Ordered mesoporous hematite promoted by magnesium selective leaching as a highly efficient heterogeneous Fenton-like catalyst

Chunming Zheng*a, Xiangzhi Chenga, Peipei Chena, Chuanwu Yanga, Shoumin Baoa, Jun Xiaa, Minglin Guoa and Xiaohong Sun*b
aState Key Laboratory of Hollow-fiber Membrane Materials and Membrane Processes, School of Environmental and Chemical Engineering, Tianjin Polytechnic University, Tianjin 300387, P. R. China. E-mail: zhengchunming@tjpu.edu.cn; Fax: +86 022 83955140; Tel: +86 022 83955661
bKey Laboratory of Advanced Ceramics and Machining Technology, Ministry of Education, School of Materials Science and Engineering, Tianjin University, Tianjin 300072, P. R. China. E-mail: sunxh@tju.edu.cn; Fax: +86 022 27406114; Tel: +86 022 27406141

Received 17th February 2015 , Accepted 23rd April 2015

First published on 23rd April 2015


Abstract

Ordered mesoporous hematite with an ultrahigh surface area (up to 200 m2 g−1) was prepared through a hard templating method of Mg and Fe in mesoporous silica KIT-6 (meso-Mg/Fe2O3) and used for highly efficient wet peroxide oxidation of methylene blue. The obtained results showed that approximately two thirds of Mg cations were removed in the leaching process, resulting in a highly porous hematite with a significant amount of defects in the structure. The activated mesoporous iron oxide exhibited excellent catalytic activity for the degradation of methylene blue, achieving above 95% removal of 60 mg L−1 methylene blue after 3 h at reaction conditions of initial pH, 0.6 mg L−1 catalyst and 2600 mg L−1 H2O2 dosage. The apparent rate constant of used meso-Mg/Fe2O3 is 1.972 h−1, which is 1.22, 3.02 and 4.53 times those of meso-Fe2O3, con-Mg/Fe2O3 and α-Fe2O3, respectively. With the increase of reusability of the meso-Mg/Fe2O3 catalyst, both the leaching concentration of Mg and the catalytic activity of the catalyst increased, which is quite different with the catalytic mechanism of composite components in heterogeneous Fenton-like processes, such as Fe–Cu and Fe–Zn composites. The leaching concentrations of iron in the catalysts were found to be low (<5 mg L−1) in consecutive runs. Hence, the meso-Mg/Fe2O3 catalyst has proved to be an attractive alternative in the treatment of environmental refractory organic pollutants and has a unique and superb catalytic activity in the heterogeneous Fenton-like system.


Introduction

The conventional Fenton process requires strong acidic conditions (pH < 3) to prevent the hydrolysis of ferrous and ferric ions and to maintain acceptable conversion rates. In addition, non-recyclable soluble iron salts also yield large amounts of iron oxide sludge, which is regarded as a secondary pollutant and the loss of catalyst.1,2 In order to overcome the abovementioned disadvantages, heterogeneous Fenton-like catalysts have been investigated widely. In the past decade, several iron-based catalysts coupled with H2O2 for heterogeneous Fenton-like catalysis have been investigated,3 i.e. iron oxides and hydroxides4 iron composites mixed with suitable transition metal such as Co, Mn, Cu, Cr, and Zn,5 supported iron catalysts, including oxide supported iron catalysts,6 carbon material supported iron catalysts,7–10 Fe exchanged zeolites, iron exchanged resin, industrial solid waste supported Fe catalysts, and mesoporous material supported Fe catalysts.11 However, the heterogeneous Fenton-like activities of these catalysts were relatively slow if they were operated at high pH values or without external power supplies such as UV, ultrasound or microwave. Moreover, the leaching of active metals (such as Cu, Co and Cr) also leads to the depletion of their catalytic activity and environmental pollution problems. To overcome these drawbacks of heterogeneous Fenton-like systems, the newly developed heterogeneous catalysts should be recyclable in a broad pH range (pH = 3.0–9.0) and stable enough to avoid the serious leaching of active metals during operation.12 Therefore, the key problem is how to design and developed heterogeneous Fenton-like catalysts, which can effectively generate ˙OH by H2O2 decomposition with broad usability and high durability.

Recently, various strategies have been proposed to enhance the performance of heterogeneous Fenton-like catalysts,13 e.g. loading the catalysts on carriers with high surface area to improve their dispersion14 and reducing the size of the catalyst to nano-scales to increase surface energy.15 Among these methods, mesoporous iron oxides are of particular interest because they combine large internal surface area, nanosized walls, uniform pore size distributions, large pore volume, controllable compositions and surface functionalities.16 For example, Cornu C. et al. found that the loading of Fe to mesoporous SBA-15 could greatly enhance the degradation performance for methylene blue and studied the formation and location of iron species (including Fe2+ and Fe3+) in Fe/SBA-15 catalysts.17 Xia et al. introduced alumina into Fe–Cu bimetallic oxide supported MCM-41 catalysts to obtain high degradation activities for phenol.18 However, these methods still need to be improved to increase the dispersion of metals active sites on the support and overcome the vulnerable leakage problems of these metals.19,20 In addition, a wide range of iron-based metal composites with highly ordered mesostructures have been successfully synthesized with a hard templating (nanocasting) method. Some of them exhibited unique heterogeneous Fenton-like catalytic properties.21,22 Wang et al. prepared ordered mesoporous copper ferrite (meso-CuFe2O4) with a KIT-6 hard templating method, which can efficiently increase the degradation activity for imidacloprid.23 Su et al. reported that the substitution of Fe2+ by Zn2+ can remarkably increase the activity of mesoporous Fe3O4 nanoparticles.24 However, the leaching of active metals (Fe, Cu, Zn et al.) from these mesoporous catalysts was also inevitable, which undoubtedly leads to the reduction of their catalytic activity in the long term.

As we know, the leaching of Mg from mesoporous catalysts was unavoidable and even more serious than active metals such as Fe, Cu, Zn.25 Hu et al. reported that the leaching concentration of Mg from mesoporous CoMg/SBA-15 was much higher than that of Co. Interestingly, BET surface area and pore volume also increased after Mg selective leaching in this catalyst system.26 Jiao et al. reported the fabrication of mesoporous Co3O4 with Mg substitution, which exhibited high oxygen evolution activities in a visible-light-driven water oxidation system. Approximately one third of Mg cations were removed in the leaching process, resulting in a highly porous cobalt oxide with a significant amount of defects in the spinel structure.27 In this report, the Mg components in iron-based mesoporous structure are not used as the active metals for the heterogeneous Fenton-like catalysts, but implied to facilitate the leaching process of mesoporous α-Fe2O3 substituted by Mg ions. During the selective leaching of Mg ions, the porous structure of α-Fe2O3 still remained. Moreover, the surface area of the catalyst was also increased in favor of the adsorption of the macromolecules (scheme shown in Fig. 1).16,27


image file: c5ra03019f-f1.tif
Fig. 1 Schematic illustration of mesoporous Mg–Fe2O3 through Mg ions leaching. Arrows show the Mg leaching sites (i.e., defect sites).

