Tuning the electronic and structural properties of WO3 nanocrystals by varying transition metal tungstate precursors

Sara Rahimnejadabc, Jing Hui Heab, Wei Chen*adb, Kai Wu*eb and Guo Qin Xu*ab
aDepartment of Chemistry, National University of Singapore, 3 Science Drive 3, 117543, Singapore. E-mail: chmxugq@nus.edu.sg; Tel: +65-65163595
bSPURc, 1 CREATE Way, #15-01, CREATE Tower, Singapore 138602, Singapore
cDepartment of Chemistry, Yadegar-e-Imam Khomeini (RAH) Branch, Islamic Azad University, 18155-144, Tehran, Iran
dDepartment of Physics, National University of Singapore, 2 Science Drive 3, 117542, Singapore
eState Key Laboratory for Structural Chemistry of Unstable and Stable Species, College of Chemistry and Molecular Engineering, Peking University, Beijing, 100871, P. R. China

Received 17th September 2014 , Accepted 10th November 2014

First published on 11th November 2014


Abstract

Oxygen vacancy is one type of the most important defects affecting the photocatalytic performance of WO3. In this paper, WO3 nanoplates with a high density of oxygen vacancies were synthesized from MWO4 (M = Zn, Cd, Co, Ni) precursors using a sacrificial template method. The structures and morphologies of WO3 nanoplates were investigated by field emission scanning electron microscopy (FE-SEM), high resolution Transmission Electron Microscopy (HRTEM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Brunauer–Emmett–Teller (BET) analysis, Photoluminescence (PL), Diffuse Reflectance UV-Vis (DRS UV-Vis) and Time-correlated single-photon counting (TCSPC). The metal tungstates were found to not only act as the precursors but also as structure-directing agents during the growth of WO3 nanoplates. XRD data revealed that two phases of WO3·xH2O (x = 1 or 2) were obtained after acid treatment of MWO4. WO3 nanoplates derived from NiWO4 were found to have the highest ratio of WO3·2H2O, highest concentration of oxygen vacancies, narrowest band gap, longest electron–hole recombination time, and in turn the highest rate of photodegradation of azo dye methylene blue. These results show that the structural, electronic and photocatalytic properties of synthesized WO3 nanoplates can be tuned by varying the transition metal tungstate precursors.


1. Introduction

Metal oxides have been widely studied due to their optical, magnetic, electronic and photocatalytic applications.1–3 In particular, WO3 is an important semiconductor material, which has been widely used in heterogeneous catalysis,4,5 gas sensors,6,7 electrochromic8/photochromic9/field emission10,11/solar energy devices12 and photocatalysis.13–22 Among these fields, the photocatalytic property was extensively investigated23–27 due to its promising application for remediation of hazardous waste,28 water oxidation29–31 and CO2 reduction32 because WO3 has a high stability in acidic media,33 light-harvesting ability to visible light34 and long-lasting energy storage ability.35 However, the fast recombination rate of photogenerated electron (e)–hole (h+) and rather low conduction band has inherently limited its photocatalytic effiency.36 Several efforts were made to enhance the photocatalytic activity of WO3 by tailoring its particle size,37 crystal structure38 and composition.17 Recently, Yamakata et al. reported the relationship between the size of WO3 particles and their photocatalytic efficiency.37 Surprisingly, they discovered that large WO3 particles with low surface area to volume ratios are suitable for photocatalytic oxygen evolution because of the long-lived photogenerated holes. This strongly suggest that shape and size of WO3 particles may not be the critical factor.

