Hierarchical TiO2 spheres decorated with Au nanoparticles for visible light hydrogen production

Guoqiang Zhangab, Zhao Zhaoab, Huaqiao Tana, Haifeng Zhaoa, Dan Quab, Min Zhenga, Weixing Yu*c and Zaicheng Sun*a
aState Key Laboratory of Luminescence and Applications, Changchun Institute of Optics, Fine Mechanics and Physics, Changchun, 130033, Jilin, P. R. China. E-mail: sunzc@ciomp.ac.cn
bUniversity of Chinese Academy of Sciences, Beijing, P. R. China
cInsititue of Micro and Nano Optics, College of Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China

Received 5th December 2014 , Accepted 18th February 2015

First published on 18th February 2015


Abstract

Hierarchical TiO2 spheres composed of nanosheets are successfully synthesized via a simple solvothermal route. TiO2 spheres with high surface area (∼100 m2 g−1) exhibit excellent photocatalytic activity. Au nanoparticles are loaded on the surface of TiO2 nanosheets through anchor molecules – thiolglycolic acid. The LSPR absorption band at ∼550 nm of Au nanoparticles is clearly observed in the diffusion reflective UV-Vis spectra. H2 production results show that the TiO2 spheres have higher photocatalytic activity than commercial P25 TiO2. After loading with Au nanoparticles, TiO2–Au spheres display a 27.6 μmol (g−1 h−1) H2 production rate under visible light irradiation (λ > 420 nm) because the localized surface plasmon resonance (LSPR) of Au nanoparticles enhances the visible light absorption. Furthermore, the H2 production rate could be improved to 92.4 μmol (g−1 h−1) for TiO2 spheres loaded with both Au and Pt nanoparticles. Based on these results, we propose a possible mechanism. Under UV light, TiO2 absorbs UV light and generates excited electrons, passing to Au nanoparticles for H2 production. In the case of visible light irradiation, the hot electrons are generated from Au nanoparticles due to the LSPR effect. And then the hot electrons are transferred from the Au nanoparticles to TiO2 and cocatalyst Pt nanoparticles for H2 generation.


Introduction

Titanium dioxide (titania, TiO2) is one of the most widely used semiconducting oxide materials and has a wide range of applications including bio-separation, sensors, energy storage, solar cells, catalysis and photocatalysis.1–7 TiO2 offers a number of advantageous characteristics including low cost, relatively high photocatalytic activity, low toxicity, and high chemical and optical stability. Since the discovery of hydrogen production from water through electrochemical photolysis,8 TiO2 has been intensively studied and used as a photocatalyst in both fundamental research and practical applications.9–11 Due to its large band gap (Eg = 3.0–3.2 eV), it can absorb UV light and generate electron (e) and hole (h+) pairs which can induce a variety of redox reactions. Over the past few decades, tremendous efforts have been devoted to improving the catalytic activity of TiO2-based photocatalysts with well controlled characteristics.12–14 However, weak absorption in visible region lies still on the way for the practical applications like solar water splitting. Generally, a few routes have been used for improving the absorption of TiO2 in visible region. Firstly, doping TiO2 with metal (like Fe3+, Cr3+, Ti3+ etc.) or non-metal (for example C, N, S and so on) will narrow the band gap.11,15–18 Secondly, sensitizing TiO2 with narrow band gap quantum dots or dye molecules, which absorb visible light and inject electron or hole into TiO2, improves the overall photocatalytic capability of TiO2.19–23