In this study, highly ordered mesoporous magnesium-substituted α-Fe2O3 (meso-Mg/Fe2O3) was successfully synthesized through the nanocasting strategy and used as a heterogeneous Fenton catalyst with broad usability and enhanced durability. In addition, methylene blue (MB) is a typical synthetic cationic dye, which is non-biodegradable, extensively used in textile industry and usually chosen as a model contaminant in the relative studies, whose presence in wastewater may cause the risk of nausea, vomiting and burning of the eye.11 Therefore, MB was selected as a model compound to monitor the Fenton catalytic activity of meso-Mg/Fe2O3. The purpose of this study was to elucidate the effect of ordered mesoporous structure and the role of magnesium on the catalytic activity of meso-Mg/Fe2O3. For comparison, conventional α-Fe2O3, ordered α-Fe2O3 and conventional Mg–Fe mixed metal oxides were synthesized as reference catalysts. According to the detected ˙OH radical and surface reaction revealed by physico-chemical characterizations, the possible activation mechanism of meso-Mg/Fe2O3 was proposed. Finally, the stability and reusability of the catalyst was also investigated.

Experimental

1. Materials

All chemicals required to prepare the mesoporous KIT-6 silica and derived Mg substituted Fe2O3 catalysts were used as purchased. Tetraethylorthosilicate (TEOS) and triblock copolymer Pluronic P123 (Mw = 5800, EO20PO70EO20) was purchased from Sigma-Aldrich. Ferric nitrate (Fe(NO3)3·9H2O), magnesium nitrate (Mg(NO3)2·6H2O), butyl alcohol, hydrochloric acid (HCl), sodium hydroxide (NaOH), and n-hexane were obtained from Tianjin Guangfu Chemical Reagent Co., Ltd., China. H2O2 (30%, v/v). Ethanol, isopropanol and methylene blue (MB) were supplied by Shanghai Aladdin Chemical Reagent Co., Ltd, China. Deionized water was used throughout the experiments.

2. Synthesis of ordered mesoporous Mg-substituted α-Fe2O3

Ordered mesoporous Mg-substituted α-Fe2O3 was synthesized by a nanocasting strategy with mesoporous silica KIT-6.27 Preparation of mesoporous silica KIT-6 was synthesized according to the previous literature.28 In a typical synthesis, 0.6 g as-prepared mesoporous silica KIT-6 was dispersed in 5 mL toluene and stirred for 30 min, then 1.139 g Fe(NO3)3·9H2O and 0.362 g Mg(NO3)2·6H2O at an Fe/Mg molar ratio of 2 were added and refluxed for 6 h at 343 K. After refluxing, the resulting mixture was dried in an oven at 333 K overnight and calcined at 873 K for 3 h (heating rate 1 K min−1). Finally, the silica template KIT-6 was dissolved by 2 M NaOH in 333 K under continuously stirring for 24 h, and the red product was produced by washing with distilled water and ethanol and drying in a vacuum oven. Pure α-Fe2O3 mesoporous sample was also prepared by the same synthetic approach. The conventional Mg–Fe mixed metal oxides without mesoporous structure were prepared using the traditional solid-phase method, the specific steps are as follows: the corresponding mixture in stoichiometric proportions of Fe(NO3)3·9H2O and Mg(NO3)2·6H2O were calcined at 873 K for 3 h with a heating rate of 1 K min−1, and the product was denoted as con-Mg/Fe2O3.

3. Characterization

Powder X-ray diffraction (XRD) patterns were measured on a Bruker D8 Advance X-ray diffractometer using Cu Kα radiation (λ = 0.154 nm) as the X-ray source. High resolution transmission electron microscopy (HRTEM) images was collected on a Hitachi H-7650 transmission electron microscope with field emission gun at 100 kV. The surface morphology of the solid samples were investigated on a Hitachi S-4800 Scanning Electron Microscope (SEM). N2 adsorption–desorption isotherms were measured at 77 K using a Micromeritics Tristar 3000 sorptometer outgassed at 423 K for a minimum of 12 h. The specific surface area and the pore size distribution curves are obtained using the Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) methods from the adsorption data. Fourier transform infrared (FT-IR) spectroscopy were performed to assess different functional groups of the materials using KBr pressed disks with a Bruker TENSOR37 FT-IR spectrometer transmission analyzer. X-ray photoelectron spectroscopy (XPS) was recorded on a Kratos ASIS-HS X-ray photoelectronspectroscope with an Al Kα source at 150 W (15 kV, 10 mA).

4. Catalytic tests

Catalytic tests were carried out under mild conditions (atmospheric pressure and ∼298 K) in a thermostatic reactor of 250 mL with continuous stirring. The reaction was performed in the kinetic regime and no external and internal diffusional resistances occurred. The initial reaction pH was 7.2 and could be adjusted by adding 0.1 mol L−1 H2SO4 to the desired pH during the degradation processes. In a typical experiment, 50 mL of MB solution was prepared with an initial concentration of 60 mg L−1, and given amounts of hydrogen peroxide and catalyst were added at reaction time = 0 min. At regular intervals, 1.0 mL sample was collected in 1.5 mL centrifuge tubes and immediately mixed with 0.1 mL isopropanol to quench ˙OH. It was centrifuged three times (5 min each) at 13[thin space (1/6-em)]400 rpm to separate the catalysts, and the supernatant liquid was collected for analysis.