On the other hand, surface oxygen vacancies are intrinsic defects of metal oxides.39 They are the most reactive sites on the surface and able to modify the electronic and chemical properties of the surface and greatly prolong life time of photoexcited carriers.40 Therefore the amount of e and h+ on the photocatalytic surface could be a key factor in determining the photocatalytic reaction rate.41 Liu et al. also reported that the reactivity of a photocatalyst is mainly influenced by its surface geometric and electronic structures. Thus tuning the surfaces structures of WO3 photocatalysts41 to generate high-density e and h+ is essential to optimize their photocatalytic performances for targeted reactions.42 Nowadays surface oxygen vacancy engineering is able to effectively enhance the photocatalytic performances of metal oxides.43

Most of WO3 nanocrystals were synthesized from aqueous solution of transition metal tungstates.44,45 Many reports have confirmed the effects of preparation methods, the nature of supports46–50 and tungsten precursors51 on the efficiency of WO3 photocatalysts. However the influence of transition metal tungstate sources on fabricating WO3 with different optical, electronic and photocatalytic properties has not been systematically studied yet. The present study aims to investigate the effects of precursors on the surface structures and photocatalytic behaviours of WO3. The monoclinic WO3 nanoplates with different densities of oxygen vacancies were prepared from transition metal tungstates MWO4 (M = Zn, Cd, Co, Ni). Their photocatalytic efficiencies were evaluated by photo-degradation of methylene blue and are correlated with the changes of defects density in the samples prepared via varying the transition metal tungstate sources.

2. Experimental section

2.1 Chemicals and materials

Tungsten oxide (WO3 nanopowder, Sigma Aldrich), methylene blue (MB, Alfa-Aesar), nickel(II) nitrate hexahydrate 99.999%, zinc nitrate hexahydrate 99% (Sigma Aldrich), sodium tungstate dehydrate 99% (Sigma Aldrich), cadmium nitrate tetrahydrate (Sigma Aldrich), cobalt(II) nitrate hexahydrate 99%, nitric acid 70% (Sigma Aldrich) were used as received without further purification. Ultrapure deionized water was prepared by millipore purification system.

2.2 Preparation of MWO4 (M = Zn, Cd, Co, Ni)

M (NO3)2 and Na2WO4 aqueous solutions with a molar ratio 1[thin space (1/6-em)]:[thin space (1/6-em)]1 were mixed at room temperature. The resulting precipitates were transferred into a Teflon-lined stainless steel autoclave at 160 °C for 24 h. After hydrothermal treatment, the MWO4 powders collected after filtration and washing with distilled water were dried in air at 80 °C overnight.

2.3 Preparation of WO3

The synthesized MWO4 (M = Zn, Cd, Co, Ni) products from previous procedure were immersed in 8 mol L−1 HNO3 solution. The duration of acid treatment was varied from 24 h to 72 h depending on tungstate sources. Upon filtration and washing with distilled water, the acid treated products were calcined in a furnace at 500 °C for 4 h in air.

2.4 Characterization

The morphology and microstructure of the samples were determined by SEM (JEOL-6701F) and TEM (JEM-3010). XRD patterns of the samples were recorded on a Panalytical X'pert XRD system using Cu Kα radiation. The optical absorbance spectra were recorded by a UV-Visible spectrophotometer (Shimadzu UV-2600). The chemical state and valance band spectra of tungsten trioxides were characterized by X-ray photoelectron spectroscopy. Mg Kα (1253.6 eV) was utilized as the excitation light source and the signal was recorded by Omicron EA 125 at a normal emission angle at room temperature. The Brunauer–Emmett–Teller (BET) surface area was determined by nitrogen adsorption–desorption isotherm measurement at 77 K using Micromeritics ASAP 2020. Photoluminescence (PL) Spectra were measured on a HORIBA Jobin Yvon S.A.S. Fluoromax-4 spectrofluorometer at an excitation wavelength of 320 nm. Fluorescence lifetime was measured on Flurolog HR 320 HORIBA Jobin Yvon S.A.S. time-correlated single photon counting (TCSPC) instrument.

2.5 Photocatalysis measurement

The photocatalytic activities of the prepared samples were evaluated by degradation of methylene blue under simulated solar light irradiation. In a typical run, 50 mg of photocatalysts was dispersed in 50 mL of methylene blue aqueous solution (5 mg L−1). The solution was continuously stirred in the dark for 1 h to establish adsorption–desorption equilibrium before irradiation. The solution was then irradiated under illumination of a 500 W Xe lamp with light intensity of 100 mW cm−2 was used as light source for simulated solar light. At each interval of 2 min, sample aliquots were exacted from the reactor, followed by centrifugation (13[thin space (1/6-em)]000 rpm for 5 min) and filtering through a 0.45 μm PTFE syringe filter (Millipore) to remove the photocatalysts. The concentration of methylene blue was monitored by UV-Vis Spectrophotometer at the maximum absorbance peak (664 nm).

