Au decorated Fe3O4@TiO2 magnetic composites with visible light-assisted enhanced catalytic reduction of 4-nitrophenol

Ying Zhou, Yihua Zhu*, Xiaoling Yang, Jianfei Huang, Wei Chen, Xiaoming Lv, Cuiyan Li and Chunzhong Li*
Key Laboratory for Ultrafine Materials of Ministry of Education, School of Materials Science and Engineering, East China University of Science and Technology, Shanghai 200237, China. E-mail: yhzhu@ecust.edu.cn; czli@ecust.edu.cn

Received 5th May 2015 , Accepted 26th May 2015

First published on 26th May 2015


Abstract

Heterostructured Au nanoparticle decorated Fe3O4@TiO2 composite magnetic microspheres (MSs) were synthesized by grafting Au nanoparticles onto 3-aminopropyltrimethoxysilane (APTMS) modified Fe3O4@TiO2 MSs. Significantly, by varying the reaction conditions, the as-synthesized Fe3O4@TiO2@Au MSs showed high performance in the catalytic reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP) in the presence of NaBH4 under visible light. In addition, the as-prepared Fe3O4@TiO2@Au MSs can clean themselves by photocatalytic degradation of organic molecules, and can be reused for several cycles with convenient magnetic separability. This approach provided a platform based on the synergy of varying components under suitable conditions to optimize the catalytic ability.


1. Introduction

In heterogeneous catalysis, reactants are absorbed on the surface of a catalytically active solid. Most of such heterogeneous catalysts are typically composed of small particles of catalytically active materials, which can significantly enhance the contact between catalysts and surroundings, thus ensuring the effectiveness of catalysts.1 Au is usually chemically inert, however, the electronic properties of nanoparticles differ from the corresponding bulk materials, and nanosized Au can be effective in catalysis.2 Gold nanoparticles (Au NPs) have been found to feature in several chemical reactions including low-temperature CO oxidation,3 reductive catalysis of chlorinated or nitrogenated hydrocarbons,4–6 and organic synthesis.7,8 However, owing to the high surface energy, Au NPs are easy to aggregate leading to remarkable reduction even deactivation in their original catalytic activity.9 Hence, Au NPs are generally immobilized onto a variety of supports including polymer,10 metal oxides,11–13 graphene oxide,14 etc. Moreover, the synergistic interactions between metal nanoparticles and metal oxide supports contribute to the catalytic activity.15 The composite way to produce new material could realize the combination or cooperation of the characteristics of each component leading to enhanced properties. In most cases, the recycle of such catalysts is tedious and time-consuming by centrifugation/redispersion thus hinders the effective recovery and reuse.12

Fe3O4 nanoparticles have been extensively applied in many fields due to the rapid and efficient magnetic response by the aid of an external magnetic field. Their insoluble and superparamagnetic natures enable the separation of catalysts from reaction mixture trouble-free and efficient.12 Recently, Fe3O4 nanoparticles have drawn attention as robust, readily available, high-surface-area supports in catalytic transformations.16,17 However, Fe3O4 nanoparticles as supports are vulnerable to the air. Thus, it is a need to build core/shell-structured composite carriers to hold Fe3O4 nanoclusters as core and functionalized layer as shell. The shells need to be chemically stable, convenient for the immobilization of Au NPs and beneficial for the catalytic reactions. On the basis of such consideration, TiO2 can be an appropriate material, which possess high specific surface areas beneficial for the absorption of molecules, large pore volumes, as well as stable and interconnected frameworks with active pore surfaces for easy modification and functionalization.13

As reported by Kuroda et al., Au NPs directly deposited on poly(methyl methacrylate) (PMMA) beads, which showed a higher catalytic activity than Au NPs.10 Esumi et al.18 investigated the catalytic activity of dendrimer-stabilized nanoparticles. These authors demonstrated that the catalytic reaction was the diffusion controlled. So the exposed Au NPs in the reaction mixture is essential to achieve excellent catalytic efficiency. Up to now, there have been many efforts towards assembling noble-metal nanoparticles onto various supports with different structures. For example, Zhang et al.19 have reported the synthesis of Au nanostructures on graphene oxide and their catalytic activity, but the rate constant of catalysis is moderate. What's more, it remains a difficult task to separate the catalysts from reaction solution. Chang et al.20 have reported the structure of Ag decorated polyaniline nanofibers with Ag NPs exposed outside, the easy oxidation of Ag NPs may cause a reduction of catalytic activity and be negative to reuse. Zheng et al.21 have synthesized Fe3O4@SiO2@Au and such structure exhibited high catalytic activity. Efforts have been made to synthesize Au NPs decorated composite catalysts, while the response of noble metal to visible light is less reported. The localized surface plasmon resonance (LSPR) of Au NPs can bring several benefits to the catalysis when the system is exposed to visible light: heating effect, strengthened dipore–dipore reaction with polar molecule, as well as the response of TiO2 to visible light which contribute to the synergistic effect of noble catalysis and photocatalysis.22