Another promising route to enhance the photocatalytic performance of TiO2 is to decorate the surface with noble metal nanoparticles (e.g. Ag, Au, Pt, and Pd).24–28 It has been shown that the presence of the noble metal nanoparticles can effectively shift the Fermi level of TiO2, which results in the enhanced photocatalytic efficiency.24,29 On the other hand, the localized surface plasmon resonance (LSPR) of noble metal nanoparticles also enhances the absorption in the visible region of TiO2. Kamat et al. investigated the size dependence of Au–TiO2 nanocomposites and found that small Au particles lead to large apparent Fermi level shift (20 mV for 8 nm diameter and 40 mV for 5 nm and 60 mV for 3 nm gold nanoparticles), which improves the photoinduced charge separation.24 In traditional way, adding the noble metal salt into TiO2 sol–gel precursor and photo-reducing the noble metal salts, it is hard to avoid the aggregation of gold nanoparticle to form big particles.30–32 Preparation of TiO2 hierarchical nanostructures loaded with Au nanoparticles are highly desired. Recently, Lou et al. reported the hierarchical TiO2 spheres composed in nanosheets with high surface area (∼100 m2 g−1) and high reactive (001) facet.33 On the other hand, this hierarchical TiO2 spheres have both accessible macropores from intersheets and mesopores from the packing of nanoparticles within nanosheets.

Herein, hierarchical TiO2 spheres composed in nanosheets were synthesized via modified Lou's route. Based on the hierarchical TiO2 spheres, we developed a facile method to load Au nanoparticles on the surface of TiO2 nanosheets by introducing an anchor molecule thioglycolique acid (TGA), which effectively controls the size of Au nanoparticles with ∼5 nm in diameter. The obtained TiO2–Au spheres possess the following advantages: (i) hierarchical structure shows high photocatalytic performance from large surface area; (ii) the LSPR of Au nanoparticles enhances the visible light absorption and improves the photocatalytic performance of TiO2 in the visible light region. H2 production experiments indicate that the TiO2–Au spheres show a high photocatalytic performance in both UV and visible region due to Au nanoparticles on the TiO2 surface. That makes TiO2 spheres decorated with Au nanoparticles as a high active visible-light photocatalyst.

Results and discussion

Firstly, the large scale of TiO2 spheres were prepared according to Lou's report with a modification.33 Fig. S1 shows field-emission scanning electron microscopy (FE-SEM) images of the obtained TiO2 spheres, which size can be tuned from 400–800 nm in diameter by reaction conditions. Close inspection images disclose that the TiO2 spheres are composed of TiO2 nanosheets, which is about 10 nm in thickness. High resolution TEM image displays that the nanosheets are composed of TiO2 nanoparticles with ∼6 nm of diameter after long time calcination. The crystalline phase of these TiO2 spheres can be transferred into anatase phase under calcination at 400 °C for 2–4 hours. The XRD patterns as shown in Fig. 3A, the calcined TiO2 spheres exhibit a typical anatase crystalline phase pattern (JCPDS no. 21-1272). To obtain Au nanoparticles on the TiO2 nanosheets surface, here we used an anchor molecule TGA, which has two functional groups, carboxyl group (–COOH) and thiol group (–SH). The COOH group makes the anchor molecules attach on the TiO2 surface. The thiol group on the other side will expose on the surface. Au complexes (Au (TGA)) forms between TGA and HAuCl4 with the addition of HAuCl4. Then Au nanoparticles form by adding into fresh NaBH4 solution. The anchor molecule TGA has a key role for the Au nanoparticles formation. It can effectively prevent from the aggregation of Au.

Fig. 1 show the FE-SEM images of TiO2 spheres decorated with Au nanoparticles. After loading Au nanoparticles, the morphology of TiO2 spheres is well kept. With the increase of HAuCl4 amount, the more HAuCl4 can be adsorbed on the TGA modified surface of TiO2 spheres. After reducing, Au nanoparticles are denser on the surface of TiO2 spheres. The energy dispersive X-ray analysis (EDAX, Fig. S2 and S3) proves the existence of Au. According to EDAX results, the weight fraction of loaded Au nanoparticles increases from 1.07, 2.00, 2.43 to 2.82 wt% by adding 5, 10, 20 and 30 mL of 1.0 mmol L−1 HAuCl4·3H2O solution into TiO2 sphere dispersion for samples TiO2–Au1, 2, 3, and 4, respectively. Fig. 2 display transmission electron microscopy (TEM) images of TiO2 spheres decorated with Au nanoparticles. Low magnification TEM image (Fig. 2A) clearly shows the TiO2 sphere with nanosheets structures. High resolution TEM images exhibit that 0.35 and 0.24 nm are characteristic lattice fringe space of anatase TiO2 (101) and Au (111) nanoparticles, respectively. That reveals the uniform Au nanoparticles with ∼5 nm in diameter are attached on the TiO2 nanosheets. In addition, the TiO2 nanosheets are composed of TiO2 nanoparticles with diameter of 6.01 ± 1.61 nm (Fig. 2C and D).


