Enhancement of photocatalytic decomposition of perfluorooctanoic acid on CeO2/In2O3

Fang Jianga, Haitao Zhaoa, Huan Chen*a, Chenmin Xua and Jian Chenb
aKey Laboratory of Jiangsu Province for Chemical Pollution Control and Resources Reuse, School of Environmental and Biological Engineering, Nanjing University of Science and Technology, Nanjing 210094, China. E-mail: hchen404@njust.edu.cn; Fax: +86-25-84315352; Tel: +86-25-84303209
bYancheng Teachers College, Jiangsu Provincial Key Laboratory of Coastal Wetland Bioresources and Environmental Protection, Yancheng 224002, PR China

Received 16th April 2016 , Accepted 24th July 2016

First published on 25th July 2016


Abstract

CeO2-doped indium oxide photocatalysts (xCeO2/In2O3) with various CeO2 doping amounts were synthesized and systematically characterized by X-ray powder diffraction (XRD), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), UV-vis spectra, photoluminescence spectroscopy (PL) and photocurrent measurements. The as-obtained xCeO2/In2O3 photocatalysts were used in photocatalytic degradation of PFOA for the first time and much higher photocatalytic activity of xCeO2/In2O3 than CeO2 and In2O3 was obtained under UV light irradiation. The excellent photocatalytic activity of 0.86% CeO2/In2O3 could be mainly derived from effective inhibition of recombination of photo-induced electron-holes caused by the charge transfer between CeO2 and In2O3. Moreover, the 0.86% CeO2/In2O3 catalyst showed excellent photocatalytic stability in cycling runs for the photocatalytic degradation of PFOA, providing a great potential of CeO2/In2O3 catalysts for PFOA treatment.


1. Introduction

Perfluorooctanoic acid (PFOA, C7F15COOH), as a class of perfluorocarboxylic acids (PFCAs), is widely used as a surface treatment agent in manufacturing, aerospace, automotives, electronics, semiconductors and textiles.1 The chemical structure of PFOA (Fig. S1) shows that the PFOA is chemically inert due to its strong C–F bonds (530 kJ mol−1).2 As a result, PFOA is extraordinarily thermally and chemically stable. In recent years, many studies have showed that PFOA is environmentally persistent and bioaccumulative,3,4 and is frequently detected in waters, animals and even humans beings.5–7 Furthermore, toxicological studies demonstrate that exposure to PFOA can lead to developmental and reproductive toxicities, liver damage and possibly cancer.8,9 Therefore, developing effective methods to eliminate PFOA from the environment is urgent.

It has been found that natural decomposition processes and many conventional techniques, including biological degradation, oxidation and reduction, are not effective to destruct PFOA under mild conditions.10,11 In the last few years, several treatment methods have been developed to decompose PFOA, such as electrochemical treatment,12,13 sono-chemical methods,14–16 hydrothermal treatment17,18 and photocatalysis.19–22 Among these methods, UV-based photocatalysis have been demonstrated to be an effective method for PFOA decomposition under mild conditions.23 In the previous reports, TiO2 has been widely investigated to degrade most organic pollutants into less harmful products by hydroxyl radicals (OH˙) formed under ultraviolet (UV) irradiation. However, TiO2 has low photocatalytic activity to decompose PFOA.24–28 Recently, it is reported that β-Ga2O3[thin space (1/6-em)]29–31 and In2O3[thin space (1/6-em)]32–35 showed effective activities for PFOA decomposition under UV irradiation and the degradation production was found to be CO2 and F.

Indium oxide (In2O3), with a direct band gap of 3.6 eV and an indirect band gap of 2.8 eV,36 has been widely applied in the development of semiconductor gas sensors, optoelectronic devices, solar cells and photocatalysis.37,38 Moreover, In2O3 has proved to be an efficient catalyst to decompose PFOA under UV irradiation. PFOA could be decomposed on In2O3 via direct oxidation because of high adsorption capacity of In2O3 and tight coordination with PFOA.32 Moreover, the high redox potential of hole in In2O3 is suitable for oxidation of PFOA. However, the fast recombination of photo-generated electrons and holes limited the photocatalytic performance of In2O3. Therefore, considerable attempts have been used to modify In2O3 nanostructure and improve the photocatalytic activity for PFOA decomposition. For instance, In2O3 nanostructures with different morphologies, such as porous microspheres, nanocubes and nanoplates, have been developed to PFOA decomposition and high photocatalytic activities were found in In2O3 material with higher oxygen vacancy defects.33 Moreover, In2O3–graphene composites were synthesized and showed an enhanced photocatalytic activity for PFOA decomposition because of the efficient separation of photogenerated hole–electron pairs.39 To take better advantage of In2O3, it is crucial to explore effective modification methods to further optimize its photocatalytic activity toward PFOA degradation.

