Supported silver nanoparticles as photocatalysts

0 downloads 0 Views 255KB Size Report
formaldehyde and methanol in air at room temperature due to. aSchool of ... silver NPs were carried out on Philips CM200 TEM with ... An air conditioner.
View Online

PAPER

www.rsc.org/greenchem | Green Chemistry

Supported silver nanoparticles as photocatalysts under ultraviolet and visible light irradiation Xi Chen,a Zhanfeng Zheng,a Xuebin Ke,a Esa Jaatinen,a Tengfeng Xie,b Dejun Wang,b Cheng Guo,c Jincai Zhaod and Huaiyong Zhu*a

Downloaded by Institute of Chemistry, CAS on 15 February 2011 Published on 26 January 2010 on http://pubs.rsc.org | doi:10.1039/B921696K

Received 16th October 2009, Accepted 2nd December 2009 First published as an Advance Article on the web 26th January 2010 DOI: 10.1039/b921696k The significant activity for dye degradation by silver nanoparticles (NPs) on oxide supports was better than popular semiconductor photocatalysts. Moreover, silver photocatalysts can degrade phenol and drive oxidation of benzyl alcohol to benzaldehyde under ultraviolet light. We suggest that surface plasmon resonance (SPR) effect and interband transition of silver NPs can activate organic molecules for oxidation under ultraviolet and visible light irradiation.

Introduction Photocatalysts show great potential as drivers of chemical reactions when illuminated by sunlight at ambient temperatures.1-3 One of the great challenges in this field is devising new catalysts that consist of nanoparticles (NPs) usually below 10 nm size regime and possess high activity when illuminated with either visible light or ultraviolet (UV) light.1 The new photocatalysts will enable us to use sunlight, the abundant and green energy source, to drive useful chemical reactions. Once sunlight is utilized as a substitute of fossil fuel to drive reactions for production of important chemicals and environmental remediation, this will alleviate our reliance on fossil fuel energy and reduce energy consumption and CO2 emissions. Conventional semiconductor photocatalysts, in particular TiO2 based materials have been extensively investigated.2 As these photocatalysts have a large band gap, photocatalysis can only occur when UV light is absorbed. Since more than 43% of the solar energy is in the visible part of the spectrum,3 many approaches have been proposed to develop photocatalysts that can perform under visible light.1,3-4 It is well known that gold, silver and copper NPs strongly absorb visible-light due to the socalled surface plasmon resonance (SPR) effect.5 The SPR effect is the collective oscillation of conduction electrons in the NPs, which resonate with the electromagnetic field of the incident light. Also these excited electrons will return to their thermal equilibrium states and release heat to the lattice and surrounding medium.6 This heating effect may also induce reactions of the molecules adsorbed on the particles. Indeed, we found that when illuminated with visible light, gold NPs dispersed on oxide supports exhibited significant activity for oxidation of formaldehyde and methanol in air at room temperature due to a School of Physical and Chemical Sciences, Queensland University of Technology, Queensland, 4001, Australia. E-mail: [email protected]; Fax: +61 731381804; Tel: +61 731381581 b College of Chemistry, Jilin University, Changchun, 130012, China c College of Science, Nanjing University of Technology, Nanjing, 210009, China d Institute of Chemistry, Chinese Academy of Science, Beijing, 100080, China

414 | Green Chem., 2010, 12, 414–419

the SPR effect.7 Since visible light absorption heats the electrons and excites them from ground state to higher energy levels, the probability that a conduction electron participates in chemical reactions involving electron transfer is greater. Silver NPs also exhibit considerable UV light absorption due to the interband transition (the transition of 4d electrons to the 5sp band).8 Therefore, silver NPs are potentially photocatalysts that utilize the full solar spectrum. Silver NPs on the surface of semiconductors and electron-donor substances cause charge separation of photogenerated electron-hole pairs, thus enhancing the overall photocatalytic activity.9 However, the photocatalytic activity of the silver NPs themselves has not been recognized. While silver ions were reported to be photoactive for certain reactions, such as nitric oxide decomposition and carbon-hydrogen bond activation,10 to date even the precise nature of the reaction mechanism for the catalysis involving plasmonic silver materials has not been clarified. Here we verify that silver NPs at room temperature can be used to drive chemical reactions when illuminated with light throughout the solar spectrum, and in the process gain some understanding into the mechanism behind the photocatalytic process (which is different from that for the conventional semiconductor photocatalysts).