3. Result and discussion

3.1 Structure and morphology

We first studied how the structures and compositions of WO3 products are correlated with metal tungstate sources while keeping the other conditions constant. The XRD patterns of the uncalcined samples synthesized by different precursors are shown in Fig. 1. In the case of ZnWO4, the acid treatment results in pure orthorhombic tungsten oxide hydrate WO3·H2O phase (JCPDS no. 43-0679) with lattice constants a = 0.5238 nm, b = 1.7040 nm, c = 0.5120 nm. The other three precursors (MWO4, M = Cd, Co, Ni) result in WO3·2H2O phase (JCPDS no. 16-0166) with lattice constants of a = 0.7450 nm, b = 0.6920 nm, c = 0.3720 nm mixed with WO3·H2O. This could be seen from the peak assignment marked by diamond and star symbols. Among these three precursors, NiWO4 has the highest ratio of monoclinic WO3·2H2O, which will correlate with the catalytic performance, as we will demonstrate later.
image file: c4ra10650d-f1.tif
Fig. 1 X-ray diffraction (XRD) patterns of the WO3·xH2O products synthesized by different precursors before calcination.

The above showed that both monoclinic WO3·2H2O and orthorhombic WO3·H2O phases could be obtained by varying precursors. Further calcination at 500 °C converts the crystal phase of these samples to pure monoclinic WO3 (JCPDS no. 43-1035) as confirmed from the XRD results in Fig. 2. The normalized XRD peaks of the sample (Ni/WO3) derived from NiWO4 have the lowest intensities. This indicates a lower crystallinity, correlating to high oxygen-deficient surfaces.52


image file: c4ra10650d-f2.tif
Fig. 2 X-ray diffraction (XRD) patterns of the products synthesized by different precursors after calcination.

Fig. 3 shows the representative FESEM images of WO3 samples synthesized from different precursors. The result indicates that all products are of plate structure but the sizes of plates depend on precursors. In Fig. 3a, a typical plate is about 270 nm long and 220 nm wide. When the precursor was changed, the plates are of quasi-quadrangular shape (Fig. 3b–d). It is well proved that solvents, impurities and additives in solution can substantially influence the ultimate shape of the crystals by tuning the growth rate and orientation of the crystals.53 This effect can be explained by different time duration for completing the process of replacing precursor cation by proton for each precursor, possibly affecting the rate of nucleation, growing, agglomeration and consequently ultimate shape.


image file: c4ra10650d-f3.tif
Fig. 3 FESEM images of the products synthesized by different precursors: (a) Zn/WO3 (b) Cd/WO3 (c) Co/WO3 (d) Ni/WO3.

The detailed structural and morphological characteristics of tungsten oxides were further investigated by HR-TEM. Fig. 4d shows that the synthesized Ni/WO3 is not well crystallized. The HRTEM image shows the lattice fringes of 0.366 nm which can be readily assigned to (200) crystal planes. Fig. 4a–c show HR-TEM images of WO3 derived from MWO4 (M = Zn, Cd, Co). The result shows clear crystal lattice, which can be assigned to the preferential orientation at (200) and (020) directions. Since (200) and (020) of WO3 have equal surface energy and show the same adsorption ability,54 no correlation was observed between crystal facets of WO3 and dye degradation.


image file: c4ra10650d-f4.tif
Fig. 4 HRTEM images of WO3 samples by different precursors. (a) Zn/WO3 (b) Cd/WO3 (c) Co/WO3 (d) Ni/WO3.

BET surface area measurements of the samples were carried out at liquid nitrogen temperature, and the corresponding values are summarized in Table 1. The specific surface area of the sample varies insignificantly, indicated that the WO3 particle size was not crucial for photodegradation of azo dyes methylene blue.

