Facile hydrothermal synthesis and improved photocatalytic activities of Zn2+ doped Bi2MoO6 nanosheets

Xiang-Biao Zhang a, Lei Zhang *bc, Jin-Song Hu a and Xin-Hua Huang *b
aSchool of Chemical Engineering, Anhui University of Science and Technology, Huainan 232001, P. R. China
bLaboratory of Multiscale Materials and Molecular Catalysis, School of Materials Science and Engineering, Anhui University of Science and Technology, Huainan, Anhui 232001, P. R. China. E-mail: hxh0317@hotmail.com
cState Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210093, P. R. China. E-mail: yutian1224@126.com

Received 16th March 2016 , Accepted 22nd March 2016

First published on 24th March 2016


Abstract

A novel sheet-like Zn2+ doped Bi2MoO6 photocatalyst was fabricated through a facile hydrothermal strategy. The morphologies and chemical compositions of the samples were investigated with the help of powder X-ray diffraction (XRD), X-ray photoelectron spectrum (XPS), energy dispersive X-ray analysis (EDS), high resolution transmission electron microscopy (HRTEM), scanning electron microscopy (SEM), Raman and FT-IR spectra. These experimental results indicated that Zn2+ ions could be incorporated into the crystal lattices of Bi2MoO6 and substituted the Bi3+ ions. Photocatalytic experiments revealed that the as-obtained Zn2+ doped Bi2MoO6 nanosheets possessed excellent UV-visible light induced photocatalytic ability for the decomposition of Rhodamine B (RhB) molecules by comparison with undoped Bi2MoO6. The effect of some scavengers on the photocatalytic efficiency was assessed. Moreover, a possible enhancement mechanism of photocatalytic activity over Zn2+ doped Bi2MoO6 nanosheets was presented based on the systematical characterization results.


1. Introduction

Since the discovery of the Honda–Fujishima effect, semiconductor-based photocatalysis technology has afforded a potentially promising route to deal with serious energy and environmental problems.1–5 Without doubt, in order to achieve efficient photocatalytic technology, it is necessary to first consider how to obtain efficient photocatalysts.6–9 In the past several years, various photocatalysts have been developed and the most famous one is titanium dioxide.10–12 However, an inherent shortcoming of titanium dioxide, namely the relatively wide band-gap, may seriously restrict its practical application.13 Therefore, considering the full use of solar energy, it is imperative to design and develop new photocatalyst with efficient visible-light activity.14–16

Among many bismuth-based photocatalysts, Bi2MoO6 attracts keen research interest because of its fascinating physical and chemical properties, including suitable band gap, excellent chemical inertness and cheapness.17–19 As a typical Aurivillius oxide, Bi2MoO6 owes interesting layered structures in which [Bi2O2]2+ layers are sandwiched between MoO42− slabs, and can be regarded as a new kind of visible-light photocatalyst with two-dimensional layered structure. Latest researches have demonstrated that Bi2MoO6 can display photocatalytic ability for environmental treatment and water splitting under the irradiation of visible light irradiation.20–22 However, it remains a challenge to realize the practical applications of Bi2MoO6 photocatalyst owing to some immanent defects, such as relatively low quantum yield and poor charge separation efficiency.23–25

Doping impurities into a semiconductor can be regarded as one of the most commonly used methods to prolong the lifetime of photogenerated charges and improve the photocatalytic efficiency.26–28 For example, Ding et al. found that the reactive species regulation over Bi2MoO6 under visible light could be realized through a Bi self-doping technology via a simple soft-chemical method, leading to the significant enhancement of photocatalytic property.29 Alemi and coworkers reported the successful preparation of novel Ln(Gd3+, Ho3+ and Yb3+)-doped Bi2MoO6 photocatalysts and presented a possible enhancement mechanism of photocatalytic ability.30 Recently, Ren et al. demonstrated that the photocatalytic efficiency of Bi2WO6 matrix might be significantly improved by introducing Zn2+ to replace Bi3+ lattice sites based on a first-principles calculation method.31 Due to the structure similarity between Bi2WO6 and Bi2MoO6, it is reasonable to believe that the above design idea (Zn2+ doping technology) should be suitable for the photocatalytic ability improvement of Bi2MoO6. However, to the best of our knowledge, report on the Zn2+ doped Bi2MoO6 photocatalyst is still quite rare up to now. Hence, we report the successful preparation of sheet-like Zn2+ doped Bi2MoO6 photocatalyst through a facile hydrothermal strategy. The photocatalytic ability of the Zn2+ doped Bi2MoO6 sample was evaluated by decolorizing RhB under the UV-visible-light irradiation, which revealed that the photocatalytic performance of Zn2+ doped Bi2MoO6 could be greatly enhanced compared to pure Bi2MoO6. Furthermore, a possible enhancement mechanism of photocatalytic activity was proposed in detail.