Cerium oxide (CeO2), a well-known rare earth oxide, has been widely used in catalysis, electrochemistry, photochemistry and materials science due to its proper electronic structure and high stability.40–42 It has been investigated that CeO2 can be used as an efficient cocatalyst because of its wide band gap (Eg = 2.92 eV), which can absorb light near UV region.43 For example, Liu et al.44 reported the efficient photocatalytic activity of CeO2/TiO2 composite on degradation of methylene blue. Wang et al.45 found that ZrO2/CeO2 nanocomposite possessed excellent photocatalytic activity toward RhB decomposition. And Song et al.46 also found that CeO2/Ag3PO4 composite exhibited enhanced photocatalytic activity to decompose methylene blue and RhB. Thus, it was proposed that CeO2 could also be used as cocatalyst for In2O3 to form CeO2-doped In2O3 composite. Although there is no report using CeO2-doped In2O3 composite as photocatalyst so far, the CeO2–In2O3 composite has been used as an effective gas sensor. As a sensor, a charge transfer from In2O3 to CeO2 has been found in the CeO2–In2O3 composite.47 Therefore it is reasonable to speculate that CeO2–In2O3 composite can benefit the electron transfer and promote the separation of the electron–hole pairs, leading to an enhanced photocatalytic activity on PFOA degradation.

In this study, CeO2-doped In2O3 composite (xCeO2/In2O3) with different loading amounts of CeO2 were synthesized and used for PFOA decomposition under UV light irradiation for the first time. The effect of CeO2 on photocatalytic activity of xCeO2/In2O3 composite was investigated and the possible photocatalytic mechanism was discussed. The results revealed that the CeO2/In2O3 composite possessed a much higher photocatalytic activity to decompose PFOA as compared to CeO2 and In2O3.

2. Experimental

2.1 Preparation of catalyst

Indium oxide (In2O3, 99.99%) was obtained from Aladdin Chemical Reagent Co. The CeO2/In2O3 catalysts with different CeO2 loading amounts were prepared by the impregnation method. In detail, 1 g of In2O3 was impregnated in Ce(NO3)3 solution, and the mixture was evaporated to dryness under stirring at 80 °C. The obtained products were dried overnight and calcined at 600 °C for 4 h in the oven. The final products were donated as xCeO2/In2O3, where x was the weight percent of CeO2 as detected by ICP-AES.

2.2 Characterization of catalysts

Powder X-ray diffraction (XRD) patterns were obtained from a Bruker D8 advanced diffraction-meter with Cu Kα radiation and the scanning angle ranging from 10° to 80°. Transmission electron microscopy (TEM) images were acquired using JEOL JEM2100 microscopy at an acceleration voltage of 200 kV. N2 adsorption–desorption isotherm was measured on a BEL Belsorp-MAX adsorption analyzer. The samples were outgassed at 280 °C for 6 h prior to the nitrogen adsorption measurement. The specific surface areas of the catalysts were calculated by the BJH model. X-ray photoelectron spectra (XPS) were obtained on a RBD upgraded PHI-5000C ESCA system (Perkin Elmer) with Mg Kα radiation ( = 1253.6 eV). UV-vis diffuse reflectance spectroscopy was taken on a Hitachi U-3010 UV-vis spectrometer. Photoluminescence spectra were carried out on a jobinYvon SPEX Fluorolog-3-P spectroscope. Photocurrent was obtained on a CHI 660B electrochemical workstation in a standard three-electrode system. The zeta potential was analyzed using a Brookhaven zeta PALS.