Table 1 Comparison of physical properties of different photocatalysts
Sample BET surface area (m2 g−1) BJH adsorption average pore diameter (nm)
Zn/WO3 20.5 8.0
Cd/WO3 17.8 8.3
Co/WO3 15.2 8.1
Ni/WO3 18.5 10.2


3.2 Electronic and optical properties

To gain insight into the effect of different metal tungstates on the electronic and optical properties of WO3 nanoplates, the chemical states and surface chemical compositions of the resultant crystals were determined by XPS. The O1s spectra of the WO3 samples in Fig. 5 can be described as the superposition of three peaks located at 530.0, 531.2 and 533.0 eV. The O1s peak at 533.0 eV indicates loosely bound oxygen, which is from H2O molecules on the surface of WO3. The peak at 530.0 eV is attributed to the O2− ions mainly from bulk WO3. The intensity at 531.2 eV is associated with O2− in the oxygen deficient regions with the matrix of WO3.55 In the oxygen deficient surface region; OH groups are bonded to the metal cations to maintain the charge balance. Thus, the O1s intensity of OH is related to the oxygen vacancy density. Since the concentration of lattice O2− should not be sensitive to surface electronic structures; we normalize other peaks to this peak at 530.0 eV.56 As Table 2 shows, the relative peak intensity at 531.2 eV of Ni/WO3 is the highest among the four samples, correlating with the concentration of oxygen vacancies in WO3 samples.
image file: c4ra10650d-f5.tif
Fig. 5 XPS Spectra of the O1s region registered for WO3 samples, fitted with three components, at 530.0, 531.2, 533.0 eV for O2−:OL (lattice oxygen), OH:OV (oxygen vacancy or defect), H2O:Oc (chemisorbed oxygen species) respectively.
Table 2 O1s signals for WO3 based different precursors with the relative oxygen species amounts (the amounts of surface O2−/OH/H2O species were determined by XPS from the O1s peak (530.0, 531.3, 533.0 eV for O2−, OH, H2O respectively)
Peak position (eV) Zn/WO3 Cd/WO3 Co/WO3 Ni/WO3
530.0 1.0 1.0 1.0 1.0
531.3 0.19 0.40 0.48 0.62
533.0 0.11 0.19 0.18 0.19


To evaluate the effect of oxygen vacancy on the energy gap of the WO3 samples prepared under different conditions, the optical properties of WO3 samples were probed using UV-Visible diffuse reflectance spectroscopy.

The DR UV-Vis spectra of WO3 derived from MWO4 (M = Zn, Cd, Co, Ni) are shown in Fig. 6. It could be found that the optical absorption edge was estimated to be ∼460 nm for WO3 derived from ZnWO4 and ∼510 nm for WO3 derived from NiWO4, respectively. Combined with XPS results, with the increase of surface oxygen vacancy, the absorption edge of WO3 gradually moved to longer wavelength.55,57 We concluded that WO3 derived from NiWO4 has the highest O vacancy concentrations and thus the narrowest band gap as well as the best visible light response.


image file: c4ra10650d-f6.tif
Fig. 6 Diffuse reflectance UV-Vis of WO3 samples obtained from different precursors.

The relevant PL emission spectra and fluorescence life times of WO3 samples were investigated and presented in Fig. 7 and 8. All samples have similar emission profiles. The blue emission peaks at 409 and 421 nm can be assigned to oxygen vacancies in WO3.58 The WO3 sample derived from NiWO4 has a significantly lower luminescence intensity compared to other WO3 samples. Thus in this case we can observe the much lower PL intensity, which indicates the lower recombination rate of photo-induced electron–hole pair.


image file: c4ra10650d-f7.tif
Fig. 7 PL spectra (using excitation at 340 nm) of WO3 samples.

image file: c4ra10650d-f8.tif
Fig. 8 Time-correlated single photon counting of fluorescence lifetimes of WO3 samples.

The much lower PL intensity indicates the lower recombination rate of photo-induced electron–hole pair and the presence of oxygen vacancies would facilitate the charge separation process. As reported, the oxygen vacancies are demonstrated to be electron donors in semiconductor59 and can be considered to enhance the donor density in heterogeneous WO3·xH2O/WO3 sample derived from NiWO4.