2. Experimental

2.1 Preparation of pure Bi2MoO6 sample

The pure Bi2MoO6 sample was firstly prepared based on previous report.32 0.001 mol of Bi(NO3)3·5H2O and 0.0005 mol of Na2MoO4·2H2O were added in 15 mL of ethylene glycol. After agitating for 2 h, the mixture was transferred into stainless-steel autoclave with a Teflon liner of 40 mL capability, followed by heating treatment at 160 °C for 6 h. When the temperature was cooled to ordinary temperature, the product (denoted as S1) was washed five times repeatedly, and then dried at 60 °C for 7 h.

2.2 Preparation of Zn2+ doped Bi2MoO6 photocatalyst

Firstly, 0.5 g of the as-obtained pure Bi2MoO6 sample (S1) was introduced into a Teflon liner. Then, 30 mL of aqueous solution containing 0.0055 g of zinc acetate and 0.01 g of NaOH was added. After stirring for 0.5 h, the Teflon liner was sealed and heated at 180 °C for 24 h. The subsequent washing and drying procedures of the sample (denoted as S3) are consistent with that of pure Bi2MoO6 sample (Section 2.1). For the preparation of the other two samples, the amount of zinc acetate was controlled at 0.0033 (denoted as S2) and 0.0077 g (denoted as S4), respectively, keeping other experimental parameters unchanged.

2.3 Characterization

The crystal structures of the products were characterized by XRD technology (XRD-6000, Shimadzu, Japan; Cu Kα radiation, λ = 1.5418 Å). SEM and EDS images of the samples were recorded on a field emission scanning electron microanalyser (S-4800, Hitachi, Japan). HRTEM was carried out on a high resolution transmission microscope, employing an accelerating voltage of 200 kV (FEI Tecnai G20, America). FT-IR spectra were obtained on an IR spectrometer (Magna 750, Nicolet, America). Raman spectra were taken at room temperature in the spectral range of 200–1000 cm−1 using a Raman spectromicroscope (LabRAM HR800, Horiba Jobin Yvon, France). The UV-vis diffusion reflectance spectra of the samples were analyzed with a UV-vis spectrophotometer (UV-3600, Shimadzu, Japan). XPS spectrum was recorded on a multifunctional imaging electron spectrometer (Thermo ESCALAB 250XI, America). The surface photovoltage spectra of the products were obtained by surface photovoltage measurement system, which was comprised of a monochromator and a lock-in amplifier (model SR830-DSP) with an optical chopper. The Brunauer–Emmett–Teller (BET) surface area was measured using a micropore & chemisorption analyzer (ASAP 2010, Micromeritics, America). Photoluminescence spectra were recorded employing a fluorescence spectrophotometer (Edinburgh FLSP 920, England). The doped amounts of Zn2+ ions in the samples were analyzed by inductively coupled plasma-atomic emission spectroscopy (J-A1100 from Jarrell-Ash Corp).

2.4 Activity measurement

A 240 W Xe lamp (BL-GHX-Xe-300, Shanghai BILON Co., Ltd., China) was employed as a light source. Firstly, 40 mg of Zn2+ doped Bi2MoO6 was introduced into 10 mg L−1 of RhB aqueous solution (40 mL). Before turning on the light, the above mixed solution was agitated in the dark room for 1 h. Next, ∼2 mL of the reaction solution was got out at suitable intervals. Centrifugation technique was used to dislodge the solid powder. The concentration of RhB solution was calculated by UV-visible spectrophotometry (UV-3600, Shimadzu, Japan).