5. Sample analysis

The concentration of MB in the supernatant liquid was determined at an absorbance peak of 664 nm using a UV-Vis spectrophotometer (UV-752, Shanghai Rex Instrument Co., Ltd., China). ˙OH concentration was determined by the terephthalic acid (C8H6O4, TA) fluorescence method.29 The experimental procedure was similar to the measurement of Fenton catalytic activity except that the MB aqueous solution was replaced by 5 × 10−4 mol L−1 terephthalic acid aqueous solution with a concentration of 2 × 10−3 mol L−1 NaOH. The photoluminescence spectra (PL) of TA samples were obtained using a fluorescence spectrophotometer (F-380, Tianjin Gangdong Instrument Co., Ltd., China) at room temperature, and a wavelength excitation of 325 nm. The concentration of Fe(II) ions was determined by reacting the recovered solution after testing with 1,10-phenanthroline (maximum of absorption at 510 nm) by a Varian Cary 50 UV-Vis spectrophotometer. The Fe(III) concentration was determined by reaction with thiocyanate (maximum of absorption at 480 nm) under acidic conditions.30 The variation of H2O2 concentration during the reaction was analyzed colorimetrically by the same UV-Vis spectrophotometer after complexation with titanium salt.31

Result and discussion

1. Characterization of as-prepared samples

In order to investigate the structure–performance relationship of meso-Mg/Fe2O3, the morphology of as-prepared catalyst samples was characterized by SEM and the images are shown in Fig. 2. It can be seen that the size of these catalyst samples is homogeneous. By comparison between meso-Mg/Fe2O3 (Fig. 2C and E) and the KIT-6 templates (Fig. 2A), an increased roughness of catalyst surface in meso-Mg/Fe2O3 could be observed. The morphology of meso-Mg/Fe2O3 is also very different from meso-Fe2O3 and con-Mg/Fe2O3 (Fig. 2B–D). The former still maintained a similar surface morphology of KIT-6, and the latter was composed of particles with 200–1000 μm in diameter. After the recycling of meso-Mg/Fe2O3 and con-Mg/Fe2O3 (Fig. 2E and F), the catalyst particles aggregated and became larger. To further investigate the mesostructure of meso-Mg/Fe2O3, representative TEM images of as-synthesized catalyst samples are depicted in Fig. 3. From Fig. 3A–C, the mesoporous structure of meso-Mg/Fe2O3 is clearly visible even after the replication of KIT-6.32,33 Fig. 3D shows clear crystal lattice fringes, which indicate that meso-Mg/Fe2O3 was synthesized with a highly crystalline nature. After the recycling of the meso-Mg/Fe2O3 (Fig. 3E and F), the crystalline nature of the catalyst still remained.21
image file: c5ra03019f-f2.tif
Fig. 2 SEM images of KIT-6 (A), meso-Fe2O3 (B), meso-Mg/Fe2O3 (C), con-Mg/Fe2O3 (D), postreaction meso-Mg/Fe2O3 (E) and postreaction con-Mg/Fe2O3 (F).

image file: c5ra03019f-f3.tif
Fig. 3 TEM images of KIT-6 (A), meso-Fe2O3 (B), meso-Mg/Fe2O3 (C), high resolution meso-Mg/Fe2O3 (D), postreaction meso-Mg/Fe2O3 (E) and high resolution postreaction meso-Mg/Fe2O3 (F).

The successful preparation of meso-Mg/Fe2O3 is further confirmed by low-angle powder XRD analysis, as shown in Fig. 4A. The prepared meso-Mg/Fe2O3 catalyst indicated the excellent replication of KIT-6 templates. Mesoporous α-Fe2O3 was also in good agreement with the values reported in the literature.34 Compared with as-made meso-Mg/Fe2O3, there was a decrease in intensity of the diffraction peaks for the post-reaction meso-Mg/Fe2O3, suggesting the increased hematite particle dispersion, which may be attributed to the decrease in electron density contrast due to the increased Mg leaching during the reaction process.26 Wide-angle XRD results (including meso-Fe2O3, meso-Mg/Fe2O3, con-Mg/Fe2O3 and post-reaction meso-Mg/Fe2O3) are shown in Fig. 4B. The 2θ values of meso-Mg/Fe2O3 at 24.09°, 33.07°, 35.55°, 40.76°, 49.34°, 53.93°, 57.31°, 62.28° and 63.85° could be indexed to the (0 1 2), (1 0 4), (1 1 0), (1 1 3), (0 2 4), (1 1 6), (1 2 2), (2 1 4) and (3 0 0) planes of hexagonal hematite (α-Fe2O3, JCPDS no. 33-0664), respectively. The diffraction pattern for meso-Mg/Fe2O3 closely matches the standard α-Fe2O3, confirming the success of magnesium substitution to iron in the spinel. The peaks of meso-Mg/Fe2O3 are relatively broader than the peaks of con-Mg/Fe2O3, which is possibly due to the pore walls that were composed of small particles. The average crystallite size of meso-Mg/Fe2O3 and post-reaction meso-Mg/Fe2O3 was determined and calculated to be 7.3 and 5.1 nm, respectively, using the Scherrer equation:

image file: c5ra03019f-t1.tif
where Dhkl, λ, βhkl, and θhkl were the volume-averaged particle diameter, X-ray wavelength, full width at half maximum (FWHM) and diffraction angle, respectively.


image file: c5ra03019f-f4.tif
Fig. 4 (A) Low-angle XRD patterns of meso-Fe2O3, meso-Mg/Fe2O3 and postreaction meso-Mg/Fe2O3; (B) high-angle XRD patterns of α-Fe2O3, con-Mg/Fe2O3, meso-Fe2O3, meso-Mg/Fe2O3 and postreaction meso-Mg/Fe2O3.

Textural properties of mesoporous catalysts were investigated by N2 adsorption–desorption measurements, as shown in Fig. 5. The corresponding parameters, including BET surface area, pore volume and average pore size, are listed in Table 1. As shown in Fig. 5A, the isotherm of meso-Mg/Fe2O3 showed a type IV isotherm at P/P0 about 0.5, indicating a typical shape of mesoporous materials (Fig. 5A). The hysteresis shaping started from about 0.65 and extended almost to 0.95, indicating their large porosity.35 The specific surface area (90.7 m2 g−1) and the pore diameter (9.1 nm) are consistent with previous reports on others mesoporous metal oxides, which indicates the successful fabrication of mesoporous Mg/Fe2O3.


image file: c5ra03019f-f5.tif
Fig. 5 (A) N2 adsorption–desorption isotherms for meso-Mg/Fe2O3, acid washed meso-Mg/Fe2O3 and postreaction meso-Mg/Fe2O3; (B) pore size distributions from the desorption branches through the BJH method for meso-Mg/Fe2O3, acid washed meso-Mg/Fe2O3 and postreaction meso-Mg/Fe2O3.
Table 1 The structural and chemical properties of meso-Mg/Fe2O3 and selected iron-based catalystsa
Catalysts SBET (m2 g−1) Vpore (cm3 g−1) Vmicro (cm3 g−1) Dp (nm) Mg wt% Fe wt% Mg to Fe ratio
a Fe content in wt%: as determined by ICP-AES analysis; Vpore: mesopore volume measured at P/P0 = 0.97; Vmicro: micropore volume calculated from t-plot; Dp: BJH average pore size calculated on the adsorption branch.
KIT-6 575 1.08 0.23 6.8 N/A N/A N/A
Meso-Fe2O3 96.3 0.32 0.043 8.6 N/A 76.9 N/A
Meso-Mg/Fe2O3 90.7 0.37 0.056 9.1 11.0 43.3 0.59
Postreaction meso-Mg/Fe2O3 190.4 0.44 0.064 12.7 3.67 61.2 0.14
Acid washed meso-Fe2O3 103.7 0.35 0.049 9.4 N/A 72.3 N/A
Acid washed meso-Mg/Fe2O3 203.9 0.41 0.061 10.2 3.98 58.7 0.16
Con-Mg/Fe2O3 9.3 0.06 0.007 16.3 68.7 0.54