image file: c4ra15818k-f1.tif
Fig. 1 Field emission scanning electron microscopy (FE-SEM) images of TiO2 spheres decorated with different amount of Au nanoparticles. (A) 1.07 wt% (TiO2–Au1), (B) 2.00 wt% (TiO2–Au2), (C) 2.43 wt% (TiO2–Au3) and (D) 2.82 wt% (TiO2–Au4) measured from EDAX.

image file: c4ra15818k-f2.tif
Fig. 2 Transmission electron microscopy (TEM) images of TiO2 spheres loaded with 2.43 wt% Au nanoparticles (A and C). (B) is the high resolution TEM image of TiO2 loaded with Au nanoparticles. 0.35 nm and 0.24 nm are characteristic lattice fringe space of anatase TiO2 (101) and Au (111) nanoparticles. The histogram of TiO2 nanoparticles calculated from TEM images of (C).

Fig. 3A displays the X-ray diffraction patterns of TiO2 loaded with different amount of Au nanoparticles. It clearly shows the diffraction peaks at 25.3, 37.8, 48.0 degree, which are related to the (101), (004), (200) of anatase phase TiO2 (JCPDS no. 21-1272). According to the Scherrer equation, the TiO2 nanocrystal size was calculated to be 10 nm using (101) reflection peak. It is consistent with the diameter of TiO2 measured from TEM images. Diffused reflective UV-Vis spectroscopy is presented in the Fig. 3B. Compared with TiO2 spheres, a broad absorption band at ∼550 nm, related to the plasmon resonance of Au, is observed for the TiO2 spheres loaded with Au nanoparticles. The hierarchical TiO2 spheres exhibit high surface area. The Brunauer–Emmett–Teller (BET) surface area and pore size of the TiO2 spheres are characterized using nitrogen adsorption–desorption isotherm shown in Fig. 3C and D. It gives a type-IV isotherm with a type-H3 hysteresis loop, indicating a mesoporous structure.34 The BET surface area of TiO2 spheres with and without Au nanoparticles are determined to be 97 and 103 m2 g−1, respectively. According to previous reports,35 a bimodal mesopore size distribution was implied in this type of hysteresis loop. The hysteresis loop in the lower relative pressure range (0.4 < P/P0 < 0.8) is attributed to the mesopores from the packing of nanoparticles within nanosheets and that in the high relative pressure range (0.8 < P/P0 < 1.0) is related to the macropores from intersheets.


image file: c4ra15818k-f3.tif
Fig. 3 X-ray diffraction pattern (A) and diffusion reflection UV-Vis spectra (B) of TiO2 spheres loaded different amount of Au nanoparticles (TiO2–Au1–4 for Au weight fraction 1.07, 2.00, 2.43, 2.82 wt% respectively). (C and D) are the N2 sorption curves of TiO2 sphere without and with Au nanoparticles, respectively. Insets are the pore size distribution curves.

TiO2 spheres exhibit high photocatalytic activity compared with commercial P25 TiO2. Fig. S4A shows the H2 production of TiO2 spheres and P25 TiO2 loaded with 1 wt% Pt as cocatalyst under UV-Vis light (300 W Xe lamp). The H2 evolution rate is about 335 and 204 μmol hour−1 for 10 mg TiO2 sphere and P25 TiO2, respectively. The main reason is that TiO2 spheres have larger surface area than P25 TiO2 (∼50 m2 g−1). No H2 production is observed for 10 mg of TiO2 sphere under UV-visible light (Fig S4B). It exhibits obvious photocatalytic activity after loading Au nanoparticles. The H2 production rate is ∼23.9, 53.6, 165 and 87.3 μmol hour−1 for 10 mg sample of TiO2–Au1, 2, 3, and 4, respectively. The photocatalytic activity reaches maximum when the Au nanoparticles amount is about 2.43 wt% (TiO2–Au3). These results indicate that Au nanoparticles could work as a cocatalyst for H2 production. TiO2 adsorbs UV light and the electron was excited from valance band (VB) to conduction band (CB), then the photogenerated electron can inject into Au nanoparticles. The proton accepts the electron and forms H2.