3. Results and discussion

Fig. 1a depicts the XRD patterns of pure Bi2MoO6 (S1) and Zn2+ doped Bi2MoO6 with different zinc concentrations (S2, S3 and S4). It is observed that all the diffraction patterns corresponding to the above four samples belong to the orthorhombic phase of Bi2MoO6 (JCPDS card no. 76-2388). The diffraction peaks attributed to some impurities fail to be found in Fig. 1a, implying that the existence of Zn2+ ions has no significant impact on the micro-scale structures of Bi2MoO6. The partial enlarged detail of the four XRD patterns in the range of 25°–32° is presented in Fig. 1b. This clearly shows that the (131) diffraction peak slightly shifts toward the higher angles with the increasing of the zinc concentrations. For the rest diffraction peaks, similar phenomenon can also be observed. As is well known, the differences in ionic radius will affect the lattice parameters.33 Usually, the ionic radius of Zn2+ and Bi3+ ions are 0.074 and 0.103 nm, respectively. If some Bi3+ lattice sites in the Bi2MoO6 matrix are replaced by the smaller Zn2+ ions, the lattice parameters of Bi2MoO6 will decrease and lead to the slight shift of diffraction peaks toward higher angles, which is consistent with the XRD observation.34 The calculated cell parameters of various samples (S1, S2, S3 and S4) are shown in Table S1 (ESI).
image file: c6ra06972j-f1.tif
Fig. 1 XRD patterns of pure Bi2MoO6 (S1) and doped Bi2MoO6 samples with different Zn2+ content (S2, S3 and S4).

Raman spectroscopy is a very sensitive test technology to characterize the local structure of micro/nanomaterials. By means of it, some changes of bonding states in the coordination polyhedron can be observed.17 As for pure Bi2MoO6 (S1) and Zn2+ doped sample (S3), many characteristic bands for Bi2MoO6 sample can be easily found: 282, 324, 351, 399, 714, 798 and 844 cm−1 (Fig. 2a). The A1g peak at 798 cm−1 is originated from the symmetric stretch of a MoO6 octahedron, while the A1g modes at 714 and 844 cm−1 show the character of orthorhombic distortions of the MoO6 octahedron in Bi2MoO6. The presence of modes below 400 cm−1 can be attributed to both Bi–O stretches and lattice modes. This result is consistent with previous report.25 Although no obvious shift of Raman peaks can be observed, the relative intensity of Raman peak has changed. As shown in Fig. 2a, the intensity of band at 351 cm−1 remains the same all the time, while the intensity of 282 and 324 cm−1 increase when the Zn2+ ions is introduced into the reaction system. The reason for these differences is unclear up to now. However, it is reasonable to believe that such change may be attributed to the tiny changes in the crystal structure of Bi2MoO6, even if both two samples possess the same orthorhombic phase. Actually, similar experimental phenomenon in Bi2MoO6 system has been demonstrated by Zhu's research team.35


image file: c6ra06972j-f2.tif
Fig. 2 Raman (a) and FT-IR spectra (b) of pure Bi2MoO6 (S1) and doped Bi2MoO6 samples (S3).

Fig. 2b shows the FTIR spectra of undoped (S1) and doped (S3) Bi2MoO6 samples. For S1, the bands located at about 841 and 796 cm−1 can be attributed to the asymmetric and symmetric stretching mode of MoO6 involving vibrations of the apical oxygen atoms, respectively.3 The absorption bands located at 555 and 733 cm−1 are attributed to the bending vibration and asymmetric stretching modes of MoO6, respectively. The band at 438 cm−1 corresponds to the Bi–O stretching mode. In addition, similar FT-IR spectrum can be easily found in S3. However, it should be noted that the intensity of FT-IR spectrum over S3 significantly reduce and some peaks located at 438 and 555 cm−1 shift slightly, which may be ascribed to the interactions between Zn2+ and Bi2MoO6 in the Zn2+ doped Bi2MoO6 system.36