The surface physico-chemical characteristics of the as-prepared catalysts were further explored by FT-IR analysis. According to the infrared active modes polarized to the c-axis of the hexagonal crystal system of α-Fe2O3, the typical adsorption peaks of α-Fe2O3 had a broad absorption at around 560 cm−1 and a narrower one at around 480 cm−1.36,37 In Fig. 6A, the dominant bands of the meso-Mg/Fe2O3 (at about 537 cm−1 and 471 cm−1) are typical characteristic of crystalline α-Fe2O3. Interestingly, the dominant absorption peak at 568 cm−1 of α-Fe2O3 shifts to somewhat lower frequencies at 551 cm−1 of meso-Fe2O3 and 537 cm−1 of meso-Mg/Fe2O3. Moreover, the above absorption peak also obviously becomes broader than α-Fe2O3 from Fig. 6A. This might be due to the changes in the structure of the catalyst, since the catalyst consists of large numbers of small subparticles in the mesoporous structure.38 Thus, when the average crystallite diameter of subparticles is reduced, the absorption peaks become broader. These results are consistent with XRD and TEM. In Fig. 6B, the absorption peak at 540 cm−1 of post-reaction meso-Mg/Fe2O3 hardly changes compared with the one of as-made meso-Mg/Fe2O3, suggesting that the catalyst has not been affected by our experimental conditions.


image file: c5ra03019f-f6.tif
Fig. 6 FT-IR spectra of (A) bulk Fe2O3, meso-Fe2O3, meso-Mg/Fe2O3, con-Mg/Fe2O3, and (B) con-Mg/Fe2O3, meso-Mg/Fe2O3 and postreaction meso-Mg/Fe2O3 (for interpretation of the references in color for this figure legend, the reader is referred to the web version of this article).

2. Efficient MB degradation performance in the presence of meso-Mg/Fe2O3

In the previous studies, α-Fe2O3 has been proven to be easily recyclable and an efficient catalyst for the degradation of MB, therefore control experiments for the meso-Mg/Fe2O3–H2O2–MB system were evaluated.39,40 The external and internal diffusional resistances of the degradation processes were excluded, and the reaction was performed in the kinetic regime before the analysis. Prior to the addition of H2O2, the mixed solution of MB and the catalysts was stirred in the dark for 30 min to establish the equilibrium adsorption state.26

A series of experiments were performed by varying the catalyst loading of meso-Mg/Fe2O3 from 0.2 g L−1 to 0.6 g L−1 and the result is depicted in Fig. 7A. From Fig. 7A, the degradation efficiency increased with the meso-Mg/Fe2O3 catalyst loading. At 0.6 g L−1 of meso-Mg/Fe2O3, MB degradation reached almost 98% in 180 min.41 Moreover, the data for MB concentration decay were further analyzed by the first order kinetic equation, which could be expressed as

image file: c5ra03019f-t2.tif
where k′ is the apparent reaction constant and C0 and C are the initial concentration and the concentration at time t, of MB, respectively. The degradation of MB follows pseudo-first order reaction kinetics. The apparent rate constant of used meso-Mg/Fe2O3 is 1.972 h−1, which is 1.22, 3.02 and 4.53 times than those of meso-Fe2O3, con-Mg/Fe2O3 and α-Fe2O3. The rate constant k′ increased linearly with increasing catalyst dosage (see Fig. 7B and C).


image file: c5ra03019f-f7.tif
Fig. 7 (A) Removal efficiency of methylene blue in the presence of meso-Mg/Fe2O3 0.4 g L−1 (a), H2O2 (2600 mg L−1) (b), meso-Mg/Fe2O3 (0.2 g L−1) + H2O2 (c), meso-Mg/Fe2O3 (0.4 g L−1) + H2O2 (d) and meso-Mg/Fe2O3 (0.6 g L−1) + H2O2 (e); (B) kinetic curves of methylene blue degradation on the dependence of ln(C/C0) versus time; (C) kinetic analysis of methylene blue degradation on the dependence of meso-Mg/Fe2O3 catalyst loading (initial pH, H2O2 dosage = 2600 mg L−1 and initial MB concentration = 60 mg L−1).

During the heterogeneous Fenton degradation of MB, it should be noted that the adsorption of MB by meso-Mg/Fe2O3 could achieve about 9% (Fig. 7A, curve (a)), which is due to the high specific surface area of meso-Mg/Fe2O3 (90.7 m2 g−1). The experiment in homogeneous systems was also performed. After vigorous agitation for 4 h, the meso-Mg/Fe2O3 catalyst was removed and H2O2 was added into the clear filtrate (380 μL, 30 wt%). As shown in Fig. 7A and 8A, the removal of MB by adsorption was 9%, the removal of MB by pure H2O2 at 6 h was 12%, while the removal of MB catalyzed by leaching Fe ions (concentration ∼5 mg L−1) was 25%. Therefore, the contribution of homogeneous Fenton reaction was 3%. The results of this experiment clearly indicate that the production of hydroxyl radicals observed primarily occurs at the nanoparticle surface rather than being catalyzed by the Fe ions leached into the filtrate. To accurately estimate the catalytic activity of meso-Mg/Fe2O3, comparable amounts of FeSO4 (Fe ions concentration 60 mg L−1) were used to catalyze the degradation of MB. The results show that 60 mg L−1 MB could be completely degraded in 30 min (Fig. 8A). These results indicated that the contribution of the Fenton-like reaction was mainly dominated by the meso-Mg/Fe2O3 catalysts. From Fig. 8B, four types of catalysts, with the same Fe loading, display distinctly different catalytic activities (meso-Mg/Fe2O3 > meso-Fe2O3 > con-Mg/Fe2O3 > α-Fe2O3). The excellent catalytic activity of mesoporous catalysts strongly originates from the ordered mesoporous network structure and large pore size, which contributes to the raised numbers of accessible active sites of the catalysts and increased mass transfer of MB in the catalysts porosity.42


image file: c5ra03019f-f8.tif
Fig. 8 (A) The degradation efficiency of methylene blue with different catalyst during the reaction. Homogeneous Mg2+ (5 mg L−1) (a), homogeneous Fe2+ (5 mg L−1) (b), α-Fe2O3 (c), con-Mg/Fe2O3 (d), meso-Fe2O3 (e), meso-Mg/Fe2O3 (0.6 g L−1) (f) and homogeneous Fe2+ (60 mg L−1) (g); (B) the kinetic curves of methylene blue degradation on the dependence of In(C/C0) versus time (initial pH, H2O2 dosage = 2600 mg L−1 and initial MB concentration = 60 mg L−1).