Fig. 4 shows H2 production rate of TiO2 sphere with different amount Au nanoparticles under visible light (λ > 420 nm) in presence and absence of Pt cocatalyst. No H2 is produced from the pure TiO2 sphere without loading Au nanoparticles. That indicates that TiO2 spheres have no response to the visible light. Fig. S5 shows TiO2 spheres loaded both Au and Pt nanoparticles. When Au nanoparticles are attached onto the TiO2 sphere surface, H2 is detected. That means that Au nanoparticles absorb the visible light, generate hot electron due to the LSPR effect and transfer hot electron to TiO2 for H2 generation. The H2 evolution rate increases with the increasing of Au nanoparticles amount. However, too much amount of Au nanoparticles (2.84 wt%) results in a decrease of H2 production rate. The optimal loading amount of Au nanoparticles is about 2.43 wt% and H2 production rate is about 27 μmol (g−1 h−1). Pt is often chosen as a cocatalyst to lower the over potential of H2O splitting. Furthermore, when the cocatalyst Pt nanoparticles are loaded onto the TiO2–Au spheres via photo-reduction reaction, the H2 production rate has obvious improvement. The H2 production rate increases from 27 μmol (g−1 h−1) to 92 μmol (g−1 h−1), which is about 3 folds improvement.


image file: c4ra15818k-f4.tif
Fig. 4 (A) H2 production of TiO2 spheres loaded with different amount Au nanoparticles under visible light (λ > 420 nm). (B) Normalized H2 production rate of TiO2 sphere with Au nanoparticles. (C) H2 production of TiO2 spheres loaded with Au and Pt nanoparticles under visible light (λ > 420 nm). (D) Normalized H2 production rate of TiO2 spheres loaded Au and Pt nanoparticles.

Based on the above results, we propose the following mechanism for TiO2–Au composites (Scheme 1). Under UV light illumination, TiO2 adsorbs light and is excited to generate exciton. The electron is transferred to Au nanoparticles, which work as cocatalyst to promote H2 generation (Scheme 1A). It is known that a Schottky barrier is formed at the Au–TiO2 interface when Au nanoparticles make direct physical contact with TiO2.36 Upon excitation of the Au SPR with λ > 420 nm, intense SPR-enhanced EM field are generated on the surface of Au nanoparticles, increasing the yield of interfacial “hot electrons”. That will induce efficient transfer of “hot electrons” to the CB of TiO2. The Schottky barrier at the interface also helps the transferred “hot electrons” accumulating in the TiO2 CB, preventing them from traveling back to Au nanoparticles. Since no holes are generated in the valence band (VB) of TiO2 under λ > 435 nm excitation, the transferred “hot electrons” in the TiO2 CB should have much longer lifetimes, fostering the reduction of H2O to produce H2.37,38 Further, the reduction of H2O is promoted by the loading of cocatalyst Pt nanoparticle. The “hot electron” is transferred from Au nanoparticle to the CB of TiO2, then further to the Pt nanoparticles. Pt nanoparticles prefer to adsorb the proton and lower the over potential of H2O reduction.39 The hot electrons prefer transfer from Au nanoparticles to Pt nanoparticles where hydrogen evolution occurs. That is the reason why H2 production rate has an obvious improvement after loading Pt nanoparticles.


image file: c4ra15818k-s1.tif
Scheme 1 Possible photocatalytic H2 production mechanism of TiO2–Au under full spectrum light (A) and visible light (λ > 420 nm), where B is TiO2–Au and C is TiO2–Au and Pt nanoparticles.