Experimental Silver NPs preparation Solution-phase reduction methods11 were used to prepare the silver NPs supported on different oxides. ZrO2 , Zeolite Y and amorphous SiO2 powders were chosen as supports because of their band gaps (above ~5.0 eV), which are much larger than the energies of the photons of visible light (below 3.0 eV). Hence, the light cannot excite electrons from the valence band to the conduction band of the support. It is also impossible for the silver NPs on the support to reduce its band gap enough for visible light photons to be absorbed. Thus, the observed visible light absorption and catalytic activity by the photocatalysts is due to the supported silver NPs. The AgNO3 solution (3 ¥ 10-3 M) containing suspended 0.5 g ZrO2 , SiO2 or zeolite Y This journal is © The Royal Society of Chemistry 2010

View Online

(surface areas are 34, 47, 653 m2 g-1 , respectively) was irradiated with six UV lamps (20 W/tube, from Philips). The irradiated mixture was then centrifuged for 2 h, and the obtained Ag/oxide precipitate was washed with deionized water, dried at 80 ◦ C and heated at 450 ◦ C for 6 h.

Downloaded by Institute of Chemistry, CAS on 15 February 2011 Published on 26 January 2010 on http://pubs.rsc.org | doi:10.1039/B921696K

Sample characterization Transmission electron microscopy (TEM) studies of supported silver NPs were carried out on Philips CM200 TEM with an accelerating voltage of 200 kV. The silver content was determined by energy dispersive X-ray spectroscopy (EDS) using FEI Quanta 200 Environmental SEM. X-ray photoelectron spectroscopy (XPS) analysis was performed on a Kratos Analytical Axis Ultra X-ray photoelectron spectrometer. X-ray diffraction (XRD) was carried out using a PANalytical with CuKa radiation. The surface photocurrent (SPC) measurements were performed on the system constituted of a source of monochromatic light, a lock-in amplifier (SR830-DSP) with a light chopper (SR540), and a photovoltaic cell. A 500 W xenon lamp (CHFXQ500W, Global xenon lamp power) and a double-prism monochromator (Hilger and Watts, D300) provide monochromatic light. A comb-like ITO electrode with an external bias (10.0 V) on its two sides was used. The sample chamber for transient photovoltage (TPV) measurements consists of an ITO electrode, a 10 mm thick mica spacer as electron isolator, and a platinum wire gauze electrode (with a transparency of about 50%). The construction is a sandwich-like structure of ITO electrode-sample-mica-gauze platinum electrode. During the measurement, the gauze platinum electrode was connected to the core of a BNC cable which input signals to the oscilloscope. The samples were excited from platinum wire gauze electrode with a laser radiation pulse (wavelength of 532 nm and pulse width of 5 ns) from a third-harmonic Nd:YAG laser (Polaris II, New Wave Research, Inc.). The intensity of the pulse was regulated with a neutral gray filter and determined with an EM500 singlechannel joulemeter (Molectron, Inc.). The TPV signals were registered with a 500 MHz digital phosphor oscilloscope (TDS 5054, Tektronix).

TiO2 at 550 ◦ C in N2 gas for 4 h to test SRB photodegradation. In the photocatalytic phenol degradation experiment, aqueous suspensions of organic compounds (100 mL, 1 mM) and 100 mg of silver photocatalyst powders were placed in the vessel. Then the vessel was in a chamber with 6 UV light tubes as the light source (20 W/tube, NEC, light intensity was 0.014 W cm-2 , wavelength around 365 nm). The filtrates were analysed at 270 nm in the UV-vis spectra using the Varian 50. In the benzyl alcohol degradation experiment, 50 ml toluene suspensions of benzyl alcohol (10%) and 50 mg of silver photocatalyst powders were placed in the glass vessel. Then the vessel was in a chamber with 6 UV light tubes as the light source. In order to increase photocatalytic activity, 30 mg NaOH was added into the benzyl alcohol solution and the vessel was filled with pure oxygen as the reaction atmosphere. The filtrates were analysed in GC HP6890 Prometheus to measure the concentration change of benzyl alcohol.