The valence states and compositions of the samples were investigated by XPS. Fig. 3a depicts the survey spectrum of Zn2+ doped Bi2MoO6 (S3). It can be easily concluded that S3 possesses four elements, including Bi, Mo, Zn and O. The Bi 4f peaks located at 158.98 (Bi 4f7/2) and 164.28 eV (Bi 4f5/2) indicate that the valence state of Bi element in the product is 3+ (Fig. 3b).37 In the Mo 3d XPS spectrum (Fig. 3c), the two peaks at 232.26 (Mo 3d5/2) and 235.42 eV (Mo 3d3/2) indicate that Mo is in its [MoO4]2− (Mo6+) state.32 Two weak peaks centered at 1021.61 eV (Zn 2p3/2) and 1044.68 eV (Zn 2p1/2) can be attributed to the Zn 2p binding energy (Fig. 3d), which implies that the presence of Zn element is in the form of Zn2+ oxidation state.38 As shown in Fig. 3e, a strong oxygen 1s peak centered at 529.9 eV corresponds to the oxygen in its O2− oxidation state.32 Moreover, the molar percentage of Zn2+/Bi3+ in the three Zn2+ doped Bi2MoO6 samples can be calculated to be 0.9, 1.5 and 2.0% by inductively coupled plasma-atomic emission spectroscopy, which corresponds to the S2, S3 and S4, respectively. Therefore, it is reasonable to believe that Zn2+ ions have been successfully entered into the Bi2MoO6 matrix and substituted the Bi3+ ions, leading to the formation of Zn2+ doped Bi2MoO6 based on above experimental results.


image file: c6ra06972j-f3.tif
Fig. 3 XPS spectra of doped Bi2MoO6 sample (S3): (a) survey spectrum; (b) Bi 4f; (c) Mo 3d, (d) Zn 2p and (e) O 1s.

To characterize the structural and morphological properties, SEM and TEM analysis technologies were performed. As shown in Fig. 4, it can be seen that the as-obtained S3 are composed of many irregular nanosheets. For the other two doped samples such as S2 and S4, similar sheet-like microstructures can be easily found (Fig. S1, ESI). However, it should be noted that undoped product (S1) is comprised of nanoflakes-assembled Bi2MoO6 hierarchical microspheres (Fig. S2, ESI). This transformation from hierarchical microspheres to nanosheets may be ascribed to the following Zn2+ ions doping treatment.


image file: c6ra06972j-f4.tif
Fig. 4 SEM (a–c) and TEM images (d) of doped Bi2MoO6 sample (S3).

Fig. 5 depicts the HRTEM images of pure Bi2MoO6 (S1) and doped Bi2MoO6 samples with different Zn2+ content (S2, S3 and S4). The clear lattice boundaries in the HRTEM images illustrate the high crystallinity of the Bi2MoO6 nanosheets. The observed lattice fringes of four Bi2MoO6 samples are determined to be 3.156 (S1), 3.148 (S2), 3.138 (S3) and 3.127 Å (S4), respectively, which correspond to the (131) plane of Bi2MoO6 crystal. However, it should be noted that a slight decrease in the lattice spacing of (131) crystal plane can be easily found, which may be ascribed to the lattice contraction due to the substitution of Bi3+ by Zn2+ cations. These experimental facts are in good agreement with the XRD results. Moreover, the EDS spectra of pure Bi2MoO6 (S1) and doped Bi2MoO6 samples (S3) are shown in Fig. S3 (ESI), which further prove the successful preparation of Zn2+ doped Bi2MoO6 nanosheets.


image file: c6ra06972j-f5.tif
Fig. 5 HRTEM images of pure Bi2MoO6 (S1, a) and doped Bi2MoO6 samples with different Zn2+ content (S2: b; S3: c; S4: d).