2.3 Photocatalysis experiments

The photocatalytic decomposition of PFOA was conducted in a quartz tube reactor at 25 °C. In a typical experiment, 0.08 g catalyst was added into 200 mL of 100 mg L−1 PFOA solution. The mixture was kept stirring in dark for 30 min to reach adsorption–desorption equilibrium before a 500 W Hg lamp was turned on. At given time intervals of irradiation, 2 mL of suspension was withdrawn and then filtered. The concentrations of PFOA in the filtrate were determined by a high performance liquid chromatography (e2695, Waters) equipped with a 432 conductivity detector. The reaction products were separated on an X-Bridge C18 column (4.6 × 250 mm). The mobile phase was the mixture of methanol/0.02 M NaH2PO4 (75/25, v/v) at a flow rate of 1.0 mL min−1 and the injected sample volume was 100 μL. The concentration of fluoride ion (F) was determined by an ICS-2000 Dionex ion chromatography. The total organic carbon (TOC) was measured using an Elementar vario TOC analyzer.

3. Results and discussion

3.1 Catalyst characterization

Fig. 1 depicts the XRD patterns of In2O3 and xCeO2/In2O3 catalysts. For In2O3, the diffraction peaks appeared at 21.42°, 30.60°, 35.59°, 45.50°, 50.96° and 60.76° were correspond to the (211), (222), (400), (134), (440) and (622) planes of In2O3, respectively (JCPDS no. 65-3170). All of these peaks could be seen in the XRD patterns of xCeO2/In2O3, indicating that loading of CeO2 did not destroy the crystal structure of In2O3. Interestingly, the sharper peaks could be observed in xCeO2/In2O3 as compared to In2O3, revealing that fine crystallite of In2O3 was obtained in xCeO2/In2O3. During the preparation process of xCeO2/In2O3, the catalysts were calcined at 600 °C, which may influence the crystallinity of In2O3. Thus, XRD pattern of In2O3 after calcined at 600 °C was also measured and it was found that the calcined In2O3 showed better crystallinity than In2O3 (Fig. S2), indicating that calcination at high temperature was beneficial to crystal formation. In addition, no diffraction peaks of CeO2 species were observed, because the CeO2 particles were dispersed well on the surface of the xCeO2/In2O3 catalysts.
image file: c6ra09856h-f1.tif
Fig. 1 XRD patterns of In2O3 and CeO2/In2O3 samples with different CeO2 contents.

BET surface areas of In2O3 and xCeO2/In2O3 were calculated from the results of N2 adsorption–desorption isotherms. The BET surface area of In2O3 was 39.7 m2 g−1, while the surfaces areas of xCeO2/In2O3 (25.7–32.1 m2 g−1) were slightly smaller than that of In2O3. It revealed that the introduction of CeO2 may slightly reduce the BET specific surface of In2O3, for some pores of In2O3 were blocked by CeO2 particles.

TEM characterization was used to investigate the microstructure of xCeO2/In2O3 catalysts. Uniform distribution of CeO2/In2O3 nanoparticles with porous structure could be seen from TEM image (Fig. 2a) of 2.33% CeO2/In2O3. The average particle size of CeO2/In2O3 is about 30–40 nm. The high-resolution (HRTEM) image of 2.33% CeO2/In2O3 is shown in Fig. 2b. The basal distances of the lattice fringes are calculated to be about 0.15 nm and 0.29 nm, matching well with the lattice spacing of (622) and (211) planes of In2O3, respectively. And the lattice fringes of 0.27 nm is corresponded to the interplanar spacing of (200) planes of CeO2. It demonstrated that the conjunctions were well-formed without changing the crystal of CeO2 and In2O3.


image file: c6ra09856h-f2.tif
Fig. 2 (a) TEM image of 2.33% CeO2/In2O3 and (b) HRTEM image of 2.33% CeO2/In2O3.