Results and discussion In this study, we loaded silver NPs onto various typical oxide supports12 and used these photocatalysts for degrading a range of organic compounds in aqueous solution under either visible light or UV irradiation at room temperature. TEM images of silver NPs supported on ZrO2 (Ag/ZrO2 ), amorphous SiO2 (Ag/SiO2 ) and Zeolite Y (Ag/Zeolite Y) are shown in Fig. 1. These images indicate that silver exists in these samples as NPs. Most of the silver particles (the dark-colour substance) on these supports were found to have dimensions below 10 nm, which can lead to changes in surface and electronic structure providing an opportunity to control catalytic activity and selectivity.13 The silver contents in Ag/Zeolite Y, Ag/ZrO2 and Ag/SiO2 samples were found by EDS to be 7.39, 7.48 and 7.56 wt% of the overall photocatalyst mass, respectively. XPS analysis shown in Fig. 2 indicated that the silver exists in metal state. However, no silver peaks can be identified through XRD pattern of our photocatalysts (shown in Fig. 3), probably due that the loaded silver did not form large particles, but was dispersed in the support structure.

Photocatalytic tests In the photocatalytic dye sulforhodamine-B (SRB) degradation experiment under air atmosphere, aqueous suspensions of SRB (50 mL, 2 ¥ 10-5 M) and 50 mg of photocatalyst were placed in a glass vessel, which was in a chamber with 6 light tubes (18 W/tube, Philips, light intensity 0.011 W cm-2 , wavelength around 450 nm) as the light source. The pH of the solutions was adjusted to 2.5 with 0.1 M HNO3 . An air conditioner was installed in the chamber to maintain the temperature at 25 ◦ C as the light illumination could cause an increase in temperature of the vessel. Before irradiation the suspensions were magnetically stirred in the dark for 30 min to establish adsorption/desorption equilibrium between the dye and the catalyst. At given irradiation time intervals, 4 mL aliquots were collected, centrifuged, and then filtered through a Millipore filter (pore size 0.45 mm) to remove the catalyst particulates. The filtrates were analyzed at the wavelength of maximal absorption (565 nm) in the UV-vis spectra of SRB using a Varian 5000. For comparison we also product nitrogen-doped TiO2 by annealing This journal is © The Royal Society of Chemistry 2010

Fig. 1 Transmission electron microscopy images of the photocatalysts of silver NPs on supports. (a) Silver on zeolite Y, Ag-zeolite Y. (b) Silver on zirconia, Ag-ZrO2 . (c) Silver on silica, Ag-SiO2 .

Dyes are of special interest as their use in the textile and industrial industries is becoming a significant source of environmental contamination.14 Under visible light irradiation silver NPs dispersed on oxide supports exhibited significant activity for SRB degradation at 25 ◦ C (An air conditioner was installed to maintain the temperature as the light illumination could cause an increase in the vessel temperature), which is even better than can be achieved with the widely reported nitrogen Green Chem., 2010, 12, 414–419 | 415

View Online

Downloaded by Institute of Chemistry, CAS on 15 February 2011 Published on 26 January 2010 on http://pubs.rsc.org | doi:10.1039/B921696K

Fig. 2 Binding energy of Ag 3d5/2 and Ag 3d3/2 for silver photocatalysts. (1) Ag/Zeolite Y; (2) Ag/ZrO2 ; (3) Ag/SiO2 .