The UV-vis diffusion reflectance spectra were employed to detect the photo-responsive performance of the samples. The bandgap value of a semiconductor-based photocatalyst can be determined using the following formula:

 
αhν = C(Eg)n(1)
where α, h, ν, C and Eg are the absorption coefficient, Planck constant, light frequency, a constant and band gap, respectively.39 Among them, the exponent n rests with the type of the transition in a semiconductor. For example, n can be assigned a value of 1/2 and 2 for direct inter-band transition and indirect inter-band transition, respectively.40 Therefore, the Eg of the resulting photocatalysts can be calculated to be 2.60, 2.71, 2.73 and 2.82 eV for the above four samples (S1, S2, S3, and S4) based on the intercept of the line on abscissa [(αhν)2 = 0], respectively (Fig. 6).


image file: c6ra06972j-f6.tif
Fig. 6 The UV-vis diffusion reflectance spectra of pure Bi2MoO6 (S1, a) and doped Bi2MoO6 samples with different Zn2+ content (S2: b; S3: c; S4: d).

To estimate the potential application of Zn2+ doped Bi2MoO6 in the decomposition of some toxic pollutants, four samples including S1, S2, S3 and S4, were adopted as photocatalyst to degrade RhB molecule under the irradiation of UV-visible light. The relevant experimental results can be found in Fig. 7a. Blank experiment implies that RhB molecules are relative stable and the decomposition rate can be negligible. As for undoped Bi2MoO6 sample (S1), the degradation value of RhB is 12.9%. When Zn2+ doped Bi2MoO6 samples are adopted as photocatalysts, all three samples (S2, S3 and S4) possess efficient photocatalytic ability under the irradiation of UV-visible light and S3 possesses the highest photocatalytic degradation efficiency (98.9%). Fig. 7b depicts the linear relation between ln(C0/C) and t, which implies that the photocatalytic decomposition process reaction of RhB should follows the pseudo-first order model.3 These experimental results obviously reveal that the reaction rates is 0.006, 0.076, 0.128, 0.033 min−1 for S1, S2, S3 and S4 respectively and S3 possesses the optimal photocatalytic performance, which is almost 21.3, 1.7, and 3.9 times higher than those of S1, S2 and S4, respectively.


image file: c6ra06972j-f7.tif
Fig. 7 (a) Photo-degradation performance of RhB solution as a function of irradiation time under exposure to UV-visible light and (b) kinetics of RhB decolorization in solutions.

To clarify the decomposition process of RhB molecule, many trapping agents including isopropanol, sodium oxalate and benzoquinone were adopted as scavengers of ·OH, h+ and ·O2, respectively.41 When S3 was employed as photocatalyst, blank experiment revealed that the decomposition efficiency of RhB was 98.9% in the absence of any scavenger (Fig. 8). If benzoquinone, isopropanol or sodium oxalate were adopted as trapping agents, respectively, it could be easily found that sodium oxalate (decomposition efficiency: 1.5%) plays the most key role than isopropanol (decomposition efficiency: 81.1%) and benzoquinone (decomposition efficiency: 30.9%). Hence, h+ plays as more crucial roles than ·OH and ·O2 in the decomposition of RhB.


image file: c6ra06972j-f8.tif
Fig. 8 Photocatalytic degradation of RhB over the doped Bi2MoO6 sample (S3) with or without quenchers.

As is well known, the stronger the surface photovoltage response and the weaker photoluminescence intensity, the higher the photoinduced charge carriers and photocatalytic ability.42,43 As depicted in Fig. 9, one can find that Zn2+ doped Bi2MoO6 (S3) exhibits higher surface photovoltage response and lower photoluminescence intensity than undoped Bi2MoO6 (S1), which implies that the photo-generated charges can be easily separated in the Zn2+ doped Bi2MoO6 than pure Bi2MoO6.


image file: c6ra06972j-f9.tif
Fig. 9 The surface photovoltage (a) and photoluminescence spectra (b) of pure Bi2MoO6 (S1) and doped Bi2MoO6 sample (S3).

In addition, the total density of states of valence band for the pure Bi2MoO6 (S1) and doped Bi2MoO6 (S3) were measured by valence band X-ray photoelectron spectroscopy. As shown in Fig. 10, one can find that the valence band potential of Zn2+ doped Bi2MoO6 is estimated to be 0.13 eV more positive than pure Bi2MoO6 sample. That is, a ca. 0.13 eV downward in the valence band of Bi2MoO6 is obtained upon Zn2+ doping, which may imply that the photo-induced holes of Zn2+ doped sample possesses stronger oxidizing ability than that of undoped product.44


image file: c6ra06972j-f10.tif
Fig. 10 The valence band XPS spectra of pure Bi2MoO6 (S1: a) and doped Bi2MoO6 (S3: b).