Conclusions

In summary, we have loaded Au nanoparticles on the surface of TiO2 spheres through anchor molecules thiolglycolic acid. SEM and TEM disclosed that Au nanoparticles are well dispersed on the surface of TiO2 sphere. Due to TiO2 spheres are composed of nanosheets, it possesses relatively large surface area (∼100 m2 g−1), which is beneficial to the photocatalytic performance. TiO2 spheres exhibit high photocatalytic H2 production rate (165 μmol hour−1 for 10 mg of TiO2–Au3) under full spectrum light. The TiO2 spheres loaded with Au nanoparticles also shows good photocatalytic activity under visible light illumination due to the LSPR effect of Au nanoparticles. The photocatalytic performance could be further improved to 97 μmol hour−1 for 10 mg TiO2–Au3 by loading the cocatalyst Pt nanoparticle on the TiO2.

Experimental

Chemicals and materials

HAuCl4·3H2O (AR), NaBH4 (98%), thioglycolic acid (TGA, AR, 90%), diethylenetriamine (DETA, 99%), and titanium isopropylate (TIP, 98%) were purchased from Aladdin Reagent Company. Methanol (AR), isopropanol (AR) was purchased from Beijing Chemical Reagent Company. All chemicals were used without any further purification.

Synthesis of TiO2 sphere

At room temperature, the 380 mL of isopropanol was mixed with 300 L of DETA and stirred for 30 min. Then the 12.5 mL of TIP was injected with mixture slowly. The above solution was transferred into a 500 mL Teflon lined stainless steel autoclave and was heated to 200 °C for 48 h in an electric oven. Following this, the reaction was naturally cooled to room temperature. The as-prepared samples were washed with ethanol to remove remaining organic impurity and dried at 70 °C for 12 h in an electric oven. Finally, the samples were annealed at 400 °C for 4 h to form the crystal phase of TiO2.

Synthesis of TiO2–Au

At room temperature, the 80 mg of TiO2 sphere was dispersed into 10 mL of deionized water. 100 μL of TGA was added into the mixture and stirred for 6 h to make TGA adsorb on TiO2 sphere surface. Then the above mixture was centrifuged at 9000 rpm for 15 min to remove free TGA molecules. Then, 5, 10, 20 and 30 mL of 1.0 mmol L−1 HAuCl4·3H2O solution was added into TiO2 spheres dispersion and stirred for 6 h away from light to make gold ion bonded with TGA. Then the mixture was centrifuged at 9000 rpm for 15 min to remove free gold ion. 5 mL of 0.1 mol L−1 NaBH4 solution was injected rapidly and stirred at 1000 rpm. The as-prepared samples were centrifuged and washed with deionized water. Finally, the samples were dried at 70 °C for 12 h.

Characterizations

The Brunauer–Emmett–Teller (BET) specific surface area was measured using a Micromeritics Gemini V Surface Area and Pore Size Analyzer. Scanning electron microscope (SEM) images were measured on JEOL JSM 4800F. Transmission electron microscope (TEM) images were taken using an FEI Tecnai G2 operated at 200 kV. The UV-Vis absorption spectra were recorded on a UV-3600 UV-Vis-NIR scanning spectrophotometer (Shimadzu). The crystalline structure was recorded by using an X-ray diffractometer (XRD) (Bruker AXS D8 Focus), using Cu Kα radiation (λ = 1.54056 Å).

Photocatalytic activity measurements

10 mg of TiO2–Au photocatalyst was placed into an aqueous methanol solution (120 mL, 25 vol%) in a closed gas circulation system (Perfect Light Company Labsolar-III (AG)). The UV-visible light irradiations were obtained from a 300 W Xe lamp (Perfect Light Company Solaredge 700) without and with a UVCUT-420 nm filter (CE Aulight. Inc). The evolved gases were detected in situ by using an online gas chromatograph. (GC-2014C, Shimadzu) equipped with a thermal conductivity detector (TCD). To obtain TiO2–Au–Pt photocatalyst, 10 mg of TiO2–Au photocatalyst loaded with 1.0 wt% Pt (H2PtCl6) was placed into an aqueous methanol solution in a closed gas circulation system. Before collecting the evolution gas, the reaction was irradiated with UV-Vis light over an hour for reduction Pt nanoparticles onto the TiO2–Au samples.