Fig. 3 XRD analysis of silver photocatalysts. (1) Ag/Zeolite Y; (2) Ag/ZrO2 ; (3) Ag/SiO2 .

doped TiO2 photocatalysts. The comparison of degradation curves of SRB in aqueous solutions using silver photocatalysts after 180 min of blue light irradiation are shown in Fig. 4. Silver supported on zeolite exhibited the highest degradation ability among these materials. SRB content decreased by 74% in 3 h under blue light irradiation with intensity of 0.011 W cm-2 . The photocatalysts of silver supported on zirconia and silica (Ag/ZrO2 and Ag/SiO2 ) are also very active for SRB degradation (shown in Fig. 4). After 3 h of blue light irradiation, 71% and 66% of the dye were degraded with Ag/SiO2 and Ag/ZrO2 , respectively, slightly lower than the degradation rate with Ag/Zeolite Y. We also found a trend that the ability of the photocatalysts to degrade SRB increased with the increasing silver content. The Ag/Zeolite Y samples containing 4.5 wt% and 1.5 wt% of silver were also prepared. The SRB content decreased to 56% and 40% under blue light irradiation in 3 h, respectively. In the dark, SRB was not decomposed with any one of the three silver photocatalysts. A blank experiment under the otherwise identical conditions but without silver NPs (solutions with these oxide supports - Zeolite Y, ZrO2 or SiO2 powder only) was also conducted, and no SRB conversion above 3% was observed. Moreover, silver photocatalysts were stable under repeated application. About 82% SRB conversion catalyzed by

Fig. 4 Degradation curves of SRB under blue light using different photocatalysts. (1) N-doped TiO2 . (2) Ag/Zeolite Y. (3) Ag/SiO2 . (4) Ag/ZrO2 .

416 | Green Chem., 2010, 12, 414–419

Ag/Zeolite Y can be maintained within 5 photodegradation recycles. We also studied the effect that the intensity of the light irradiation had on the SRB degradation reaction. No concentration changes were detected if the experiment was conducted without light irradiation. When the light intensity was reduced to 0.008 and 0.006 W cm-2 (by turning off 2 or 4 blue light tubes, respectively), the SRB conversion by Ag/SiO2 decreased from 71% to 50% and 27%, respectively. The wavelength of the irradiation also affects the photocatalytic activity. Under red light irradiation (of 6 light tubes from Philips, with overall light intensity 0.010 W cm-2 and wavelength around 650 nm) about 56% and 39% of SRB was converted by using Ag/SiO2 and Ag/ZrO2 as catalysts in 3 h, which is substantially lower than that under blue light (71% and 66%). These observations show that the SRB degradation is undoubtedly driven by visible light. Next, as the calcination temperature of silver samples was 300 ◦ C, only 49% SRB was degraded over Ag/Zeolite Y under blue light for 3 h due to weak interaction between silver and support. While high calcination temperature (600 ◦ C) can lead to big silver paricles. 53% SRB was converted under the same experimental condition except the calcination temperature. Moreover, these silver photocatalysts also can degrade organic compounds under UV light irradiation at room temperature. The UV light absorption by silver NPs can excite interband transition and be utilized to drive photoreactions. TiO2 is the most widely studied photocatalyst under UV light irradiation, and nitrogen-doped TiO2 can exhibit high activity when illuminated by visible light.4 We compare the SRB photodegradation activity of the supported silver particles with that of TiO2 material in Table 1. SRB conversions from replicate runs agree to each other within ±3%. The results confirm that silver NPs supported on oxides are superior to TiO2 based photocatalysts for SRB degradation under both blue and UV light irradiation, given that the silver particles are the active photocatalysis component and that silver accounts for about 7.5 wt% of the catalyst mass. The silver photocatalysts exhibited better catalytic activity under UV irradiation than under blue light. They not only decompose dye molecules under UV light faster than under visible light, but also are able to oxidize phenol in aqueous solution, which they cannot catalyze under visible light. After 120 h of UV light irradiation 41% of phenol was degraded by Ag/ZrO2 . The photocatalytic conversion of phenol by Ag/Zeolite and Au/SiO2 was 37% and 38%, respectively. The conversion of blank experiment (without catalyst under UV irradiation) was below 1%. These experimental results indicate that silver particles supported on oxides can catalyze the degradation of organic compounds without involving photosensitization process like dyes in aqueous solution at ambient temperature.15 The photocatalytic decomposition of organic compounds discussed above involves transfer of multiple electrons from the organic molecules to oxygen—the oxidant. In principle, selective (or partial) oxidation of an organic compound can be achieved with these photocatalysts if we regulate the electron transfer process by tuning the experimental conditions. Indeed, supported silver NPs were found to be effective catalysts for oxidation of benzyl alcohol in toluene to benzaldehyde. 11% benzyl alcohol conversion was achieved in 48 h under UV light irradiation with 62% of the product being benzaldehyde, when This journal is © The Royal Society of Chemistry 2010