In order to reveal the reason for activity differences in S1, S2, S3 and S4, the BET surface area of these four photocatalysts were measured (Fig. 11). Compared to pure Bi2MoO6 sample (S1: 63.87 m2 g−1), doped Bi2MoO6 products (S2: 11.21 m2 g−1; S3: 8.87 m2 g−1 and S4: 15.74 m2 g−1) possess a relatively lower specific surface area. Usually, high specific surface area means a strong photocatalytic activity. However, photocatalytic experimental results indicate that S1 owes the worst photocatalytic activity. Therefore, compared to pure Bi2MoO6 sample, the higher photocatalytic activity of Zn2+ doped Bi2MoO6 can be ascribed to below two factors: (1) the recombination of photo-generated charges can be effectively suppressed by Zn2+ doping technology. Under the irradiation of UV-visible light, photo-generated electrons are aroused and transferred from valence band to conduction band across the band gap, leading to the formation of photo-generated electron–hole pairs. e in the conduction band of Zn2+ doped Bi2MoO6 might be trapped by O2 to generate ·O2, which may significantly increase the life of the photo-generated holes.45 As a result, more photo-induced carriers will be involved in the photocatalytic reaction, thereby improving the photocatalytic efficiency. (2) When Zn2+ ions are incorporated into the crystal lattices of Bi2MoO6 and substituted the Bi3+ ions, some changes have taken place in the crystal and band structures of Bi2MoO6 matrix. Thus, the oxidizability of photo-induced holes can be significantly enhanced based on the valence band XPS analysis results. Therefore, photo-generated h+ with higher activity can efficiently oxidize RhB molecule to the final degradation products.


image file: c6ra06972j-f11.tif
Fig. 11 Nitrogen adsorption–desorption isotherm of pure Bi2MoO6 (S1: a) and doped Bi2MoO6 samples with different Zn2+ content (S2: b; S3: c and S4: d).

Moreover, for three doped Bi2MoO6 samples, S3 possesses the lowest specific surface area and the highest photocatalytic activity, which implies that there are other affecting factors on the photocatalytic activities. Recently, many reports have demonstrated that the doping amount can be regarded as a decisive factor associated with the photocatalytic activity in some photocatalytic systems such as Zn2+ doped BiOBr,46 Zn2+ doped BiOCl,47 Sr-doped Bi2WO6,48 and Er3+ doped Bi2MoO6.49 Currently, the generally accepted view is that appropriate amount of doping element can facilitates the rapid migration and separation of photo-generated carriers, leading to the enhancement of photocatalytic ability, which may accounts for the difference of photocatalytic decomposition ability in the above three Zn2+ doped Bi2MoO6 products (S2, S3 and S4).46

4. Conclusion

In summary, Zn2+ doped Bi2MoO6 nanosheets have been constructed through a facile hydrothermal route. Compared with pure Bi2MoO6 product, Zn2+ doped Bi2MoO6 possessed higher photocatalytic ability and S3 owed the highest photocatalytic performance. The elevated photocatalytic efficiency might be ascribed to the easier transfer and separation of the photogenerated carriers. Investigation of the primary reactive species and their effects over the Zn2+ doped Bi2MoO6 photocatalysis revealed that the h+ served as more crucial roles than ·OH and ·O2 in this photocatalytic reaction.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No. 21501002 and 21301005), the Natural Foundation of Anhui Province (No. 1408085QB31), the Open Fund of State Key Laboratory of Coordination Chemistry (SKLCC1604), the Young and Middle-Aged Academic Backbone Training Project of Anhui University of Science and Technology.