Acknowledgements

The authors thank the National Natural Science Foundation of China (no. 21301166, 21201159, and 61361166004), Science and Technology Department of Jilin Province (no. 20130522127JH, and 20121801) are gratefully acknowledged. Z. S. thanks the support of the “Hundred Talent Program” of CAS. Supported by open research fund program of State Key Laboratory of Luminescence and Applications (CIOMP, CAS) and Key Laboratory of Functional Inorganic Material Chemistry (Heilongjiang University), Ministry of Education, P. R. China.

Notes and references

  1. X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891–2959 CrossRef CAS PubMed.
  2. M. Graetzel, R. A. J. Janssen, D. B. Mitzi and E. H. Sargent, Nature, 2012, 488, 304–312 CrossRef CAS PubMed.
  3. J. Du, J. Qi, D. Wang and Z. Tang, Energy Environ. Sci., 2012, 5, 6914–6918 CAS.
  4. K. Nakata and A. Fujishima, J. Photochem. Photobiol., C, 2012, 13, 169–189 CrossRef CAS PubMed.
  5. Q. Zhang, E. Uchaker, S. L. Candelaria and G. Cao, Chem. Soc. Rev., 2013, 42, 3127–3171 RSC.
  6. T. Froschl, U. Hormann, P. Kubiak, G. Kucerova, M. Pfanzelt, C. K. Weiss, R. J. Behm, N. Husing, U. Kaiser, K. Landfester and M. Wohlfahrt-Mehrens, Chem. Soc. Rev., 2012, 41, 5313–5360 RSC.
  7. H. G. Moon, Y.-S. Shim, H. W. Jang, J.-S. Kim, K. J. Choi, C.-Y. Kang, J.-W. Choi, H.-H. Park and S.-J. Yoon, Sens. Actuators, B, 2010, 149, 116–121 CrossRef CAS PubMed.
  8. A. Fujishima and K. Honda, Nature, 1972, 238, 37–38 CrossRef CAS.
  9. X. Chen, S. Shen, L. Guo and S. S. Mao, Chem. Rev., 2010, 110, 6503–6570 CrossRef CAS PubMed.
  10. X. Chen, L. Liu, P. Y. Yu and S. S. Mao, Science, 2011, 331, 746–750 CrossRef CAS PubMed.
  11. F. Zuo, K. Bozhilov, R. J. Dillon, L. Wang, P. Smith, X. Zhao, C. Bardeen and P. Feng, Angew. Chem., Int. Ed., 2012, 51, 6223–6226 CrossRef CAS PubMed.
  12. J. Lee, M. Christopher Orilall, S. C. Warren, M. Kamperman, F. J. DiSalvo and U. Wiesner, Nat. Mater., 2008, 7, 222–228 CrossRef CAS PubMed.
  13. M. Yin, Y. Cheng, M. Liu, J. S. Gutmann and K. Müllen, Angew. Chem., Int. Ed., 2008, 47, 8400–8403 CrossRef CAS PubMed.
  14. H. G. Yang, C. H. Sun, S. Z. Qiao, J. Zou, G. Liu, S. C. Smith, H. M. Cheng and G. Q. Lu, Nature, 2008, 453, 638–641 CrossRef CAS PubMed.
  15. W. Choi, A. Termin and M. R. Hoffmann, J. Phys. Chem., 1994, 98, 13669–13679 CrossRef.
  16. E. Borgarello, J. Kiwi, M. Graetzel, E. Pelizzetti and M. Visca, J. Am. Chem. Soc., 1982, 104, 2996–3002 CrossRef CAS.
  17. R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, Science, 2001, 293, 269–271 CrossRef CAS PubMed.
  18. D. Pan, J. Zhang, Z. Li, C. Wu, X. Yan and M. Wu, Chem. Commun., 2010, 46, 3681–3683 RSC.