View Online

Table 1 Absorption of irradiation energy and catalytic activity of the photocatalysts to degrade SRB Under blue light a

Under UV light b

Conversion

Absorb energy

TOF

Photocatalyst

%

W cm-2

atom-1 s-1

Ag-zeolite Y Ag-ZrO2 Ag-SiO2 N-doped TiO2 TiO2 (P25)

68 48 50 34 14

0.007 0.007 0.004 0.001 0.001

1.6 ¥ 10-5 1.1 ¥ 10-5 1.2 ¥ 10-5 1.5 ¥ 10-7 8.8 ¥ 10-9

Conversiona

Absorb energy

TOFb

Normalized TOFb

cm2 J-1 (Ag or TiO2 )-atom-1 s-1

%

W cm-2

atom-1 s-1

cm2 J-1 (Ag or TiO2 )-atom-1 s-1

2.3 ¥ 10-3 1.6 ¥ 10-3 3.0 ¥ 10-3 1.5 ¥ 10-4 8.8 ¥ 10-6

75 55 64 49 75

0.007 0.006 0.004 0.005 0.006

1.8 ¥ 10-5 1.3 ¥ 10-5 1.5 ¥ 10-5 2.2 ¥ 10-7 3.3 ¥ 10-7

2.6 ¥ 10-3 2.2 ¥ 10-3 3.8 ¥ 10-3 4.4 ¥ 10-5 5.5 ¥ 10-5

Normalized TOF

b

SRB conversions from replicate runs agree to within ±3%. b Turnover frequency data in the table were calculated from the conversion after 1 h of irradiation.

Downloaded by Institute of Chemistry, CAS on 15 February 2011 Published on 26 January 2010 on http://pubs.rsc.org | doi:10.1039/B921696K

a

Ag/zeolite Y was used as the photocatalyst. Partial oxidation for producing aldehydes from corresponding alcohols is a very important process for the fine chemical industry.16 We also increased the selectivity for producing benzaldehyde to 100% by adjusting the pH with NaOH solution17 and using pure oxygen instead of air, though the overall conversion fell to 4% in 48 h. Based on these facts, we conclude that the photocatalytic process with supported silver NPs does not occur via the same mechanism as found for semiconductor photocatalysts, such as TiO2 .2 In our study the silver NPs were supported on zeolite, ZrO2 and SiO2 . ZrO2 has a band gap of about 5 eV.7 The band gaps of zeolite Y and SiO2 are slightly larger than that for ZrO2 , according to the UV-vis absorption measurements shown in Fig. 5a by Cary5000 UV-Vis spectrometer. These supports alone (in the absence of the NPs) exhibit little light absorption. When light wavelengths are above 330 nm, the illumination cannot excite electrons of the supports from the valence band to the conduction band. Thus, all the photogenerated charges that lead to catalytic activity originate from the silver NPs. The silver in the photocatalysts remains in a metal state as indicated by the XPS analysis (Fig. 2), which is dissimilar to AgCl which can donate electrons and exhibit photocatalytic activity by the oxidation of Cl- ions to Cl0 atoms under light irradiation.9 The photocatalytic reactions (degradation and selective oxidation) involve electron transfer from the molecules of the oxidized reactant to those of the reduced reactant. We believe that the silver NPs initialize and mediate the electron transfer for the photooxidation reactions. It has been reported that the