It is well known that pH value is one of most significant factors in the Fenton reaction system.43,44 Investigation and extension of pH range on the performance of the heterogeneous Fenton-like catalyst is of crucial importance for the development of new heterogeneous Fenton-like processes.16 As shown in Fig. 9, the initial rate was relatively slow at pH 3.5, and then it followed a fast degradation rate. In case of pH 7.5, the rate of reaction was very fast and the conversion of MB was 98% over a period of 180 min. The results indicated that the reaction performance is highest at around pH 7.5 and decreased with increasing pH. In general, the homogeneous Fenton system shows best performance at pH ≈ 3.45 At neutral and even higher pH, formation of Fe(OH)3 precipitate (pKsp = 37.4) prevents the catalysis reaction proceeding. However, the addition of Mg into heterogeneous Fenton catalysts could expand its degradation pH range during the reaction processes. The substitution of Mg on heterogeneous Fenton catalysts makes the slightly acidic surface of catalysts become slightly basic.26 These basic sites on the catalyst surface can promote the formation of the surface Fe–OH complex, which is crucial in activating H2O2.46,47 Therefore, compared with the conventional Fenton reaction, meso-Mg/Fe2O3 could be used over a wide pH range.


image file: c5ra03019f-f9.tif
Fig. 9 Effect of initial pH on degradation of methylene blue with Mg/Fe2O3 suspensions: pH 3.5 (a), pH 5.5 (b), pH 7.5 (c), pH 8.5 (d) and pH 9.5 (e) (H2O2 dosage = 2600 mg L−1, initial MB concentration = 60 mg L−1 and catalyst dosage = 0.6 g L−1).

The remained TOC after total MB degradation with different catalyst could be attributed to organic acids accumulated in the solution (Fig. 10). It can be observed that pure H2O2 and α-Fe2O3 did not produce a significant TOC reduction. Con-Mg/Fe2O3 and meso-Fe2O3 showed 22% and 31% TOC reduction, respectively. In addition, meso-Mg/Fe2O3 produced the highest TOC reduction of about 54% after 280 min reaction. In order to identify the end products, degraded samples of MB were analyzed by ion exclusion chromatography (results not reported). Oxalic acid and formic acid were identified as byproducts of the aromatic ring cleavage and are often detected after degradation of MB.43


image file: c5ra03019f-f10.tif
Fig. 10 Temporal changes of TOC during the degradation of methylene blue with different catalyst. H2O2 (2600 mg L−1) (a), α-Fe2O3 (b), con-Mg/Fe2O3 (c), meso-Fe2O3 (d) and meso-Mg/Fe2O3 (0.6 g L−1) (e) (initial pH, H2O2 dosage = 2600 mg L−1 and initial MB concentration = 60 mg L−1).

In order to compare the effect of the meso-Mg/Fe2O3–H2O2 Fenton system and H2O2 on the ˙OH radicals concentration, experiments using different heterogeneous Fenton catalysts are researched under the same conditions. The generated ˙OH radicals in the catalytic process were investigated according to the reaction ˙OH and terephthalic acid producing 2-hydroxyterephthalic acid. The PL intensity at 425 nm gradually rises with increasing irradiation time in the presence of different catalysts (Fig. 11). From Fig. 11D, the ˙OH intensity of meso-Mg/Fe2O3 is higher than that of the individual meso-Fe2O3 and con-Mg/Fe2O3. Therefore, the degradation activity of the meso-Mg/Fe2O3 could also be expected to be higher than that of meso-Fe2O3 and con-Mg/Fe2O3. The abovementioned results also account for the higher catalytic activity of meso-Mg/Fe2O3 in comparison with that of α-Fe2O3. From Fig. 11C, the generation rates of ˙OH radicals in con-Mg/Fe2O3 and meso-Mg/Fe2O3 were slower than that of α-Fe2O3 and meso-Fe2O3 at the beginning of the reaction. After 60 min, in sharp contrast, the abovementioned Mg-substituted samples (especially meso-Mg/Fe2O3) showed much higher generation rates compared to the pure α-Fe2O3 and meso-Fe2O3. This indicates that the initial surface structure of the as-made catalyst plays a critical role at the beginning of the reaction. However, if the reaction time is long enough, the meso-Mg/Fe2O3 catalyst experienced the Mg leaching under catalytic environment. The departure of Mg cations from the catalysts could create defects or vacancies in the meso-Mg/Fe2O3 structure, and therefore activate the catalyst.27 A more stable surface of the catalyst is formed for prolonged catalytic oxidation, leading to the increased catalyst activity (shown in Table 1).21 The leaching process is slow and could not be observed in a short reaction time (30 min). The leaching/activation mechanism could be summarized in Fig. 1.


image file: c5ra03019f-f11.tif
Fig. 11 Total concentrations of ˙OH formed as a function of time. α-Fe2O3 (A), meso-Fe2O3 (B), con-Mg/Fe2O3 (C) and meso-Mg/Fe2O3 (D) (pH = 3.00, H2O2 dosage = 1300 mg L−1, catalyst dosage = 0.3 g L−1 and benzoic acid used as a probe at 1200 mg L−1).

In order to clarify the origin of the difference in activities, detailed structural characterizations of meso-Mg/Fe2O3 were studied before and after the degradation of MB. The chemical composition of meso-Mg/Fe2O3 and related catalysts was confirmed by ICP-AES, and the Mg/Fe molar ratio is shown in Table 1. From Table 1, the Mg/Fe molar ratio was reduced to 0.14 compared to the initial Mg/Fe ratio of 0.59. The post-reaction meso-Mg/Fe2O3 exhibited typical isotherms for traditional ordered mesoporous materials and a much higher BET surface area (190.4 m2 g−1) than the as-made catalyst (90.7 m2 g−1). Thus, Mg leaching during the reaction processes increases the surface area of meso-Mg/Fe2O3 by approximately 100 m2 g−1. To verify the contributions of Mg cations to the increased surface area, the as-made meso-Mg/Fe2O3 and meso-Fe2O3 were treated in a diluted aqueous HNO3 solution (pH = 3) for 30 min and investigated by N2 adsorption measurements. The BET surface areas for acid-treated meso-Mg/Fe2O3 and meso-Fe2O3 are 203.9 m2 g−1 and 103.7 m2 g−1, respectively. These results are consistent with our conclusion that the Mg leaching contributes ∼100 m2 g−1 surface area. The departure of Mg cations from mesoporous Fe2O3 surface may create defects or vacancies in the iron oxide surface and leads to a highly porous structure.27