Nitrophenols are among the common pollutants in the industrial and agricultural wastewater.16 The removal of them can be realized by processes such as adsorption,23 photocatalytic degradation,24 microwave-assisted catalytic oxidation,25 electro-Fenton method,26 electrochemical treatment,27 and so on. On the other hand, 4-aminophenol (4-AP) is of great importance in the manufacture of analgesic and antipyretic drugs.28 4-AP is also widely used as photographic developer, corrosion inhibitor, anticorrosion-lubricant, and hair-dyeing agent.29 The reduction of 4-nitrophenol (4-NP) over Au nanocatalysts in the presence of NaBH4 offers an effective way to the production of 4-AP. Also, the reduction has no byproducts formed and is easily monitored via UV-vis spectroscopy, which is suitable for the investigation of catalytic activity.

Herein, we report the synthesis of stable, reproducible and recyclable Fe3O4@TiO2@Au composites. The as-prepared multifunctional microspheres (MSs) have Fe3O4 core inside, TiO2 interlayer and Au shell outside with uniform size range and areal distribution. These catalysts with the multicomponent nanostructure, designated Fe3O4@TiO2@Au, are fabricated by uniform coating Fe3O4 with TiO2 through sol–gel process and then with the modification of Fe3O4@TiO2 MSs by 3-aminopropyltrimethoxysilane (APTMS), followed by coating with a layer of citrate-stabilized Au NPs. The Au decorated Fe3O4@TiO2 magnetic composites with visible light-assisted enhanced catalytic reduction of 4-NP were investigated. The catalytic activities of these composites have been tested by using the reduction of 4-NP to 4-AP as model reaction and monitored by UV-vis spectroscopy. The unique nanostructure makes the composite MSs stable, recyclable and high-enhancement in catalysis.

2. Experimental

2.1. Reagents and materials

FeCl3·6H2O, chloroauric acid, trisodium citrate, citric acid, sodium acetate, tetrabutyl titanate (TBOT), ethanol, ethylene glycol, NH4F, aqueous ammonia (28 wt%) were of analytical grade and purchased from Shanghai Chemical Corp. 3-Aminopropyltrimethoxysilane (APTMS) were purchased from Sigma-Aldrich Chemicals Co. All chemicals were used as received. Ultrapure water (18 MΩ cm−1) was used for all experiments.

2.2. Synthesis of Fe3O4 MSs

Fe3O4 MSs were synthesized by the solvent-thermal method reported by Liu et al.30 In a typical synthesis, 3.25 g of FeCl3·6H2O, 1.3 g of trisodium citrate and 6.0 g of sodium acetate were added into 100 mL of ethylene glycol under magnetic stirring. The mixture was then transferred and sealed into a Teflon-lined stainless-steel autoclave (200 mL in capacity). Then the reaction was allowed to proceed at 200 °C for 10 h. The obtained black products were collected using an external magnetic field, rinsed with deionized water and ethanol for 3 times, respectively. Finally, the obtained Fe3O4 MSs were dried in vacuum for 12 h.

2.3. Synthesis of Fe3O4@TiO2 MSs

The Fe3O4 MSs with a uniform porous TiO2 shell outside were performed following the sol–gel process described by Zhao group.31 Typically, 0.075 g of Fe3O4, 0.3 mL of aqueous ammonia (28 wt%) were dissolved in 90 mL of ethanol and stirred for 15 min at 600 rpm at 45 °C. Then under continuous mechanical stirring, 0.75 mL of TBOT previously dissolved in 10 mL of ethanol was added dropwise into the mixture. The reaction was allowed to proceed at 45 °C for 24 h. The resultant products were separated and collected using an external magnetic field, rinsed with ethanol and water and then dried in vacuum for 12 h. To improve crystallinity, 0.4 g of Fe3O4@TiO2 MSs and 0.37 g of NH4F were dispersed in a mixture of 66 mL of ethanol and 34 mL H2O under mechanical stirring for 60 min. The mixture was hydrothermally treated at 180 °C in a Teflon-lined stainless steel autoclave (200 mL in capacity) for 24 h. The as-prepared products were washed with ultrapure water for 3 times to remove any ionic impurities, and dried at 80 °C overnight.