silver doped on TiO2 surface can interact strongly with the oxygen atoms and give rise to an electron transfer to the Ti 3d states.18 In order to determine whether light irradiation can induce electron transfer from the silver particle to the oxygen molecules (or oxygen adsorbed on the support), SPC and TPV spectra of the samples were also analysed and shown in Fig. 5. A surface current and a transient photovoltage will arise whenever excess light-induced charge carriers are separated in space, with the signal intensity being proportional to the number of the photogenerated charges.19 These spectra also explicitly exhibit the dependence of the electron transfer on the illumination wavelength allowing the spectral regions under which electron transfer occurs to be identified. For typical isolated spherical silver NPs the SPR absorption is generally around 380 nm, but the absorption band of the supported NPs is significantly red-shifted to above 410 nm.20 Aggregation of the silver NPs and non-spherical shaped particles, which we observed in the samples in the present study, broaden the absorption resonance at low light intensity.21 The silver NPs exhibit UV absorption, due to the interband excitation of electrons from 4d to 5sp. From the SPC spectra (Fig. 5b) it is apparent that UV absorption produces a much larger surface photocurrent than that induced by the SPR absorption under visible light irradiation. Also, a large initial photovoltage is observed in the TPV spectrum (Fig. 5c) which was measured with a 532 nm laser that has a much higher intensity than the light used for the SPC measurement. This indicates that visible light irradiation does generate electrical

Fig. 5 (a) Light absorption of silver photocatalysts, zeolite Y and oxides in UV and visible light range. (b) Surface photocurrent spectra of silver NPs supported on zeolite Y (trace in blue) and zeolite Y (trace in black). (c) Transient photovoltage spectra of silver NPs supported on zeolite Y (trace in blue) and zeolite Y (trace in black).

This journal is © The Royal Society of Chemistry 2010

Green Chem., 2010, 12, 414–419 | 417

Downloaded by Institute of Chemistry, CAS on 15 February 2011 Published on 26 January 2010 on http://pubs.rsc.org | doi:10.1039/B921696K

View Online

surface charges when sufficiently intense. The SPC spectrum also indicates that the interband absorption (UV) results in a much larger proportion of electron transfer from silver NPs to the oxygen molecules than the SPR absorption (visible). Consequently, more positive charges (holes) are left in the silver NPs under UV illumination. To explain these photocatalytic observations, we propose a tentative transfer mechanism as illustrated in Scheme 1. Before light illumination the silver electron occupancy obeys the Fermi– Dirac distribution. Blue light irradiation will excite the SPR and is strongly absorbed. Silver electrons are excited from within the outermost sp band to higher energy states.5 Soon after light absorption the plasmon loses energy causing rapid heating of electron gas to an elevated temperature (about 400–2000 K) within a time scale of the order of 100 fs or less through electron–electron collision.22 Then the electrons share the heat energy from the ‘hot’ electron gas with the NP lattice through electron–phonon collisions. The time scale varies from 500 fs to 10 ps.23 Therefore, it is possible that the electrons with enough energy may be captured by a body once the capture process takes within 1 ps, according to Furube’s finding that the electron transfer from gold NPs to titanium oxide takes less than 240 fs.24 We suggested that under visible light irradiation at moderate intensity, a very small number of silver electrons with high temperature gain sufficient energy (above the green line in Scheme 1) to be captured by oxygen molecules. However, most of the electrons are excited to lower energy levels (below

