Zhang, G. B. Ji, M. A. Gondal, Y. S. Liu, X. M. Zhang,
X. F. Chang and N. W. Li, J. Nano. Res., 2013, 15, 1–9.
References 26 J. Xia, S. Yin, H. M. Li, H. Xu, L. Xu and Q. Zhang, Colloids
Surf., A, 2011, 387, 23–28.
Published on 08 July 2014. Downloaded by State University of New York at Stony Brook on 16/09/2014 11:30:11.

1 K. Maeda, T. Takata, M. Hara, N. Saito, Y. Inoue, 27 K. Ren, K. Zhang, J. Liu, H. Luo, Y. Huang and X. Yu,
H. Kobayashi and K. Domen, J. Am. Chem. Soc., 2005, 127, CrystEngComm, 2012, 14, 4384–4390.
8286–8287. 28 S. X. Wu, J. Z. Fang, X. T. Hong, K. S. Hui and Y. F. Chen,
2 A. Hameed, T. Montini, V. Gombac and P. Fornasiero, J. Am. Dalton Trans., 2014, 43, 2611–2619.
Chem. Soc., 2008, 130, 9658–9659. 29 J. Di, J. X. Xia, S. Yin, H. Xu, L. Xu, Y. G. Xu and H. M. Li,
3 H. Yamashita, M. Harada, J. Misaka, M. Takeuchi, K. Ikeue J. Mater. Chem. A, 2014, 2, 5340–5351.
and M. Anpo, J. Photochem. Photobiol., A, 2002, 148, 257–261. 30 N. Huang, A. Hu and M. Li, Mater. Lett., 2013, 109, 247–252.
4 E. Bae and W. Choi, Environ. Sci. Technol., 2003, 37, 147–152. 31 Z. K. Cui, M. M. Si, Z. Zheng, L. W. Mi, W. J. Fa and H. M. Jia,
5 H. M. Jia, W. J. Xiao, L. Z. Zhang, Z. Zheng, H. L. Zhang and Catal. Commun., 2013, 42, 121–124.
F. Deng, J. Phys. Chem. C, 2008, 112, 11379–11384. 32 D. K. Ma, S. M. Zhou, X. Hu, Q. R. Jiang and S. M. Huang,
6 K. Vinodgopal and P. V. Kamat, Environ. Sci. Technol., 1995, Mater. Chem. Phys., 2013, 140, 11–15.
29, 841–845. 33 J. Cao, X. Li, H. L. Lin, S. F. Chen and X. Fu, J. Hazard. Mater.,
7 T. Oncescu, M. I. Stefan and P. Oancea, Environ. Sci. Pollut. 2012, 239, 316–324.
Res., 2010, 17, 1158–1166. 34 J. Cao, B. Y. Xu, B. D. Luo, H. L. Lin and S. F. Chen, Catal.
8 M. Kong, Y. Li, X. Chen, T. Tian, P. Fang, F. Zheng and Commun., 2011, 12, 660–664.
X. Zhao, J. Am. Chem. Soc., 2011, 133, 16414–16417. 35 D. Mitoraj and H. Kisch, Angew. Chem., Int. Ed., 2008, 47,
9 S. G. Kumar and L. G. Devi, J. Phys. Chem. A, 2011, 115, 9975–9978.
13211–13241. 36 P. V. Ioannou, Main Group Chem., 2011, 10, 255–264.
10 Y. N. Wang, K. J. Deng and L. J. Deng, J. Phys. Chem. C, 2011, 37 A. Kay and I. Cesar, J. Am. Chem. Soc., 2006, 128, 15714–
115, 14300–14308. 15721.
11 R. Hao, X. Xiao, X. Zuo, J. Nan and W. Zhang, J. Hazard. 38 Y. Li, J. Liu, X. Huang and G. Li, Cryst. Growth Des., 2007, 7,
Mater., 2012, 209, 137–145. 1350–1355.
12 Y. Lei, G. Wang, S. Song, W. Fan and H. Zhang, 39 H. F. Cheng, B. B. Huang, Y. Dai, X. Y. Qin and X. Y. Zhang,
CrystEngComm, 2009, 11, 1857–1862. Langmuir, 2010, 26, 6618–6624.
13 S. X. Ge and L. Z. Zhang, J. Nano Res., 2012, 14, 1–11. 40 H. F. Cheng, B. B. Huang and Y. Dai, Nanoscale, 2014, 6,
14 Y. Y. Li, J. S. Wang, H. C. Yao, L. Y. Dang and Z. J. Li, J. Mol. 2009–2026.
Catal. A: Chem., 2011, 334, 116–122. 41 S. X. Yang, Q. Yue, Z. H. Chen, J. H. Yang and J. B. Li, J. Alloys
15 T. B. Li, G. Chen, C. Zhou, Z. Y. Shen, R. C. Jin and J. X. Sun, Compd., 2014, 597, 91–94.
Dalton Trans., 2011, 40, 6751–6758. 42 J. Xia, S. Yin, H. Li, H. Xu, Y. Yan and Q. Zhang, Langmuir,
16 L. Q. Ye, Y. R. Su, X. L. Jin, H. Q. Xie and C. Zhang, 2010, 27, 1200–1206.
Environmental Science: Nano, 2014, 1, 90–112. 43 Z. Zhang, C. C. Wang, R. Zakaria and J. Y. Ying, J. Phys. Chem.
17 L. Q. Ye, L. H. Tian, T. Y. Peng and L. Zan, J. Mater. Chem., B, 1998, 102, 10871–10878.
2011, 21, 12479–12484. 44 N. J. Huo, Q. Yue, S. X. Yang and J. B. Li, ChemPhysChem,
18 L. Ye, J. Chen, L. Tian, J. Liu, T. Peng, K. Deng and L. Zan, 2013, 14, 4069–4073.
Appl. Catal., B, 2013, 130, 1–7. 45 J. Ryu and W. Choi, Environ. Sci. Technol., 2004, 38, 2928–
19 W. L. Huang and Q. Zhu, J. Comput. Chem., 2009, 30, 183– 2933.
190. 46 L. Mohapatra, K. Parida and M. Satpathy, J. Phys. Chem. C,
20 J. Hu, S. Weng, Z. Zheng, Z. Pei, M. Huang and P. Liu, 2012, 116, 13063–13070.
J. Hazard. Mater., 2014, 264, 293–302. 47 X. Xiao, R. Hu, C. Liu, C. Xing, C. Qian, X. Zuo and L. Wang,
21 X. Chang, J. Huang, Q. Tan, M. Wang, G. Ji, S. Deng and Appl. Catal., B, 2013, 140, 433–443.
G. Yu, Catal. Commun., 2009, 10, 1957–1961. 48 S. Ge and L. Zhang, Environ. Sci. Technol., 2011, 45, 3027–
22 J. Jiang, X. Zhang, P. Sun and L. Zhang, J. Phys. Chem. C, 3033.
2011, 115, 20555–20564. 49 H. Lee and W. Choi, Environ. Sci. Technol., 2002, 36, 3872–
23 X. Zhang, Z. Ai, F. Jia and L. Zhang, J. Phys. Chem. C, 2008, 3878.
112, 747–753.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 31393–31399 | 31399

You might also like