XPS experiments were carried out to examine the surface component and chemical state of In2O3 and 0.86% CeO2/In2O3 catalysts. As depicted in Fig. 3a, the XPS spectra of In3d measured from pure In2O3 were spilt into double peaks centered at binding energies of 443.7 and 451.3 eV, respectively. Compared with In2O3, 0.86% CeO2/In2O3 showed slight shift to high binding energy, possibly owing to the electron transport from In2O3 to CeO2. Fig. 3b shows the XPS spectra of O1s in In2O3 and 0.86% CeO2/In2O3. The two peaks in 443.9 and 531.6 eV were corresponded to the lattice oxygen (O′) and chemisorbed oxygen (O′′) in crystalline In2O3, respectively. The ratio of lattice oxygen (O′/(O′ + O′′)) in 0.86% CeO2/In2O3 was larger than that of In2O3, indicating the CeO2 doping could improve the degree of crystallinity of In2O3. The result was consistent with the results of XRD.


image file: c6ra09856h-f3.tif
Fig. 3 XPS spectra of (a) In3d and (b) O1s for In2O3 and 0.86% CeO2/In2O3.

The photoelectrochemical properties of CeO2/In2O3 were clarified through UV-visible absorption spectroscopy, photoluminescence spectroscopy (PL) and transient photocurrent response measurements. As shown in UV-vis absorption spectra (Fig. 4a), 0.86% CeO2/In2O3 exhibited higher absorption intensity in the UV light region as compared to pure In2O3. The absorption of 0.86% CeO2/In2O3 showed a slight red-shift than that of In2O3. The PL spectra (excited in 250 nm) of In2O3 and 0.86% CeO2/In2O3 are shown in Fig. 4b. In comparison of In2O3, the 0.86% CeO2/In2O3 catalyst exhibited a lower emission intensify, indicating the recombination of the photogenerated charge carrier was inhibited. Moreover, the results of transient photocurrent responses (Fig. 4c) showed that the photocurrent of 0.86% CeO2/In2O3 (5.35 μA) was much higher than that of In2O3 (2.32 μA). The phenomena suggested that doping CeO2 could improve the separation efficiency of photogenerated electrons and holes, probably due to the interfacial charge transformation from In2O3 to CeO2.


image file: c6ra09856h-f4.tif
Fig. 4 (a) UV-vis absorption spectra; (b) photoluminescence spectra and (c) transient photocurrent responses of In2O3 and 0.86% CeO2/In2O3.

3.2 Photocatalytic decomposition of PFOA

Fig. 5a shows the photocatalytic decomposition process of aqueous PFOA (100 mg L−1) in the presence of 0.08 g of xCeO2/In2O3 catalyst under UV light irradiation. The decomposition efficiencies in 60 min were 43.2%, 53% and 88% for P25, CeO2 and In2O3, respectively, while the xCeO2/In2O3 catalysts possessed higher degradation efficiency. The control experiment without light irradiation is exhibited in Fig. S3. It is clear that little perfluorooctanoic acid has been adsorbed by xCeO2/In2O3, indicating adsorption is not the main procedure in the photocatalytic decomposition process.
image file: c6ra09856h-f5.tif
Fig. 5 (a) Photocatalytic degradation of PFOA over P25, CeO2, In2O3 and xCeO2/In2O3 catalysts and (b) the pseudo-first-order rate constant (kobs) for the photocatalytic degradation of PFOA by P25, CeO2, In2O3 and xCeO2/In2O3 catalysts.

The kinetic of the photocatalytic decomposition of PFOA could fit well to the pseudo-first-order model, which could be described as follows:

 
ln(C0/Ct) = kt (1)
where k is the rate constant (min−1), C0 and Ct are the concentrations of PFOA aqueous solution at irradiation times of 0 and t min, respectively. As described in Fig. 5b, the rate constants for PFOA decomposition over P25, CeO2, In2O3, 0.61% CeO2/In2O3, 0.86% CeO2/In2O3, 1.34% CeO2/In2O3 and 2.33% CeO2/In2O3 were found to be 0.54 h−1, 0.72 h−1, 2.10 h−1, 2.22 h−1, 3.78 h−1, 2.82 h−1 and 2.4 h−1, respectively. It can be concluded that the activities of xCeO2/In2O3 catalysts depended on the composition of CeO2 since a gradual increase in CeO2 content led to an enhancement of decomposition efficiencies with 0.86% CeO2/In2O3 exhibiting the highest photocatalytic activity. Further increase in CeO2 content led to a decrease of decomposition efficiency, probably due to the agglomeration of CeO2 caused a weaker UV light harvesting. The defluorination curve of perfluorooctanoic acid (PFOA) with 0.86% CeO2/In2O3 is displayed in Fig. S4a. The defluorination ratio of PFOA is calculated by the formulation:
 