The significant difference between the WO3 samples is in the PL decay lifetime, as shown in Table 3. The PL decay of metal oxides comes from the recombination of nonradiative (T1) and radiative (T2) processes. The radiative process originates from the recombination of photogenerated electrons and holes.60 Therefore we calculated the T2 values of different samples through double exponential decay fitting. The T2 values of as prepared are all greater than that of the commercial products (5.78 ns). In addition, Ni/WO3 has the longest radiative decay time (8.84 ns). This longest lifetime corresponding to the slowest PL decay clearly shows a prolonged e–h+ recombination process.

Table 3 The calculated decay time constant for the commercial and as-prepared WO3 samples
Sample WO3 commercial Zn/WO3 Cd/WO3 Co/WO3 Ni/WO3
T2 (ns) 5.78 4.64 6.23 6.84 8.84


Correlating with the formation mechanism of WO3 samples, WO3·xH2O was formed by the acid treatment and survived somehow after calcination. This facilitates the transfer of electrons to the surface of photocatalysts due to its highly conducting properties. Therefore the probability of photogenerated e–h+ recombination decreases greatly by increasing the lifetime of photoexcited holes. On the other hand, the higher amount of O vacancies on the WO3 sample from WO3·xH2O which was formed after NiWO4 acid treatment can increase the amount of charge carrier trapping sites on the surface which preventing the rate of e–h+ recombination and increase the life time of e and h+ hole recombination.

The photocatalytic activities of resultant WO3 crystals were evaluated by monitoring photodegradation of azo dye methylene blue. Among the WO3 samples prepared from different precursors, the Ni/WO3 exhibits the highest reaction rate; the rate constant is 5.01 × 10−2 min−1 as shown in Fig. 9. The rates constant were 3.49 × 10−2, 2.46 × 10−2, and 2.41 × 10−2 min−1 for other three samples. Correlating with Table 1, we found specific surface area is not the sole determining factor. However, the photocatalytic degradation is a pseudo first-order reaction judging from the good linearity of the curves in Fig. 9. The reaction rate constants were found to have the same trend as T2 values in PL spectra and the O1s peak intensity at 531.2 eV in XPS (Fig. 10) when the transition metal is varied from Zn to Cd, Co and Ni. Considering their physical meaning of T2 and 531.2 eV peak, this clearly demonstrates that the photoefficiency correlates with the e–h+ combination time and the concentration of surface oxygen vacancies.


image file: c4ra10650d-f9.tif
Fig. 9 Photocatalytic performance of methylene blue over WO3 samples. The reaction rate constants (k) were calculated based on pseudo-first order reaction kinetics. The light source was a 500 W Xenon lamp.

image file: c4ra10650d-f10.tif
Fig. 10 Correlation of photoefficiency with the radiative time (T2) in PL process (Table 3) and concentration of surface oxygen vacancies from XPS O1s peak at 531.3 eV (Table 2).

The relation between the different metal sources and reaction rate constant may mainly come from the crystallographies of uncalcined samples. In the process of acid treatment on MWO4 precursors, the hydrogen atoms exchange with lattice metal ions of MWO4 crystals, forming WO3·xH2O (x = 1 or 2). Up on annealing these samples, the contained water molecules desorb and WO3 nanoplates are formed. The crystalliferous water as in WO3·H2O has weak interaction and is facile to release with no distortion on the WO6 octahedrons during conversion to WO3. Therefore the formed WO3 nanoplates have less concentration of oxygen defects. In contrast, in WO3·2H2O (in fact [WO3 (H2O)]·H2O),61 there is one coordination water per W octahedron. This water molecule serves as a ligand, which is in fact part of the WO6 octahedrons. Desorption of this coordinated water should leave a ligand defect, which will finally become an oxygen defect given no external oxygen atoms healing this vacancies during the formation of WO3 nanoplates. The comparison between WO3·2H2O and WO3·H2O will explain Zn/WO3, which is from the pure WO3·H2O, has the lowest concentration of surface oxygen defects and thus the lowest photocatalytic efficiency and Ni/WO3 is of the highest. Nevertheless, the uncalcined samples derived from Co, although of higher ratio of WO3·H2O than that of Cd (Fig. 1) is unexpectedly of the higher oxygen vacancies. Therefore, other properties from the metal ions, rather than the sole crystallography, also play roles to control the oxygen vacancies, which are interesting in our future study.