2.4. Synthesis of Fe3O4@TiO2@Au MSs

The as-prepared Fe3O4@TiO2 MSs were dissolved in 200 mL of ethanol and 0.15 mL of APTMS, then the mixture was refluxed at 85 °C for 4 h. The products were washed with ethanol for 3 times and redispersed in 100 mL of ethanol for further use. Thus the APTMS modified Fe3O4@TiO2 MSs were obtained. Citrate-stabilized Au NPs with an average diameter of ∼10 nm were prepared according to the methods reported.32,33 The pH of the Au NPs and modified Fe3O4@TiO2 solution were adjusted to 5.0 by titration with 0.01 M HCl. Then 2.0 mL of modified Fe3O4@TiO2 solution was added dropwise to 200 mL of the raw Au NPs solution and stirring for 15 min to obtain the proper density of Au NPs on the surface of the Fe3O4@TiO2 MSs.

2.5. Characterization

The morphologies and sizes of the as-prepared samples were characterized by transmission electron microscopy (TEM) using a JEOL 2011 microscope (Japan) and scanning electron microscope (SEM) images using a JEOL JSM-6360LV microscope (Japan). The crystalline structure was investigated by X-ray power diffraction (RIGAKU, D/MAX 2550 VB/PC, Japan). The magnetic behavior was investigated using a vibrating sample magnetometer (VSM, Lake Shore 7304, Lake Shore, USA) with an applied field between −15 kOe and 15 kOe at room temperature. The UV-vis spectra were recorded on a UV-vis spectrometer (TU-1901, Beijing, China).

2.6. Catalytic reactions with Fe3O4@TiO2@Au MSs

The catalytic performance of the multifunctional Fe3O4@TiO2@Au MSs were investigated by the reduction of 4-NP to 4-AP in the presence of NaBH4 as a model reaction. In a typical run, 5.85 mL of ultrapure water, 0.15 mL 5 mM of 4-NP solution, and 3.0 mL 20 mM of freshly prepared NaBH4 aqueous solution were added into a 25 mL quartz beaker respectively, and the solution color turned to bright yellow rapidly. The pH of solution during the whole experimental process was controlled to 8.0 by titration with 0.01 M NaOH. Subsequently, 1.0 mL 5 wt% Fe3O4@TiO2@Au MSs was added to start the reaction. Agitation was provided by mechanical stirring. A UV lamp (365 nm, 70 W cm−2) or a high pressure Xenon lamp with a cutoff filter to block the UV light (>400 nm, 50 mW cm−2) was used as the light source, which was placed 10 cm away from the reaction vessel. After irradiation, the sample was collected with a magnet and the supernatant was extracted for the examination, and the intensity of absorption peak at 400 nm was monitored by UV-vis spectroscopy as a function of time. As the reaction was proceeding, the gradual change of the solution color from yellow to colorless can be observed. Fig. 1 shows the schematic diagrams for (a) the preparation of the multifunctional Fe3O4@TiO2@Au MS and (b) the catalytic reduction of 4-NP with the multifunctional Fe3O4@TiO2@Au MSs in the presence of NaBH4 exposed to visible light.
image file: c5ra08243a-f1.tif
Fig. 1 Schematic diagrams for (a) the preparation of the multifunctional Fe3O4@TiO2@Au MS and (b) the catalytic reduction of 4-NP with the multifunctional Fe3O4@TiO2@Au MSs in the presence of NaBH4 exposed to visible light.