image file: c6ra09856h-t1.tif(2)
where CF is the molar concentration of fluoride ion, and C0 is initial molar concentration of PFOA. It can be observed that the defluorination ratio of PFOA is 53.3% in 60 min. Fig. S4b depicts the TOC removal curve of perfluorooctanoic acid (PFOA) with 0.86% CeO2/In2O3, in which 54.8% of the PFOA was effectively mineralized after irradiation for 60 min. The result confirmed that PFOA was decomposed to inorganic production and F, and the mechanism was discussed below.

In this study, xCeO2/In2O3 catalysts displayed much higher photocatalytic activities as compared to CeO2 and In2O3, due to a synergistic effect between CeO2 and In2O3 in xCeO2/In2O3 catalysts. As known, the conduction band (CB) bottom and the valence band (VB) top of In2O3 lie at −0.63 and 2.17 eV versus NHE, respectively.48 Meanwhile, the CB bottom and the VB top of CeO2 lie at −0.39 and 2.50 eV versus NHE, respectively.49 Thus, charge transfer could occur between CeO2 and In2O3 for the interfacial electric field under the UV light irradiation. The photoexcited electrons on the CB of In2O3 could easily transfer to the CB of CeO2. At the same time, photoinduced holes in VB of CeO2 transferred to In2O3. The charge transfer process allowed CeO2 to scavenge electrons to promote the separation of charge carriers. Moreover, the lifetime of charge carriers related to the radiative process was also prolonged for xCeO2/In2O3, and thus improved the photocatalytic activity of decomposition of PFOA. It has been described by Panchangam et al.50 that PFOA was decomposed in a stepwise way.

Firstly, an electron of C7F15COOads was attracted by hvb+, producing C7F15COO˙ radical:

 
hvb+ + C7F15COOads → C7F15COO˙ (3)

Then this unstable C7F15COO˙ radical was transformed to C7F15 radical after Kolbe decarboxylation reaction:

 
C7F15COO˙ → C7F15 + CO2 (4)

The C7F15 radical reacted with ˙OH radicals to form thermodynamically unstable C7F15OH. And C7F15OH underwent hydrolysis and HF elimination, which formed C6F13COO with the loss of CF2:

 
C7F15OH + H+ → C6F13COF + 2HF (5)
 
C6F13COF + ˙OH → C6F13COO + HF (6)

Similarly, C6F13COO was decomposed into C5F11COO by repeating this process. PFOA was finally mineralized to CO2 and fluoride ions in the stepwise manner. The proposed PFOA decomposition mechanism by CeO2/In2O3 is exhibited in Fig. 6. The main oxidative species in the photocatalytic reaction were further investigated by adding radical scavengers, using KI, tert-butanol (t-BuOH) and benzoquinone (BQ) as hole, hydroxyl radical and superoxide radical scavengers, respectively. As shown in Fig. 7, the photocatalytic reaction of PFOA was greatly inhibited by adding KI and BQ, while slight inhibited by adding t-BuOH, revealing that photogenerated holes and superoxide radicals were the main active species in the photocatalytic process.


image file: c6ra09856h-f6.tif
Fig. 6 Proposed photocatalytic mechanism of PFOA degradation over 0.86% CeO2/In2O3 under UV light irradiation.

image file: c6ra09856h-f7.tif
Fig. 7 Effects of scavengers on PFOA degradation over 0.86% CeO2/In2O3.