XPS analysis of meso-Mg/Fe2O3 was carried out to better understand the roles of iron and magnesium in Fenton-like catalytic oxidation (as shown in Fig. 12). The O1s core level spectrum shows the dominant oxide peaks at around 530.1 eV, which are in good agreement with the literature values of α-Fe2O3.48 In the meso-Mg/Fe2O3 and post-reaction meso-Mg/Fe2O3 catalysts, this peak practically disappears and shift to 531.7 eV and 530.8 eV. During the heterogeneous Fenton oxidation processes, the Mg ions were removed from the meso-Mg/Fe2O3 structure. The departure of Mg cations from catalyst structure may prompt the O1s peak of meso-Mg/Fe2O3 gradually close to the peak of pure α-Fe2O3 due to strong interaction between Fe3+ and related metal oxides. A very complex line-shape is observed in the Fe2p spectra (Fig. 12B) of the same catalysts. As expected, the typical photoelectron peaks at 711 and 725 eV are the characteristic doublets of Fe3+2p3/2 and Fe3+2p1/2 core-level spectra of α-Fe2O3. In addition, the spectra of meso-Mg/Fe2O3 and post-reaction meso-Mg/Fe2O3 were fitted with six contributions.4 The peaks at 724.5 eV and 711.1 eV in the meso-Mg/Fe2O3 are assigned to Fe3+2p3/2 and Fe3+2p1/2, suggesting the presence of Fe3+ cations. The other peaks are due to shake-up satellites of the 2p3/2 and 2p1/2 peaks. In post-reaction meso-Mg/Fe2O3, the Fe3+2p3/2 and Fe3+2p1/2 peaks are shifted to 724.1 eV and 710.7 eV, respectively, due to the interaction between the clusters of iron oxides and related oxides. Furthermore, the Fe2p peaks were gradually close to the peaks of pure α-Fe2O3, indicating the valence of Fe was not changed and its impact on Fe was gradually reduced with the leaching of Mg cations.


image file: c5ra03019f-f12.tif
Fig. 12 XPS spectra of the α-Fe2O3, meso-Mg/Fe2O3 and postreaction meso-Mg/Fe2O3 samples: O1s (A) and Fe2p (B).

Once mesoporous α-Fe2O3 is fabricated through substitution and selective leaching of magnesium, a highly porous mesoporous Mg modified Fe2O3 with a significant amount of defects was made. The catalyst showed an enhanced activity in the MB degradation under mild conditions. Based on the experimental results of the XPS and the measured amount of ˙OH radicals, a plausible mechanism related to the active ˙OH radicals in the meso-Mg/Fe2O3 system is proposed. Firstly, MB was adsorbed on the surface of meso-Mg/Fe2O3. Once H2O2 was added, Fe(III) initiated H2O2 decomposition started from the reduction of Fe(III) and the formation of HO2˙ (O2˙). Then, OH radicals and reduced Fe(II) species were generated according to the Haber–Weiss mechanism.49,50 Next, the ˙OH radicals reacted with MB both adsorbed on the catalyst surface and dissolved in the bulk solution, leading to the further degradation and mineralization of pollutants. Nevertheless, the decomposition of H2O2 to generate ˙OH radicals can be described in the following reactions:18

 
≡Fe(III) + H2O2 → ≡Fe(III)(H2O2)ads (1)
 
≡Fe(III)(H2O2)ads → ≡Fe(II) + HO2˙ +H+ (2)
 
≡Fe(II) + H2O2 → ≡Fe(III) + ˙OH + OH (3)

It can be noted that at pH ≥ 4 in a heterogeneous system, the catalytic reaction mainly occurs on the catalyst surface. Therefore, the adsorption of H2O2 onto the metal centers plays an important role in this heterogeneous Fenton-like reaction. In the case of meso-Mg/Fe2O3, the surface nature of Mg is also highly hydroxylated, which can facilitate the formation of the surface ˙OH complexes.46,51 This is crucial for the adsorption of H2O2 and consequently contributes to a more rapid performance of the oxidation processes.47 Moreover, MB hardly degraded in the presence of dissolved Mg2+ ions and H2O2, indicating that Mg2+ ions is not responsible for the MB degradation. Hence, a much better catalytic activity of meso-Mg/Fe2O3 compared with meso-Fe2O3 was ascribed to the synergetic effects of the desirable basic sites on the surface of the Mg and increased BET surface area and broader pore size distributions.

3. Excellent chemical reusability and stability of meso-Mg/Fe2O3 catalyst

Fig. 13 shows the MB degradation on meso-Mg/Fe2O3 in three different batch runs. After each run, the catalyst was recovered by filtration and washed with deionized water and dried at 80 °C overnight. It can be noted that the activity of meso-Mg/Fe2O3 increased after regeneration. As mesoporous α-Fe2O3 is the active component for heterogeneous Fenton-like reactions, substitution of Fe with alkaline metals will significantly decrease the activity at first batch run. Then, with the leaching of Mg and surface restructuring under a catalytic environment, the Mg-substituted catalysts (especially meso-Mg/Fe2O3) were activated (the enlarged part of Fig. 13).27 This indicates that meso-Mg/Fe2O3 possesses enhanced chemical and catalytic stability. Notably, with the increase of reusability of the meso-Mg/Fe2O3 catalyst, both the leaching concentration of Mg and the catalytic activity of the catalyst increased, which is quite different with the catalytic mechanism of composite components, such as Fe–Cu and Fe–Zn composites, in heterogeneous Fenton processes.23,24 The addition of Cu2+ is thermodynamically favorable for the reduction of Fe3+, which is beneficial for the redox cycles of Fe2+/Fe3+ and even for the regeneration of the active species Fe2+. In the meso-Mg/Fe2O3–H2O2–MB system, approximately two thirds of Mg cations were removed in the leaching/activating process. Especially, the meso-Mg/Fe2O3 with unique mesoporous structure possesses ultrahigh surface area and enlarged pore size, which effectively reduced the mass transfer resistance and enhance the number of active sites of the catalyst.
image file: c5ra03019f-f13.tif
Fig. 13 Degradation of methylene blue in different batch runs in the H2O2–meso-Mg/Fe2O3 system. First run (a), second run (b) and third run (c) (initial pH, H2O2 dosage = 2600 mg L−1, initial MB concentration = 60 mg L−1 and catalyst dosage = 0.6 g L−1).