4. Conclusions

We have investigated the relationship between photocatalytic activities of WO3 samples derived from different transition metal tungstate sources (M = Zn, Cd, Co, Ni) for dye degradation. We found that the concentration of oxygen vacancies of WO3 sample generated from precursors is the decisive factor to the photocatalytic efficiency. The sample derived from NiWO4 is of the highest ratio of OH to O2−, longest PL lifetime decay and thus the highest photocatalytic efficiency. Our study may inspire the study of preparing precursors to optimize the photocatalytic efficiency of WO3.

Acknowledgements

We are grateful for the financial support from the CREATE-SPURc project, National Research Foundation, Singapore (R143-001-205-592).

Notes and references

  1. R. Ferrando, J. Jellinek and R. L. Johnston, Chem. Rev., 2008, 108, 845 CrossRef CAS PubMed.
  2. M. Fernández-García, A. Martínez-Arias, J. C. Hanson and J. A. Rodriguez, Chem. Rev., 2004, 104, 4063 CrossRef PubMed.
  3. D. V. Talapin, J.-S. Lee, M. V. Kovalenko and E. V. Shevchenko, Chem. Rev., 2009, 110, 389 CrossRef PubMed.
  4. H. J. Liu, S. J. Huang, L. Zhang, S. L. Liu, W. J. Xin and L. Y. Xu, Catal. Commun., 2009, 10, 544 CrossRef CAS PubMed.
  5. D. H. Koo, M. Kim and S. Chang, Org. Lett., 2005, 7, 5015 CrossRef CAS PubMed.
  6. A. Ponzoni, E. Comini, G. Sberveglieri, J. Zhou, S. Zhi Deng, N. Sh Xu, Y. Ding and Z. L. Wang, Appl. Phys. Lett., 2006, 88, 203101 CrossRef PubMed.
  7. J. Ma, J. Zhang, S. Wang, T. Wang, J. Lian and X. Duan, J. Phys. Chem. C, 2011, 115, 18157 CAS.
  8. J. H. Zhu, S. Y. Wei, M. Alexander, T. D. Dang, T. C. Ho and Z. H. Guo, Adv. Funct. Mater., 2010, 20, 3076 CrossRef CAS.
  9. C. S. Blackman and I. P. Parkin, Chem. Mater., 2005, 17, 1583 CrossRef CAS.
  10. Y. B. Li, Y. Bando and D. Golberg, Adv. Mater., 2003, 15, 1294 CrossRef CAS.
  11. J. G. Liu, Z. J. Zhang, Y. Zhao, X. Su, S. Liu and E. G. Wang, Small, 2005, 1, 310 CrossRef CAS PubMed.
  12. H. D. Zheng, Y. Tachibana and K. Kalantar-zadeh, Langmuir, 2010, 26, 19148 CrossRef CAS PubMed.
  13. J. Shi, G. Hu, R. Cong, H. Bu and N. Dai, New J. Chem., 2013, 37, 1538 RSC.
  14. K. Sayama, H. Hayashi, T. Arai, M. Yanagida, T. Gunji and H. Sugihara, Appl. Catal., B, 2010, 94, 150 CrossRef CAS PubMed.
  15. D. Chen and J. Ye, Adv. Funct. Mater., 2008, 18, 1922 CrossRef CAS.
  16. Z. G. Zhao and M. Miyauchi, Angew. Chem., Int. Ed., 2008, 47, 7051 CrossRef CAS PubMed.
  17. I. M. Szilagyi, B. Forizs, O. Rosseler, A. Szegedi, P. Nemeth, P. Kiraly, G. Tarkanyi, B. Vajna, K. Varga-Josepovits, K. Laszlo, A. L. Toth, P. Baranyai and M. Leskela, J. Catal., 2012, 294, 119 CrossRef CAS PubMed.