the green line in Scheme 1), which cannot be captured. Thus, a weak SPC signal and slow dye degradation are observed under moderate visible light illumination. The holes left in the 5sp band can capture electrons from excited SRB molecules (SRB*), due that the photosensitization process under light irradiation which involves initial excitation of the dye molecules can be helpful for injecting dye electrons.15 When silver particles absorb UV irradiation, electrons of the 4d band can be excited to the 5sp tates band,8 and many photogenerated electrons are in high energy (the green line or above in Scheme 1) where oxygen molecules can seize them, yielding large surface photovoltage signals under UV light. The holes left in the inner d band have a greater affinity for capturing electrons from the adsorbed organic molecules than those in the outermost sp band. Thus the presence of holes in the d band allows any attached phenol to be degraded by accepting its electrons. Since different mechanisms are responsible for degrading organic compounds under visible and UV light, the photocatalysts ability to oxidize specific compounds will depend on the illumination wavelength. This property can be utilized for two reaction regimes. First, under UV illumination, the photocatalysts can oxidize the compounds that they cannot oxidize under visible light, such as phenol. However, further research can be done with other reduction substances whose reduction potential is lower than that of O2 /O2 - . Probably this will provide visible light driven catalytic activity of phenol degradation to silver photocatalysts. Second, the illumination wavelength can be used as a control parameter to determine whether a specific reaction will take place or not. The ability for the photocatalysts to capture electrons under UV light is a useful feature that can be used for producing desired chemicals under the experimental conditions which can prevent further oxidation, such as selective oxidation of benzyl alcohol to benzaldehyde by silver photocatalysts.

Conclusions In summary, silver NPs are good photocatalysts under ambient temperature for degrading organic compounds. Silver NPs supported on oxides can exhibit significant oxidation activity for a synthetic dye under visible light illumination. These photocatalysts can also catalyze phenol degradation as well as selective oxidation of benzyl alcohol under UV light. The findings indicate conceptually that it is possible to drive various chemical reactions with visible light. Therefore, as in the case of photochemistry driven by surface plasmon, it is a distinct possibility that environmental remediation and fine chemical production can be performed using the most efficient light source available—visible light. This will alleviate our reliance on fossil fuel energy and concerns in regards to global warming. Our findings also show the potential to switch on or off specific reactions by tuning the light wavelength. This development will lead to a new direction in photocatalysis research.

Acknowledgements Scheme 1 The diagram of the band structures of the supported silver NPs and the proposed photocatalysis mechanism.

418 | Green Chem., 2010, 12, 414–419

Financial Supports from the Australian Research Council (ARC), and 973 program (2007CB613306) and NSFC (20537010) of China are gratefully acknowledged. This journal is © The Royal Society of Chemistry 2010

View Online

Downloaded by Institute of Chemistry, CAS on 15 February 2011 Published on 26 January 2010 on http://pubs.rsc.org | doi:10.1039/B921696K

Notes and references 1 N. S. Lewis, Nature, 2001, 414, 589–590; T. L. Thompson and J. T. Yates, Chem. Rev., 2006, 106, 4428–4453; S. U. M. Khan and M. Al-Shahry, Science, 2002, 297, 2243–2245; M. Zhang, C. C. Chen, W. H. Ma and J. C. Zhao, Angew. Chem., Int. Ed., 2008, 47, 9730– 9733. 2 M. I. Litter, Appl. Catal., B, 1999, 23, 89–114; A. L. Linsebigler, G. Q. Lu and J. T. Yates, Chem. Rev., 1995, 95, 735–758; X. B. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891–2959. 3 Z. G. Zou, J. H. Ye, K. Sayama and H. Arakawa, Nature, 2001, 414, 625–627. 4 R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, Science, 2001, 293, 269–271. 5 H. Yuan, Chem. Mater., 2007, 19, 1592–1600; G. C. Schatz, Acc. Chem. Res., 1984, 17, 370–376; X. H. Huang, I. H. El-Sayed, W. Qian and M. A. El-Sayed, J. Am. Chem. Soc., 2006, 128, 2115–2120; X. L. Zheng, W. Q. Xu, C. Corredor, S. P. Xu, J. An, B. Zhao and J. R. Lombardi, J. Phys. Chem. C, 2007, 111, 14962–14967; S. Eustis and M. A. El-Sayed, Chem. Soc. Rev., 2006, 35, 209–217; K. L. Kelly, E. Coronado, L. L. Zhao and G. C. Schatz, J. Phys. Chem. B, 2003, 107, 668–677; P. Mulvaney, Langmuir, 1996, 12, 788–800; L. M. LizMarz´an, Langmuir, 2006, 22, 32. 6 S. Link, C. Burda, Z. L. Wang and M. A. El-Sayed, J. Chem. Phys., 1999, 111, 1255–1264; K. Yamada, K. Miyajima and F. Mafune, J. Phys. Chem. C, 2007, 111, 11246–11251. 7 X. Chen, H. Y. Zhu, J. C. Zhao, Z. F. Zheng and X. P. Gao, Angew. Chem., Int. Ed., 2008, 47, 5353–5357. 8 P. V. Kamat, J. Phys. Chem. B, 2002, 106, 7729–7744; P. P. Kiran, B. N. Bhaktha, D. N. Rao and G. De, J. Appl. Phys., 2004, 96, 6717– 6723; C. Voisin, N. Del Fatti, D. Christofilos and F. Vallee, J. Phys. Chem. B, 2001, 105, 2264–2280. 9 Y. H. Zheng, C. Q. Chen, Y. Y. Zhan and X. Y. Lin, J. Phys. Chem. C, 2008, 112, 10773–10777; P. Wang, B. B. Huang, X. Y. Qin, X. Y. Zhang, Y. Dai, J. Wei and M. Whangbo, Angew. Chem., Int. Ed., 2008, 47, 7931–7933.