The stability of the catalyst was also evaluated under the reaction conditions of the meso-Mg/Fe2O3–H2O2–MB system. Table 2 shows the influence of pH, H2O2 concentration, catalyst dosage and reusage on dissolved iron concentrations during the reactions. It can be seen that the dissolved iron species increased from 3.45 mg L−1 to 6.72 mg L−1 after 3 h when the pH decreased from 7.5 to 2.0. Table 2 also shows that dissolved iron species increased from 3.17 mg L−1 to 4.53 mg L−1 after 3 h reaction when the H2O2 concentration increased from 1600 mg L−1 to 4600 mg L−1 at an initial pH of 7.5.52,53 The leaching concentration of Mg decreased after the second and third cycles, which indicates that the Mg cations in the meso-Mg/Fe2O3 also reduced. Magnesium is an essential element required by the human body. It is not regulated in aqueous environments, which implies that the Mg ions leached from the meso-Mg/Fe2O3 catalyst will not result in any extra environmental and health problems. The treated waste water could be directly discharged into the sewage system.

Table 2 Effect of pH values on the methylene blue degradation efficiency (%) and iron leaching (mg L−1)a
Run no. pH0 [H2O2] (mg L−1) Catalyst dosage (g L−1) Degradation efficiency (%) Fe leaching (mg L−1) Mg leaching (mg L−1)
a pH0: initial pH adjusted at the beginning of the degradation of methylene blue, [H2O2]: initial H2O2 dosage, catalyst dosage: meso-Mg/Fe2O3 used in the reaction, Fe leaching and Mg leaching in mg L−1: as determined by ICP-AES analysis.
1 2 2600 0.6 2.3 6.72
2 4 2600 0.6 9.5 5.36
3 6 2600 0.6 28.2 4.31
4 7.5 1600 0.6 91.7 3.17
5 7.5 4600 0.6 96.1 4.53
6 7.5 2600 0.9 98.7 5.41
7 7.5 2600 0.6 93.2 3.45 217.6
First reused meso-Mg/Fe2O3 7.5 2600 0.6 96.5 5.13 47.3
Second reused meso-Mg/Fe2O3 7.5 2600 0.6 97.8 6.84 36.4


Conclusions

The ordered mesoporous magnesium-substituted α-Fe2O3 (meso-Mg/Fe2O3) was successfully synthesized using KIT-6 as a template through a hard templating strategy. The texture, morphology and structure of the meso-Mg/Fe2O3 catalyst before and after degradation of MB were characterized by SEM, HRTEM, XRD, N2 adsorption–desorption, FT-IR spectroscopy and XPS analysis. The pseudo-first order reaction kinetics fitted well to all experiments. With the increase of reusability of the meso-Mg/Fe2O3 catalyst, both the leaching concentration of Mg and the catalytic activity of the catalyst increased, which is quite different with the catalytic mechanism of composite components, such as Fe–Cu and Fe–Zn composites, in heterogeneous Fenton-like processes. Post-reaction meso-Mg/Fe2O3 showed a much higher BET surface area and broader pore size distribution than the as-made sample, which is likely due to Mg leaching during the Fenton-like catalysis process. The structural framework of as-prepared catalyst was not changed during the reaction. In addition, the released amounts of Fe were found to be low. Hence, the meso-Mg/Fe2O3 catalyst has proven to have a unique and superb catalytic activity in the heterogeneous Fenton-like system and it will be an attractive alternative in the treatment of environmental problems.

Acknowledgements

This study was supported by the National Natural Science Foundation of China, NSFC (21101113, 51172157, 51202159, 51208357, 51472179), the Doctoral Program of Higher Education, the Ministry of Education (20120032120017), and the Municipal Natural Science Foundation of Tianjin (13JCYBJC16900, 13JCQNJC08200).