3. Results and discussion

3.1. Characterization of the triplex Fe3O4@TiO2@Au core–shell MSs

The triplex Fe3O4@TiO2@Au core–shell MS was composed of a Fe3O4 magnetic core, a TiO2 interlayer, and a layer of Au NPs shell outside. The well-defined Fe3O4 MSs with a narrow size distribution and a mean diameter of ∼400 nm were first synthesized through a robust solvothermal reaction (Fig. 2a), the magnetic core ensure the easy separate of the catalysts from the reaction mixture. Close observation revealed that the uniform gray interlayer of TiO2 with a thickness of about 35 nm provide a continuous coverage of the magnetic Fe3O4 MS (Fig. 2b), which not only are beneficial to the absorption of reactants but also ensure the degradation of organic products so that the MSs can be reused. Then, the resultant Fe3O4@TiO2 core–shell MSs are modified with APTMS to render the particle surfaces with amino groups. Finally, to mix the modified core–shell Fe3O4@TiO2 MSs with citrate-stabilized Au NPs leading to the graft of Au NPs onto the surfaces of Fe3O4@TiO2 MSs (Fig. 2c and d). The Au NPs with a mean diameter of 10 nm were well-dispersed on the surface of Fe3O4@TiO2 core–shell MS, and no aggregation was observed.
image file: c5ra08243a-f2.tif
Fig. 2 (a) SEM image of Fe3O4 MS. (b) TEM image of Fe3O4@TiO2 core–shell MSs with a uniform TiO2 shell thickness of 35 nm. (c) SEM image of Fe3O4@TiO2@Au (10 nm) and (d) the corresponding magnified TEM image.

Fig. 3 shows X-ray diffraction (XRD) patterns of (blue) Fe3O4 MSs, (red) Fe3O4@TiO2 MSs and (black) Fe3O4@TiO2@Au MSs during different stages. The characteristic broad diffraction peaks shown in wide-angle XRD patterns can be indexed to the spinel Fe3O4, anatase TiO2 and cubic phase Au NPs in the composite MSs. The specific XRD of Fe3O4 is characterized by six peaks positioned at 2θ values of 30.0°, 35.3°, 42.9°, 53.5°, 57.0° and 62.4°, corresponding to the [220], [311], [400], [422], [511] and [440] lattice planes of the cubic phase of Fe3O4 (JCPDS card no. 01-075-0449), respectively. As shown in Fig. 3(red), after the coating of a ∼35 nm TiO2 layer and subsequent hydrothermal treatment, by comparison with the patterns of Fe3O4, that of the Fe3O4@TiO2 show several additional peaks, which can be attributed to the anatase phase of TiO2. The reflections from the [101], [004], [200] and [105] planes of the anatase phase (JCPDS card no. 01-075-2545) contribute the extra peaks located at 25.3°, 37.9°, 48.0° and 53.9°. It can be clearly shown in Fig. 3(black) that after loading 10 nm Au NPs on the surface of the Fe3O4@TiO2 MS, the additional diffraction peaks can be indexed to the cubic phase of Au (JCPDS card no. 00-004-0784). The characteristic XRD patterns of Fe3O4@TiO2@Au MSs imply that the current system is suitable for the synthesis.


image file: c5ra08243a-f3.tif
Fig. 3 XRD patterns of (blue) Fe3O4, (red) Fe3O4@TiO2 and (black) Fe3O4@TiO2@Au MSs.

Fig. 4 shows the magnetic hysteresis loops of the bare Fe3O4, Fe3O4@TiO2, and Fe3O4@TiO2@Au MSs at room temperature. As shown, they exhibited superparamagnetic behavior and little hysteresis, remanence and coercivity due to the fact that the particles were composed of ultrafine magnetite nanocrystals. The magnetic saturation (Ms) value of Fe3O4 MSs is about 53.2 emu g−1, and that of Fe3O4@TiO2 and Fe3O4@TiO2@Au are 41.1 and 22.2 emu g−1, respectively. By comparison, decrease in magnetization is mainly attributed to the decrease in the density of Fe3O4 in the obtained composites after coating with TiO2 and then Au. However, it should be noted that the Fe3O4@TiO2@Au still showed strong magnetization, which suggested their suitability for magnetic separation. With the use of external magnetic field, the materials were quickly attracted to one side of the vial with several seconds and the solution become transparent, as illustrated in the physical photograph (Fig. 4 inset).


image file: c5ra08243a-f4.tif
Fig. 4 Magnetization curves of the Fe3O4, Fe3O4@TiO2 and Fe3O4@TiO2@Au MSs, respectively. The inset pattern is a photograph of the magnetic separation.