  18. Y. Guo, X. Quan, N. Lu, H. Zhao and S. Chen, Environ. Sci. Technol., 2007, 41, 4422 CrossRef CAS.
  19. D. Sanchez-Martinez, A. Martinez-de la Cruz and E. Lopez-Cuellar, Mater. Res. Bull., 2013, 48, 691 CrossRef CAS PubMed.
  20. J. Huang, X. Xu, C. Gu, G. Fu, W. Wang and J. Liu, Mater. Res. Bull., 2012, 47, 3224 CrossRef CAS PubMed.
  21. S. Rajagopal, D. Nataraj, D. Mangalaraj, Y. Djaoued, J. Robichaud and O. Y. Khyzhun, Nanoscale Res. Lett., 2009, 4, 1335 CrossRef CAS PubMed.
  22. L. Zhang, X. Tang, Z. Lu, Z. Wang, L. Li and Y. Xiao, Appl. Surf. Sci., 2011, 258, 1719 CrossRef CAS PubMed.
  23. Y. Liu, Q. Li, S. Gao and J. Ku Shang, CrystEngComm, 2014, 16, 7493 RSC.
  24. M. Mokhtar, S. A. Ahmed and K. S. Khairou, Appl. Catal., B, 2014, 150–151, 63 Search PubMed.
  25. H. Zhang, J. Yang, D. Li, W. Guo, Q. Qin, L. Zhu and W. Zheng, Appl. Surf. Sci., 2014, 305, 247 CrossRef PubMed.
  26. K. Jothivenkatachalam, S. Prabhu, A. Nithya and K. Jeganathan, RSC Adv., 2014, 4, 21221 RSC.
  27. C. Gomez-Solis, D. Sanchez-Martinez, I. Juarez-Ramirez, A. Martinez-de la Cruz and L. M. Torres-Martinez, J. Photochem. Photobiol., A, 2013, 262, 28 CrossRef CAS PubMed.
  28. M. Aslam, I. M. I. Ismail, S. Chandrasekaran and A. Hameed, J. Hazard. Mater., 2014, 276, 120 CrossRef CAS PubMed.
  29. G. Liu, J. Han, X. Zhou, L. Huang, F. Zhang, X. Wang, C. Ding, X. Zheng, H. Han and C. Li, J. Catal., 2013, 307, 148 CrossRef CAS PubMed.
  30. H. Ishihara, G. K. Kannarpady, K. R. Khedir, J. Woo, S. Trigwell and A. S. Biris, Phys. Chem. Chem. Phys., 2011, 13, 19553 RSC.
  31. K. C. Leonard, K. M. Nam, H. C. Lee, S. H. Kang, H. S. Park and A. J. Bard, J. Phys. Chem. C, 2013, 117, 15901 CAS.
  32. X. Chen, Y. Zhou, Q. Liu, Z. Li, J. Liu and Z. Zou, ACS Appl. Mater. Interfaces, 2012, 4, 3372 CAS.
  33. J. M. Spurgeon, J. M. Velazques and M. T. McDowell, Phys. Chem. Chem. Phys., 2014, 16, 3623 RSC.
  34. A. J. E. Rettie, K. C. Klavetter, J. F. Lin, A. Dolocan, H. Celio, A. Ishiekwene, H. L. Bolton, K. N. Pearson, N. T. Hahn and C. B. Mmullins, Chem. Mater., 2014, 26, 1670 CrossRef CAS.
  35. J. Li, Y. Liu, Z. Zhu, G. Zhang, T. Zou, Z. Zou, S. Zhang, D. Zeng and C. Xie, Sci. Rep., 2013, 3, 2409,  DOI:10.1038/srep02409.
  36. W. Gongming, L. Yichuan, W. Hanyu, Y. Xunyu, W. Changchun, Z. Z. Jin and L. Yat, Energy Environ. Sci., 2012, 5, 6180 Search PubMed.
  37. F. Amano, E. Ishinaga and A. Yamakata, J. Phys. Chem. C, 2013, 117, 22584 CAS.