  19. D. R. Baker and P. V. Kamat, Adv. Funct. Mater., 2009, 19, 805–811 CrossRef CAS.
  20. H. Park, W. Choi and M. R. Hoffmann, J. Mater. Chem., 2008, 18, 2379–2385 RSC.
  21. C. Karakaya, Y. Türker and Ö. Dag, Adv. Funct. Mater., 2013, 23, 4002–4010 CrossRef CAS.
  22. E. Bae and W. Choi, Environ. Sci. Technol., 2002, 37, 147–152 CrossRef.
  23. Sustainable Nanotechnology and the Environment: Advances and Achievements, ed. N. Shamim and K. Sharma Virender, American Chemical Society, 2013 Search PubMed.
  24. V. Subramanian, E. E. Wolf and P. V. Kamat, J. Am. Chem. Soc., 2004, 126, 4943–4950 CrossRef CAS PubMed.
  25. Y. Lai, Y. Tang, J. Gong, D. Gong, L. Chi, C. Lin and Z. Chen, J. Mater. Chem., 2012, 22, 7420–7426 RSC.
  26. R. Su, R. Tiruvalam, Q. He, N. Dimitratos, L. Kesavan, C. Hammond, J. A. Lopez-Sanchez, R. Bechstein, C. J. Kiely, G. J. Hutchings and F. Besenbacher, ACS Nano, 2012, 6, 6284–6292 CrossRef CAS PubMed.
  27. H. Li, Z. Bian, J. Zhu, Y. Huo, H. Li and Y. Lu, J. Am. Chem. Soc., 2007, 129, 4538–4539 CrossRef CAS PubMed.
  28. C. Zhou, L. Shang, H. Yu, T. Bian, L.-Z. Wu, C.-H. Tung and T. Zhang, Catal. Today, 2014, 225, 158–163 CrossRef CAS PubMed.
  29. K. Mogyorósi, Á. Kmetykó, N. Czirbus, G. Veréb, P. Sipos and A. Dombi, React. Kinet. Catal. Lett., 2009, 98, 215–225 CrossRef.
  30. M. Epifani, C. Giannini, L. Tapfer and L. Vasanelli, J. Am. Ceram. Soc., 2000, 83, 2385–2393 CrossRef CAS PubMed.
  31. S. C. Chan and M. A. Barteau, Langmuir, 2005, 21, 5588–5595 CrossRef CAS PubMed.
  32. X. Z. Li and F. B. Li, Environ. Sci. Technol., 2001, 35, 2381–2387 CrossRef CAS.
  33. J. S. Chen, Y. L. Tan, C. M. Li, Y. L. Cheah, D. Luan, S. Madhavi, F. Y. C. Boey, L. A. Archer and X. W. Lou, J. Am. Chem. Soc., 2010, 132, 6124–6130 CrossRef CAS PubMed.
  34. M. Kruk and M. Jaroniec, Chem. Mater., 2001, 13, 3169–3183 CrossRef CAS.
  35. K. Kaneko, J. Membr. Sci., 1994, 96, 59–89 CrossRef CAS.
  36. Y. K. Lee, C. H. Jung, J. Park, H. Seo, G. A. Somorjai and J. Y. Park, Nano Lett., 2011, 11, 4251–4255 CrossRef CAS PubMed.
  37. J. S. DuChene, B. C. Sweeny, A. C. Johnston-Peck, D. Su, E. A. Stach and W. D. Wei, Angew. Chem., Int. Ed., 2014, 53, 7887–7891 CrossRef CAS PubMed.
  38. K. Qian, B. C. Sweeny, A. C. Johnston-Peck, W. Niu, J. O. Graham, J. S. DuChene, J. Qiu, Y.-C. Wang, M. H. Engelhard, D. Su, E. A. Stach and W. D. Wei, J. Am. Chem. Soc., 2014, 136, 9842–9845 CrossRef CAS PubMed.
  39. D. Duonghong, E. Borgarello and M. Graetzel, J. Am. Chem. Soc., 1981, 103, 4685–4690 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: More SEM, EDAX, H2 production of TiO2 spheres. See DOI: 10.1039/c4ra15818k

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