3.2. Application of the Fe3O4@TiO2@Au MSs for the catalytic reduction of 4-NP

4-NP is one of the most common organic pollutants in industrial and agricultural wastewater that is difficult to degrade,16 while 4-AP is a kind of useful raw material in medicine and the chemical industry. To investigate the catalytic activity of Fe3O4@TiO2@Au MSs, the catalytic reduction of 4-NP to their corresponding daughter derivatives 4-AP in the presence of NaBH4 was chosen as the model reaction. The reduction reaction is easy to follow for the reason that only one kind of product (4-AP) is obtained and the system can be monitored by UV-vis absorbance at 400 nm. This process can be monitored by the UV-vis spectrum for the reason that 4-NP and 4-AP have different absorption wavelengths (400 nm and 300 nm, respectively) in the UV-vis range.34 Fig. 5 shows the pure 4-NP solution exhibited a distinct spectral profile with an maximal absorbance at 317 nm, the absorbance peaks shift to 400 nm was observed immediately as soon as the NaBH4 solution was added, which correspond to a color change from light yellow to bright yellow for the formation of 4-nitro-phenolate ion in alkaline condition,35,36 as illustrated in the physical photograph (Fig. 5 inset). And there is no change in the absorbance at 400 nm over time, as shown in Fig. 6, which confirmed that the reaction cannot be pushed forward with NaBH4 solution only.
image file: c5ra08243a-f5.tif
Fig. 5 UV-vis absorption spectra of 4-NP (red line) with and (black line) without NaBH4 solution.

image file: c5ra08243a-f6.tif
Fig. 6 Ct/C0 versus reaction time for the reduction of 4-NP with NaBH4 (black line), Fe3O4@TiO2 MSs (red line), Fe3O4@TiO2–NH2 MNPs (blue line) and Fe3O4@TiO2@Au MSs (green line), respectively.

When evaluating the catalytic activity of Fe3O4@TiO2@Au MSs, that of Fe3O4@TiO2 and APTMS modified Fe3O4@TiO2 was investigated in advance. As shown in Fig. 6, there is no change in the UV-vis absorbance at 400 nm meaning that neither Fe3O4@TiO2 nor APTMS modified Fe3O4@TiO2 has catalytic activity to the reduction of 4-NP. The time-dependent UV-vis spectra of the reaction running in the presence of Fe3O4@TiO2@Au MS as catalysts as well as being exposed to light was unambiguously exhibited in Fig. 7a. The absorption intensity of 4-NP at 400 nm significantly decrease the elapsed time when the catalysts were introduced accompanied with the appearance of absorption peak at 300 nm. As the reaction went on, the absorption intensity at 400 nm decreased along with the increase of that at 300 nm. The full reduction of 4-NP was completed within 3 min by the observation of a fading and ultimate bleaching of the bright yellow colour of the reaction mixture. The synergistic effect of each component could be explored by the contrast of catalytic reduction run with Fe3O4@TiO2@Au MS, TiO2@Au composite spheres, Fe3O4 MS and Au NPs subjected to different conditions including the exposure to visible light or not as shown in Fig. 7. As shown in Fig. 7b, the full reduction could be completed within 6 minutes in dark. When the system is in dark, the effect of photocatalysis was suppressed and noble metal catalysis play a dominant role. When the system was exposed to visible light but accompanied by the addition of TiO2@Au composites spheres as catalyst, as shown in Fig. 7d, the reaction is completed within 10 min with full reduction of 4-NP, and by comparison, the full reduction was completed within 13 min in dark (Fig. 7e). When the system was performed with Fe3O4 MS as catalysts, small reduction of 4-NP was observed exposed to visible light (Fig. 7g) and no obvious reduction of 4-NP was noticed (Fig. 7h) even more than 180 min. When Au NPs (10 nm) alone served as catalysts, 4-NP has hardly any reduction with illumination or not (Fig. 7j and k). As the initial concentration of NaBH4 is far higher than reactants, it is considered to remain constant throughout the whole reaction process. So the reaction is assumed to be independent of the concentration of NaBH4, and a pseudo first-order kinetic equation could be applied to evaluate the catalytic rate.37–39 In addition, most of the photocatalytic reactions follow the Langmuir–Hinshelwood adsorption model,40 and the L–H model can also be simplified to a pseudo-first-order expression. The absorption intensity of 4-NP is proportional to the concentration in the medium. Hence, the kinetic equation for the reduction could be written as follows:

image file: c5ra08243a-t1.tif


image file: c5ra08243a-f7.tif
Fig. 7 Catalytic reduction of 4-NP conducted in different conditions: Fe3O4@TiO2@Au MS with illumination (a), in dark (b); TiO2@Au with illumination (d), in dark (e); Fe3O4 MS with illumination (g), in dark (h); Au NPs with illumination (j), in dark (k), ln(Ct/C0) versus the reaction time for the reduction of 4-NP over Fe3O4@TiO2@Au MS (c), TiO2@Au (f), Fe3O4 MS (i) and Au NPs (l) with illumination (black) and in dark (red) at 25 °C. C0 and Ct is the absorption peak at 400 nm initially and at time t.

Ct is the concentration of 4-NP at time t and k is the apparent rate constant. Linear relationships between ln(Ct/C0) and the reaction time are displayed in Fig. 7c, f, i and l, which matched well with the first-order reaction kinetics. The rate constant k (Fig. 7cblack) was calculated to be 1.180 min−1 at 25 °C exposed to visible light, which was higher than that reported for the same catalytic conversion using catalysts based on Au or Ag nanoparticles with varying morphologies and sizes.14,41,42 And the values of k are 0.421 min−1 (Fig. 7cred) in dark, 0.452 min−1 with illumination and 0.247 min−1 in dark (Fig. 7f), 0.00515 min−1 with illumination and 6.08 × 10−4 min−1 in dark (Fig. 7i), 0.00153 min−1 with illumination and 0.00123 min−1 in dark (Fig. 7l) respectively.

Fig. 8 depicts the performance of the Fe3O4@TiO2@Au MSs as catalysts in the reduction of 4-NP to 4-AP by NaBH4 exposed to visible light. One of prominent features of noble metal is the LSPR. The LSPR refers to the oscillation of the metal's free electrons in phase with the varying electric field of the incident light.43,44 When the system is irradiated by visible light, the electron density decreases on one side of the particle and increases on the other, facilitating the adsorption of BH4 onto Au NPs. The reduction action occurred via electrons transfer from the donor BH4 to the acceptor 4-NP after both adsorbed onto the particles surface. The hydrogen atom from the hydride, attacked 4-NP molecules to reduce it to 4-AP after the electron transfer to Au NPs.45 The photocatalysis occurred simultaneously accompanied with the generation of degradation products.


image file: c5ra08243a-f8.tif
Fig. 8 Schematic and photographic representation of the performance of Fe3O4@TiO2@Au MSs as catalysts in the reduction of 4-NP by NaBH4 exposed to visible light.