3.3 Effect of pH

Fig. 8a exhibits the impact of initial pH on the catalytic decomposition of PFOA over 0.86% CeO2/In2O3. With the initial pH increasing from 2.84 to 9.51, the photocatalytic degradation efficiency of PFOA gradually decreased from 100% to 54.9%. Results of zeta potentials of 0.86% CeO2/In2O3 (Fig. 8b) indicated that zeta potentials continuously decreased with the increasing of solution pH and the points of zero charge (pHpzc) of 0.86% CeO2/In2O3 was 3.8. PFOA in the aqueous solution are in an anionic form in the range of tested pH because the pKa of PFOA is around −0.1.51 When the solution pH values were lower than 3.8, the surfaces of the 0.86% CeO2/In2O3 were positively charged owing to the protonation. Herein, PFOA could strongly interact with catalyst by electrostatic interaction, resulting in a high PFOA removal efficiency. At pH 3.8–9.51, the negative charges on the surfaces of 0.86% CeO2/In2O3 were observed due to the deprotonation, leading to an electrostatic repulsion between PFOA and 0.86% CeO2/In2O3. Thus, the degradation efficiency of PFOA on 0.86% CeO2/In2O3 decreased sharply with the increase of solution pH.
image file: c6ra09856h-f8.tif
Fig. 8 (a) Influence of the initial pH on PFOA degradation by 0.86% CeO2/In2O3 sample and (b) zeta potentials of 0.86% CeO2/In2O3.

3.4 Stability of photocatalyst

In addition to photocatalytic activity, the stability of photocatalysts is another key point for their practical applications. Therefore, the stability of 0.86% CeO2/In2O3 was investigated by recycling experiments to decompose PFOA. As illustrated in Fig. 9, the catalytic efficiency of 0.86% CeO2/In2O3 decreased slightly after recycling tests for 4 times. XRD patterns of the 0.86% CeO2/In2O3 samples before and after photocatalysis were measured and the results are shown in Fig. S5. It revealed that the crystal structure of 0.86% CeO2/In2O3 had no change after photocatalytic reaction, reflecting that the 0.86% CeO2/In2O3 photocatalyst had a remarkable stability, which could greatly favor the practical application to eliminate the PFOA from wastewater.
image file: c6ra09856h-f9.tif
Fig. 9 Cycling runs for the photocatalytic degradation of PFOA over 0.86% CeO2/In2O3.

4. Conclusions

In summary, CeO2-doped In2O3 catalysts with different amounts of CeO2 (xCeO2/In2O3) were fabricated via the conventional impregnation method. The catalysts possessed higher photocatalytic activities than pure CeO2 and In2O3 for the PFOA decomposition under UV light irradiation. The remarkably increased photocatalytic activity of xCeO2/In2O3 could be ascribed to the complementary heterostructure between CeO2 and In2O3, which promoted the separation of charge carriers and prolonged the lifetime of charge carriers. Furthermore, the 0.86% CeO2/In2O3 exhibited excellent stability in cycling tests, demonstrating its promising application in the degradation of PFOA from aqueous phase.

Acknowledgements

The financial supports from the Natural Science Foundation of China (No. 51178223, 51208257, 51478223 and 21307103) and the Natural Science Foundation of Jiangsu Province (BK2012405), China Postdoctoral Science Foundation (No. 2013M541677), Jiangsu Planned Projects for Postdoctoral Research Funds (1202007B) and the Fundamental Research Funds for the Central University (No. 30915011308) are gratefully acknowledged.