This journal is © The Royal Society of Chemistry 2010

10 M. Matsuoka and M. Anpo, Curr. Opin. Solid State Mater. Sci., 2003, 7, 451–459; G. A. Ozin and F. Hugues, J. Phys. Chem., 1982, 86, 5174–5179. 11 R. Hu, L. Y. Xie, S. Ding, J. Hou, Y. Cheng and D. Z. Wang, Catal. Today, 2008, 131, 513–519; Y. Shi, H. Pan, Y. T. Zhang and W. Li, Catal. Commun., 2008, 9, 796–800; F. C. Meunier, J. P. Breen, V. Zuzaniuk, M. Olsson and J. R. H. Ross, J. Catal., 1999, 187, 493– 505. 12 M. M. Schubert, S. Hackenberg, C. Andre, M. Muhler and V. Plzak, J. Catal., 2001, 197, 113–122; L. Tosheva and V. P. Valtchev, Chem. Mater., 2005, 17, 2494–2513. 13 A. T. Bell, Science, 2003, 299, 1688–1691. 14 E. J. Weber and R. L. Adams, Environ. Sci. Technol., 1995, 29, 1163– 1170. 15 J. C. Zhao, C. C. Chen and W. H. Ma, Top. Catal., 2005, 35, 269–278. 16 V. R. Choudhary, P. A. Chaudhari and V. S. Narkhede, Catal. Commun., 2003, 4, 171–175. 17 J. J. Zhu, J. L. Figueiredo and J. L. Faria, Catal. Commun., 2008, 9, 2395–2397; S. K. Klitgaard, A. T. DeLa Riva, S. Helveg, R. M. Werchmeister and C. H. Christensen, Catal. Lett., 2008, 126, 213– 217. 18 L. Giordano, G. Pacchioni, T. Bredow and J. F. Sanz, Surf. Sci., 2001, 471, 21–31. 19 P. Wang, B. B. Huang, X. Y. Zhang, X. Y. Qin, H. Jin, Y. Dai, Z. Y. Wang, J. Wei, J. Zhan, S. Y. Wang, J. P. Wang and M. Whangbo, Chem.–Eur. J., 2009, 15, 1821–1824. 20 T. Hirakawa and P. V. Kamat, J. Am. Chem. Soc., 2005, 127, 3928– 3934. 21 J. A. Creighton and D. G. Eadon, J. Chem. Soc., Faraday Trans., 1991, 87, 3881–3891. 22 A. Takami, H. Kurita and S. Koda, J. Phys. Chem. B, 1999, 103, 1226–1232; D. K. Roper, W. Ahn and M. Hoepfner, J. Phys. Chem. C, 2007, 111, 3636–3641. 23 S. Link and M. A. El-Sayed, Int. Rev. Phys. Chem., 2000, 19, 409–453. 24 A. Furube, L. C. Du, K. Hara, R. Katoh and M. Tachiya, J. Am. Chem. Soc., 2007, 129, 14852–14853.

Green Chem., 2010, 12, 414–419 | 419