References

  1. A. Fujishima and K. Honda, Nature, 1972, 238, 37 CrossRef CAS PubMed.
  2. Y. Q. Qu and X. F. Duan, Chem. Soc. Rev., 2013, 42, 2568 RSC.
  3. J. Di, J. X. Xia, M. X. Ji, H. P. Li, H. Xu, H. M. Li and R. Chen, Nanoscale, 2015, 7, 11433 RSC.
  4. D. Yue, D. M. Chen, Z. H. Wang, H. Ding, R. L. Zong and Y. F. Zhu, Phys. Chem. Chem. Phys., 2014, 16, 26314 RSC.
  5. J. L. Zhang, L. S. Zhang, N. Yu, K. B. Xu, S. J. Li, H. L. Wang and J. S. Liu, RSC Adv., 2015, 5, 75081 RSC.
  6. Y. S. Xu and W. D. Zhang, Dalton Trans., 2013, 42, 1094 RSC.
  7. Z. Q. Hao, L. L. Xu, B. Wei, L. L. Fan, Y. Liu, M. Y. Zhang and H. Gao, RSC Adv., 2015, 5, 12346 RSC.
  8. T. Jing, Y. Dai, W. Wei, X. C. Ma and B. B. Huang, Phys. Chem. Chem. Phys., 2014, 16, 18596 RSC.
  9. S. M. Sun and W. Z. Wang, RSC Adv., 2014, 4, 47136 RSC.
  10. J. R. Ran, J. Zhang, J. G. Yu, M. Jaroniec and S. Z. Qiao, Chem. Soc. Rev., 2014, 43, 7787 RSC.
  11. M. Y. Wang, J. Ioccozia, L. Sun, C. J. Lin and Z. Q. Lin, Energy Environ. Sci., 2014, 7, 2182 CAS.
  12. K. L. Lv, B. Cheng, J. G. Yu and G. Liu, Phys. Chem. Chem. Phys., 2012, 14, 5349 RSC.
  13. W. J. Ong, L. L. Tan, S. P. Chai, S. T. Yong and A. R. Mohamed, Nanoscale, 2014, 6, 1946 RSC.
  14. S. Fukuzumi, K. Ohkubo and T. Suenobu, Acc. Chem. Res., 2014, 47, 1455 CrossRef CAS PubMed.
  15. L. Ren, Y. Z. Li, J. T. Hou, X. J. Zhao and C. X. Pan, ACS Appl. Mater. Interfaces, 2014, 6, 1608 CAS.
  16. L. Ye, J. L. Fu, Z. Xu, R. S. Yuan and Z. H. Li, ACS Appl. Mater. Interfaces, 2014, 6, 3483 CAS.
  17. Z. Dai, F. Qin, H. P. Zhao, F. Tian, Y. L. Liu and R. Chen, Nanoscale, 2015, 7, 11991 RSC.
  18. Y. Zhang, Y. K. Zhu, J. Q. Yu, D. J. Yang, T. W. Ng, P. K. Wong and J. C. Yu, Nanoscale, 2013, 5, 6307 RSC.
  19. Q. Yan, M. Sun, T. Yan, M. M. Li, L. G. Yan, D. Wei and B. Du, RSC Adv., 2015, 5, 17245 RSC.
  20. C. S. Guo, J. Xu, S. F. Wang, Y. Zhang, Y. He and X. C. Li, Catal. Sci. Technol., 2013, 3, 1603 CAS.
  21. Y. Ma, Y. L. Jia, Z. B. Jiao, M. Yang, Y. X. Qi and Y. P. Bi, Chem. Commun., 2015, 51, 6655 RSC.
  22. J. L. Li, X. J. Liu, L. K. Pan, W. Qin and Z. Sun, RSC Adv., 2014, 4, 62387 CAS.
  23. R. Shi, T. G. Xu, L. H. Yan, Y. F. Zhu and J. Zhou, Catal. Sci. Technol., 2013, 3, 1757 CAS.
  24. C. S. Guo, J. Xu, S. F. Wang, L. Li, Y. Zhang and X. C. Li, CrystEngComm, 2012, 14, 3602 RSC.
  25. Y. C. Miao, G. F. Pan, Y. N. Huo and H. X. Li, Dyes Pigm., 2013, 99, 382 CrossRef CAS.
  26. W. Wang, M. O. Tadé and Z. P. Shao, Chem. Soc. Rev., 2015, 44, 5371 RSC.