References

  1. K. Prevedouros, I. T. Cousins, R. C. Buck and S. H. Korzeniowski, Environ. Sci. Technol., 2006, 40, 32–44 CrossRef CAS PubMed.
  2. C. D. Vecitis, H. Park, J. Cheng, B. T. Mader and M. R. Hoffmann, J. Phys. Chem. C, 2008, 112, 4261–4270 CAS.
  3. J. P. Giesy and K. Kannan, Environ. Sci. Technol., 2002, 36, 146–152 CrossRef.
  4. J. P. Giesy and K. Kannan, Environ. Sci. Technol., 2001, 35, 1339–1342 CrossRef CAS PubMed.
  5. J. P. Benskin, L. W. Yeung, N. Yamashita, S. Taniyasu, P. K. Lam and J. W. Martin, Environ. Sci. Technol., 2010, 44, 9049–9054 CrossRef CAS PubMed.
  6. H. M. Shin, V. M. Vieira, P. B. Ryan, R. Detwiler, B. Sanders, K. Steenland and S. M. Bartell, Environ. Sci. Technol., 2011, 45, 1435–1442 CrossRef CAS PubMed.
  7. T. Zhang, Q. Wu, H. W. Sun, X. Z. Zhang, S. H. Yun and K. Kannan, Environ. Sci. Technol., 2010, 44, 4341–4347 CrossRef CAS PubMed.
  8. J. H. Yang, Chemosphere, 2010, 81, 548–552 CrossRef CAS PubMed.
  9. G. Zhao, J. Wang, X. Wang, S. Chen, Y. Zhao, F. Gu, A. Xu and L. Wu, Environ. Sci. Technol., 2010, 45, 1638–1644 CrossRef PubMed.
  10. E. Sinclair and K. Kannan, Environ. Sci. Technol., 2006, 40, 1408–1414 CrossRef CAS PubMed.
  11. M. M. Schultz, C. P. Higgins, C. A. Huset, R. G. Luthy, D. F. Barofsky and J. A. Field, Environ. Sci. Technol., 2006, 40, 7350–7357 CrossRef CAS PubMed.
  12. Q. Zhuo, S. Deng, B. Yang, J. Huang and G. Yu, Environ. Sci. Technol., 2011, 45, 2973–2979 CrossRef CAS PubMed.
  13. T. Ochiai, Y. Iizuka, K. Nakata, T. Murakami, D. A. Tryk, A. Fujishima, Y. Koide and Y. Morito, Diamond Relat. Mater., 2011, 20, 64–67 CrossRef CAS.
  14. H. Moriwaki, Y. Takagi, M. Tanaka, K. Tsuruho, K. Okitsu and Y. Maeda, Environ. Sci. Technol., 2005, 39, 3388–3392 CrossRef CAS PubMed.
  15. L. A. P. Thi, H. T. Do and S. L. Lo, Ultrason. Sonochem., 2014, 21, 1875–1880 CrossRef PubMed.
  16. J. C. Lin, S. L. Lo, C. Y. Hu, Y. C. Lee and J. Kuo, Ultrason. Sonochem., 2015, 22, 542–547 CrossRef CAS PubMed.
  17. Y. C. Lee, S. L. Lo, P. T. Chiueh and D. G. Chang, Water Res., 2009, 43, 2811–2816 CrossRef CAS PubMed.
  18. Y. C. Lee, S. L. Lo, P. T. Chiueh, Y. H. Liou and M. L. Chen, Water Res., 2010, 44, 886–892 CrossRef CAS PubMed.
  19. M. J. Chen, S. L. Lo, Y. C. Lee, J. Kuo and C. H. Wu, J. Hazard. Mater., 2016, 303, 111–118 CrossRef CAS PubMed.
  20. J. H. Cheng, X. Y. Liang, S. W. Yang and Y. Y. Hu, Chem. Eng. J., 2014, 239, 242–249 CrossRef CAS.
  21. M. Li, Z. Yu, Q. Liu, L. Sun and W. Huang, Chem. Eng. J., 2016, 286, 232–238 CrossRef CAS.
  22. Y. Wang and P. Zhang, J. Environ. Sci., 2014, 26, 2207–2214 CrossRef PubMed.
  23. H. Hori, E. Hayakawa, H. Einaga, S. Kutsuna, K. Koike, T. Ibusuki, H. Kiatagawa and R. Arakawa, Environ. Sci. Technol., 2004, 38, 6118–6124 CrossRef CAS PubMed.
  24. C. Song, P. Chen, C. Wang and L. Zhu, Chemosphere, 2012, 86, 853–859 CrossRef CAS PubMed.
  25. Y. Wang and P. Zhang, J. Hazard. Mater., 2011, 192, 1869–1875 CrossRef CAS PubMed.