  38. M. R. Waller, T. K. Townsend, J. Zhao, E. M. Sabio, R. L. Chamousis, N. D. Browning and F. E. Osterloh, Chem. Mater., 2012, 24, 698 CrossRef CAS.
  39. W. Tu, Y. Zhou and Z. Zou, Adv. Mater., 2014, 26, 4607 CrossRef CAS PubMed.
  40. P. Li, Y. Zhou, W. Tu, R. Wang, C. Zhang, Q. Liu, H. Li, Z. Li, H. Dai, J. Wang, S. Yan and Z. Zou, CrystEngComm, 2013, 15, 9855 RSC.
  41. X. Y. Zhang, J. Qin, Y. Xue, P. Yu, B. Zhang, L. Wang and R. Liu, Sci. Rep., 2014, 4, 4596,  DOI:10.1038/srep04596.
  42. Y. P. Xie, G. Liu, L. Yin and H. M. Cheng, J. Mater. Chem., 2012, 22, 6746 RSC.
  43. K. Xie, N. Umezawa, N. Zhang, P. Reunchan, Y. Zhang and J. Ye, Energy Environ. Sci., 2011, 4, 4211 CAS.
  44. Z. Gu, T. Zhai, B. Gao, X. Sheng, Y. Wang, H. Fu, Y. Ma and J. Yao, J. Phys. Chem. B, 2006, 110, 23829 CrossRef CAS PubMed.
  45. A. Sonia, Y. Djaoued, B. Subramanian, R. Jacques and M. Eric, Mater. Chem. Phys., 2012, 136, 80 CrossRef CAS PubMed.
  46. J. M. Wang, E. Khoo, P. S. Lee and J. Ma, J. Phys. Chem. C, 2008, 112, 14306 CAS.
  47. Z. Wang, S. Zhou and L. Wu, Adv. Funct. Mater., 2007, 17, 1790 CrossRef CAS.
  48. S. Rajagopal, D. Nataraj, D. Mangalaraj, Y. Djaoued, J. Robichaud and O. Y. Khyzhun, Nanoscale Res. Lett., 2009, 4, 1335 CrossRef CAS PubMed.
  49. X. Gao, C. Yang, F. Xiao, Y. Zhu, J. Wang and S. Xintai, Mater. Lett., 2012, 84, 151 CrossRef CAS PubMed.
  50. L. Zhang, X. Tang, Z. Lu, Z. Wang, L. Li and Y. Xiao, Appl. Surf. Sci., 2011, 258, 1719 CrossRef CAS PubMed.
  51. Y. Oaki and H. Imai, Adv. Mater., 2006, 18, 1807 CrossRef CAS.
  52. M. Lorenz, R. Bottcher, S. Friedlander, A. Poppl and D. Spemann, J. Mater. Chem. C, 2014, 2, 4947 RSC.
  53. G. Liu, J. C. Yu, G. Q. Lu and H. M. Cheng, Chem. Commun., 2011, 47, 6763 RSC.
  54. D. Zhang, S. Wang, J. Zhu, H. Li and Y. Lu, Appl. Catal., B, 2012, 123–124, 398 CrossRef CAS PubMed.
  55. J. Wang, Z. Wang, B. Huang, Y. Ma, Y. Liu, X. Qin, X. Zhang and Y. Dai, ACS Appl. Mater. Interfaces, 2012, 4, 4024 CAS.
  56. G. Wang, Y. Ling, H. Wang, X. Yang, C. Wang, J. Z. Zhang and Y. Li, Energy Environ. Sci., 2012, 5, 6180 CAS.
  57. J. Cao, B. Luo, H. Lin, B. Xu and S. Chen, Appl. Catal., B, 2012, 111–112, 288 CrossRef CAS PubMed.
  58. W. Hu, Y. Zhao, Z. Liu, C. W. Dunnill, D. H. Gregory and Y. Zhu, Chem. Mater., 2008, 20, 5657 CrossRef CAS.
  59. Y. Y. Tay, T. T. Tan, M. H. Liang, F. Boey and S. Li, Phys. Chem. Chem. Phys., 2010, 12, 6008 RSC.
  60. F. Wang, X. Chen, X. Hu, K. S. Wong and J. C. Yu, Sep. Purif. Technol., 2012, 91, 67 CrossRef CAS PubMed.
  61. J. A. Kaduk, Powder Diffr., 2008, 23, 157–158 CrossRef.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.