References

  1. C. L. Yap, S. Gan and H. K. Ng, Chemosphere, 2011, 83, 1414–1430 CrossRef CAS PubMed.
  2. X. Zhang, M. Lin, X. Lin, C. Zhang, H. Wei, H. Zhang and B. Yang, ACS Appl. Mater. Interfaces, 2014, 6, 450–458 CAS.
  3. M. C. Pereira, L. C. A. Oliveira and E. Murad, Clay Miner., 2012, 47, 285–302 CrossRef CAS PubMed.
  4. J. V. Coelho, M. S. Guedes, R. G. Prado, J. Tronto, J. D. Ardisson, M. C. Pereira and L. C. A. Oliyeira, Appl. Catal., B, 2014, 144, 792–799 CrossRef CAS PubMed.
  5. A. Dhakshinamoorthy, S. Navalon, M. Alvaro and H. Garcia, ChemSusChem, 2012, 5, 46–64 CrossRef CAS PubMed.
  6. A. D. Bokare and W. Choi, J. Hazard. Mater., 2014, 275, 121–135 CrossRef CAS PubMed.
  7. S. Navalon, A. Dhakshinamoorthy, M. Alvaro and H. Garcia, ChemSusChem, 2011, 4, 1712–1730 CrossRef CAS PubMed.
  8. L. Zhou, Y. Shao, J. Liu, Z. Ye, H. Zhang, J. Ma, Y. Jia, W. Gao and Y. Li, ACS Appl. Mater. Interfaces, 2014, 6, 7275–7285 CAS.
  9. C. Liu, J. Li, J. Qi, J. Wang, R. Luo, J. Shen, X. Sun, W. Han and L. Wang, ACS Appl. Mater. Interfaces, 2014, 6, 13167–13173 CAS.
  10. Y. Wang, G. Zhao, S. Chai, H. Zhao and Y. Wang, ACS Appl. Mater. Interfaces, 2013, 5, 842–852 CAS.
  11. M. Muruganandham, R. P. S. Suri, M. Sillanpaa, J. J. Wu, B. Ahmmad, S. Balachandran and M. Swaminathan, J. Nanosci. Nanotechnol., 2014, 14, 1898–1910 CrossRef CAS PubMed.
  12. X. J. Yang, X. M. Xu, J. Xu and Y. F. Han, J. Am. Chem. Soc., 2013, 135, 16058–16061 CrossRef CAS PubMed.
  13. S. R. Pouran, A. A. A. Raman and W. M. A. W. Daud, J. Cleaner Prod., 2014, 64, 24–35 CrossRef PubMed.
  14. F. X. Bu, M. Hu, L. Xu, Q. Meng, G. Y. Mao, D. M. Jiang and J. S. Jiang, Chem. Commun., 2014, 50, 8543–8546 RSC.
  15. T. Zeng, X. Zhang, S. Wang, Y. Ma, H. Niu and Y. Cai, Chem.–Eur. J., 2014, 20, 6474–6481 CrossRef CAS PubMed.
  16. S. Navalon, M. Alvaro and H. Garcia, Appl. Catal., B, 2010, 99, 1–26 CrossRef CAS PubMed.
  17. C. Cornu, J. L. Bonardet, S. Casale, A. Davidson, S. Abramson, G. Andre, F. Porcher, I. Grcic, V. Tomasic, D. Vujevic and N. Koprivanac, J. Phys. Chem. C, 2012, 116, 3437–3448 CAS.
  18. M. Xia, M. Long, Y. Yang, C. Chen, W. Cai and B. Zhou, Appl. Catal., B, 2011, 110, 118–125 CrossRef CAS PubMed.
  19. J. Shi, Z. Ai and L. Zhang, Water Res., 2014, 59, 145–153 CrossRef CAS PubMed.
  20. M. Zhang, Q. Yao, W. Guan, C. Lu and J. M. Lin, J. Phys. Chem. C, 2014, 118, 10441–10447 CAS.
  21. A. H. Lu and F. Schueth, Adv. Mater., 2006, 18, 1793–1805 CrossRef CAS PubMed.
  22. P. D. Yang, D. Y. Zhao, D. I. Margolese, B. F. Chmelka and G. D. Stucky, Nature, 1998, 396, 152–155 CrossRef CAS PubMed.
  23. Y. Wang, H. Zhao, M. Li, J. Fan and G. Zhao, Appl. Catal., B, 2014, 147, 534–545 CrossRef CAS PubMed.
  24. M. Su, C. He, V. K. Sharma, M. A. Asi, D. Xia, X. Z. Li, H. Deng and Y. Xiong, J. Hazard. Mater., 2012, 211, 95–103 CrossRef PubMed.
  25. M. Hartmann, S. Kullmann and H. Keller, J. Mater. Chem., 2010, 20, 9002–9017 RSC.
  26. L. Hu, F. Yang, W. Lu, Y. Hao and H. Yuan, Appl. Catal., B, 2013, 134, 7–18 CrossRef PubMed.
  27. J. Rosen, G. S. Hutchings and F. Jiao, J. Am. Chem. Soc., 2013, 135, 4516–4521 CrossRef CAS PubMed.
  28. M. Choi, K. Na, J. Kim, Y. Sakamoto, O. Terasaki and R. Ryoo, Nature, 2009, 461, 246–249 CrossRef CAS PubMed.
  29. K. Ishibashi, A. Fujishima, T. Watanabe and K. Hashimoto, J. Photochem. Photobiol., A, 2000, 134, 139–142 CrossRef CAS.
  30. H. Tamura, K. Goto, T. Yotsuyanagi and M. Nagayama, Talanta, 1974, 21, 314–318 CrossRef CAS.
  31. J. De Laat and H. Gallard, Environ. Sci. Technol., 1999, 33, 2726–2732 CrossRef CAS.
  32. X. Sun, Y. Shi, P. Zhang, C. Zheng, X. Zheng, F. Zhang, Y. Zhang, N. Guan, D. Zhao and G. D. Stucky, J. Am. Chem. Soc., 2011, 133, 14542–14545 CrossRef CAS PubMed.
  33. H. F. Yang and D. Y. Zhao, J. Mater. Chem., 2005, 15, 1217–1231 CAS.
  34. W. Yue and W. Zhou, Chem. Mater., 2007, 19, 2359–2363 CrossRef CAS.
  35. N. Najmoddin, A. Beitollahi, M. Muhammed, N. Ansari, E. Devlin, S. M. Mohseni, H. Rezaie, D. Niarchos, J. Akerman and M. S. Toprak, J. Alloys Compd., 2014, 598, 191–197 CrossRef CAS PubMed.
  36. A. L. Andrade, D. M. Souza, M. C. Pereira, J. D. Fabris and R. Z. Domingues, J. Nanosci. Nanotechnol., 2009, 9, 3695–3699 CrossRef CAS PubMed.
  37. Y. S. Wang, A. Muramatsu and T. Sugimoto, Colloids Surf., A, 1998, 134, 281–297 CrossRef CAS.
  38. Z. Jing, S. H. Wu, S. M. Zhang and W. P. Huang, Mater. Res. Bull., 2004, 39, 2057–2064 CrossRef CAS PubMed.
  39. I. Ursachi, A. Stancu and A. Vasile, J. Colloid Interface Sci., 2012, 377, 184–190 CrossRef CAS PubMed.
  40. Y. Liu, H. Yu, S. Zhan, Y. Li, Z. Lv, X. Yang and Y. Yu, J. Sol-Gel Sci. Technol., 2011, 58, 716–723 CrossRef CAS.
  41. C. Bai, W. Gong, D. Feng, M. Xian, Q. Zhou, S. Chen, Z. Ge and Y. Zhou, Chem. Eng. J., 2012, 197, 306–313 CrossRef CAS PubMed.
  42. X. Zhong, J. Barbier Jr, D. Duprez, H. Zhang and S. Royer, Appl. Catal., B, 2012, 121, 123–134 CrossRef PubMed.
  43. J. L. Wang and L. J. Xu, Crit. Rev. Environ. Sci. Technol., 2012, 42, 251–325 CrossRef CAS PubMed.
  44. V. K. Gupta, I. Ali, T. A. Saleh, A. Nayak and S. Agarwal, RSC Adv., 2012, 2, 6380–6388 RSC.
  45. J. Herney-Ramirez, M. A. Vicente and L. M. Madeira, Appl. Catal., B, 2010, 98, 10–26 CrossRef CAS PubMed.
  46. Q. Yang, H. Choi, S. R. Al-Abed and D. D. Dionysiou, Appl. Catal., B, 2009, 88, 462–469 CrossRef CAS PubMed.
  47. M. Stoyanova, I. Slavova, S. Christoskova and V. Ivanova, Appl. Catal., A, 2014, 476, 121–132 CrossRef CAS PubMed.
  48. L. Li, Y. Chu, Y. Liu and L. Dong, J. Phys. Chem. C, 2007, 111, 2123–2127 CAS.
  49. W. P. Kwan and B. M. Voelker, Environ. Sci. Technol., 2003, 37, 1150–1158 CrossRef CAS.
  50. S. S. Lin and M. D. Gurol, Environ. Sci. Technol., 1998, 32, 1417–1423 CrossRef CAS.
  51. W. Zhang, H. L. Tay, S. S. Lim, Y. Wang, Z. Zhong and R. Xu, Appl. Catal., B, 2010, 95, 93–99 CrossRef CAS PubMed.
  52. J. H. Ramirez, F. J. Maldonado-Hodar, A. F. Perez-Cadenas, C. Moreno-Castilla, C. A. Costa and L. M. Madeira, Appl. Catal., B, 2007, 75, 312–323 CrossRef CAS PubMed.
  53. X. Hu, B. Liu, Y. Deng, H. Chen, S. Luo, C. Sun, P. Yang and S. Yang, Appl. Catal., B, 2011, 107, 274–283 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.