  26. R. Dillert, D. Bahnemann and H. Hidaka, Chemosphere, 2007, 67, 785–792 CrossRef CAS PubMed.
  27. S. C. Panchangam, A. Y. C. Lin, J. H. Tsai and C. F. Lin, Chemosphere, 2009, 75, 654–660 CrossRef CAS PubMed.
  28. C. R. Estrellan, C. Salim and H. Hinode, J. Hazard. Mater., 2010, 179, 79–83 CrossRef CAS PubMed.
  29. B. Zhao and P. Zhang, Catal. Commun., 2009, 10, 1184–1187 CrossRef CAS.
  30. T. Shao, P. Zhang, L. Jin and Z. Li, Appl. Catal., B, 2013, 142, 654–661 CrossRef.
  31. B. Zhao, X. Li, L. Yang, F. Wang, J. Li, W. Xia, W. Li, L. Zhou and C. Zhao, J. Photochem. Photobiol., A, 2015, 91, 42–47 CrossRef CAS PubMed.
  32. X. Li, P. Zhang, L. Jin, T. Shao, Z. Li and J. Cao, Environ. Sci. Technol., 2012, 46, 5528–5534 CrossRef CAS PubMed.
  33. Z. Li, P. Zhang, T. Shao, J. Wang, L. Jin and X. Li, J. Hazard. Mater., 2013, 260, 40–46 CrossRef CAS PubMed.
  34. Z. Li, P. Zhang, J. Li, T. Shao and L. Jin, J. Photochem. Photobiol., A, 2013, 111–116 CrossRef.
  35. Z. Li, P. Zhang, J. Li, T. Shao, J. Wang and L. Jin, Catal. Commun., 2014, 43, 42–46 CrossRef CAS.
  36. F. Di Quarto, C. Sunseri, S. Piazza and M. C. Romano, J. Phys. Chem. B, 1997, 101, 2519–2525 CrossRef CAS.
  37. X. Yang, J. Xu, T. L. Wong, Q. D. Yang and C. S. Lee, Phys. Chem. Chem. Phys., 2013, 15, 12688–12693 RSC.
  38. J. Cui, A. Wang, N. L. Edleman, J. Ni, P. Lee, N. R. Armstrong and T. J. Marks, Adv. Mater., 2001, 13, 1476–1480 CrossRef CAS.
  39. Z. Li, P. Zhang, J. Li, T. Shao and L. Jin, J. Photochem. Photobiol., A, 2013, 271, 111–116 CrossRef CAS.
  40. X. Feng, D. C. Sayle, Z. L. Wang, M. S. Paras, B. Santora, A. C. Sutorik, T. X. T. Sayle, Y. Yang, Y. Ding, X. Wang and Y. S. Her, Science, 2006, 312, 1504–1508 CrossRef CAS PubMed.
  41. G. S. Li, D. Q. Zhang and J. C. Yu, Phys. Chem. Chem. Phys., 2009, 11, 3775–3782 RSC.
  42. X. Liu, K. Zhou, L. Wang, B. Wang and Y. Li, J. Am. Chem. Soc., 2009, 131, 3140–3141 CrossRef CAS PubMed.
  43. S. Hu, F. Zhou, L. Wang and J. Zhang, Catal. Commun., 2011, 12, 794–797 CrossRef CAS.
  44. B. Liu, X. Zhao, N. Zhang, Q. Zhao, X. He and J. Feng, Surf. Sci., 2005, 595, 203–211 CrossRef CAS.
  45. X. Wang, B. Zhai, M. Yang, W. Han and X. Shao, Mater. Lett., 2013, 112, 90–93 CrossRef CAS.
  46. Y. Song, H. Zhao, Z. Chen, W. Wang, L. Huang, H. Xu and H. Li, Phys. Status Solidi A, 2016, 1–8 Search PubMed.
  47. L. I. Trakhtenberg, G. N. Gerasimov, V. F. Gromov, T. V. Belysheva and O. J. IIgbusi, Sens. Actuators, B, 2015, 209, 562–569 CrossRef CAS.
  48. J. Lv, T. Kako, Z. Li, Z. Zou and J. Ye, J. Phys. Chem. C, 2010, 114, 6157–6162 CAS.
  49. L. Huang, Y. Li, H. Xu, Y. Xu, J. Xia, K. Wang, H. Li and X. Cheng, RSC Adv., 2013, 3, 22269–22279 RSC.
  50. S. C. Panchangam, A. Y. C. Lin, K. L. Shaik and C. F. Lin, Chemosphere, 2009, 7, 242–248 CrossRef PubMed.
  51. S. Deng, Y. Q. Zheng, F. J. Xu, B. Wang, J. Huang and G. Yu, Chem. Eng. J., 2012, 193–194, 154–160 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra09856h

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.