  27. L. Ran, D. Y. Zhao, X. P. Gao and L. W. Yin, CrystEngComm, 2015, 17, 4225 RSC.
  28. T. Surendar, S. Kumar and V. Shanker, Phys. Chem. Chem. Phys., 2014, 16, 728 RSC.
  29. X. Ding, W. K. Ho, J. Shang and L. Z. Zhang, Appl. Catal., B, 2016, 182, 316 CrossRef CAS.
  30. A. A. Alemi, R. Kashfi and B. Shabani, J. Mol. Catal. A: Chem., 2014, 392, 290 CrossRef CAS.
  31. F. Z. Ren, J. H. Zhang and Y. X. Wang, RSC Adv., 2015, 5, 29058 RSC.
  32. Z. Q. Li, X. T. Chen and Z. L. Xue, CrystEngComm, 2013, 15, 498 RSC.
  33. H. G. Yu, Z. F. Zhu, J. H. Zhou, J. Wang, J. Q. Li and Y. L. Zhang, Appl. Surf. Sci., 2013, 265, 424 CrossRef CAS.
  34. S. N. Zhang, S. J. Zhang and L. M. Song, Appl. Catal., B, 2014, 152–153, 129 CrossRef CAS.
  35. W. J. Luo, J. J. Wang, X. Zhao, Z. Y. Zhao, Z. S. Li and Z. G. Zou, Phys. Chem. Chem. Phys., 2013, 15, 1006 RSC.
  36. L. W. Zhang, T. G. Xu, X. Zhao and Y. F. Zhu, Appl. Catal., B, 2010, 98, 138 CrossRef CAS.
  37. M. Wang, Z. Y. Qiao, M. H. Fang, Z. H. Huang, Y. G. Liu, X. W. Wu, C. Tang, H. Tang and H. K. Zhu, RSC Adv., 2015, 5, 94887 RSC.
  38. L. Zhang, X. Zhang, Y. Q. Huang, C. L. Pan, J. S. Hu and C. M. Hou, RSC Adv., 2015, 5, 30239 RSC.
  39. L. Zhang, X. F. Cao, X. T. Chen and Z. L. Xue, J. Colloid Interface Sci., 2011, 354, 630 CrossRef CAS PubMed.
  40. L. Y. Huang, H. Xu, Y. P. Li, H. M. Li, X. N. Cheng, J. X. Xia, Y. G. Xu and G. B. Cai, Dalton Trans., 2013, 42, 8606 RSC.
  41. Y. Xu, S. C. Xu, S. Wang, Y. X. Zhang and G. H. Li, Dalton Trans., 2014, 43, 479 RSC.
  42. G. H. Tian, Y. J. Chen, J. Zhou, C. G. Tian, R. Li, C. J. Wang and H. G. Fu, CrystEngComm, 2014, 16, 842 RSC.
  43. C. Liu, C. M. Li, X. D. Fu, F. Raziq, Y. Qu and L. Q. Jing, RSC Adv., 2015, 5, 37275 RSC.
  44. J. Q. Yan, Y. X. Zhang, S. Z. Liu, G. J. Wu, L. D. Li and N. J. Guan, J. Mater. Chem. A, 2015, 3, 21434 CAS.
  45. M. N Guo, Y. Wang, Q. L. He, W. J. Wang, W. M. Wang, Z. Y. Fu and H. Wang, RSC Adv., 2015, 5, 58633 RSC.
  46. X. C. Song, Y. F. Zheng, H. Y. Yin, J. N. Liu and X. D. Ruan, New J. Chem., 2016, 40, 130 RSC.
  47. W. T. Li, W. Z. Huang, H. Zhou, H. Y. Yin, Y. F. Zheng and X. C. Song, J. Alloys Compd., 2015, 638, 148 CrossRef CAS.
  48. Y. Liu, W. M. Wang, Z. Y. Fu, H. Wang, Y. C. Wang and J. Y. Zhang, Mater. Sci. Eng., B, 2011, 176, 1264 CrossRef CAS.
  49. T. F. Zhou, J. C. Hu and J. L. Li, Appl. Catal., B, 2011, 110, 221 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra06972j

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