In general, heterogeneous catalysis proceed following the five independent steps:46,47 (1) transfer of the reactant to the reaction surface, (2) absorption of reactant onto the active sites, (3) redox reactions, (4) desorption of the products and (5) transfer of products away from the surface. Any process beneficial to the five steps above may contribute to the catalytic performance enhancement. The uses of noble metal nanoparticles (Au NPs) benefit the catalytic reaction from not only the catalytic effects of noble metal itself but also the distinct feature LSPR. This will bring some benefits to the catalysis. LSPR can heat up the local environment leading to the increase of the mass transfer of the reactants or products and then the enhancement of the reaction rates. The heating effects come from the collision electrons with the ionic lattice of the metal nanoparticles within 0.7 ps,48 and the phonon–phonon relaxation coupled with the metal lattice and the semiconductor lattice.49 Typically, the heating effects work efficiently in small nanoparticles (below 30 nm).50–52 When the noble nanoparticles were in the state of LSPR, the absorption of reactant molecules onto the surface will be enhanced by dipole–dipole interactions. Since the local electric filed caused by LSPR is a fast-varying dipole and 4-NP are polar, 4-NP may be attracted to a metal surface by such dipole–dipole reaction. Metal nanoparticles can also provide a fast lane for charge transfer and the surface can be served as a charge-trap center to host more active sites for the interaction with molecules absorbed on it.53 In the view of the five steps, the LSPR, the charge trapping and fast transfer and the large of contact areas contribute to the reaction rate, thus benefiting the step 3;47,54 the heating effects can not only enhance the reaction rate but also accelerate the solution mixing and benefits step 1, 3, 4 and 5;55 the enhancement of reactant absorption can benefit step 2. These contributions lead to the enhancement of the catalytic rates. In addition, the effects of semiconductor part cannot be ignored. TiO2 is one kind of high-performance photocatalytic materials but confined to the large bandgaps and only absorb the light in the near ultraviolet (UV) region (wavelength < 400 nm). The loading of Au NPs onto TiO2 can make it response to the visible-light and perform its photocatalytic activity. The synergy of noble metal-catalysis and photocatalysis contribute to the high rate constant of the reduction of 4-NP (Fig. 7a, d, cblack and fblack). When the reaction system is in dark, no photocatalysis from semiconductor performed, leaving only the noble-metal catalysis (Fig. 7b, e, cred and fred). To verify the synergy of each components, the experiment catalyzed by TiO2@Au, Fe3O4 MS and Au NPs (10 nm) was performed. And the results were shown in Fig. 7d–l. The catalytic activity of Au NPs is strongly dependent on the size, and only catalysts with Au particles below 5 nm show catalytic activity.1 In the article, the size of Au NPs loaded on the spot is around 10 nm. So, as Au NPs synthesized serve as catalysts alone for the reduction of 4-NP, it showed little catalytic activity (as shown in Fig. 7j and k), meanwhile the aggregation of Au NPs without any support may also lead to reduction of activity. When the Au NPs are loaded onto TiO2 for the formation of TiO2@Au. The composites exhibit catalytic activity. It can be explained as follows: (1) the synergy of two components, (2) the loading of Au NPs on TiO2 not only avoid the aggregation of Au NPs but also provide more active sites. The mesoporous structure of TiO2 is beneficial for the adsorption of reactant molecules, that is to say, it can increase the contact chances for reactants and catalysts which is better than the random collision for Au NPs alone. The reduction proceeded under the condition of light-on better than that of light-off can be attributed to the assistance of visible light which contribute to the photocatalytic effect of TiO2 (as shown in Fig. 7d–f). The catalytic activity of Fe3O4 is very low and the location of it inside also limited the contact with reactants, the activity cannot be compared with other catalysts (Fig. 7g–i). Fe3O4 NPs can absorb visible light. The Fe3O4 core of the composites not only exhibits the magnetic property for the easy separation of catalysts from the reaction solution but also enhance the adsorption of visible light which can promote the catalytic process (Fig. 7a and d).

3.3. Recovery ability of the Fe3O4@TiO2@Au MS

One of the main advantages of as-prepared Fe3O4@TiO2@Au MSs is the potential recyclability. By the external magnetic field, it is convenient for Fe3O4@TiO2@Au MSs to separate from the reaction medium within several minutes. In addition, if the obtained Fe3O4@TiO2@Au MSs can self-clean the molecules absorbed on it, the reuse ability can be realized.56 The self-cleaning activity was evaluated by monitoring the residue characteristic absorption band at 300 nm to measure the degradation rate of 4-AP under a UV lamp (365 nm, 70 W cm−2) or a high pressure Xenon lamp with a cutoff filter to block the UV light (<400 nm, 50 W cm−2). It has been shown recently that metal NPs, such as Au and Ag, embedded in the matrix of TiO2 can enhance the photocatalytic activity of TiO2 under visible light irradiation.57–59 Fig. 9 a show the concentration change of the aqueous 4-AP solution versus irradiation time. As can be seen, the Fe3O4@TiO2@Au MS exhibits the excellent photocatalytic performance. Most of the photocatalytic reactions follow the Langmuir–Hinshelwood adsorption model,46 and the L–H model can be simplified to a pseudo-first-order expression: ln(C0/C) = kt (where C0 and C are the initial concentration and the concentration of 4-AP at the exposure time, t, respectively, and k is the linear plots of ln(C0/C) versus irradiation time t is attained) and the photocatalytic reaction rate of Fe3O4@TiO2@Au is 0.096 min−1 as shown in Fig. 9b. Thus, the Fe3O4@TiO2@Au MS can achieve the self-cleaning property under visible light irradiation. And then the MS was washed with deionized water and collected by magnet. Then the MS were dried for the catalytic reduction again. As shown in Fig. 9c, MS can be recycled and reused for at least 8 times with a stable conversion of ∼92%. The decrease of conversion after 8 cycles may be attributed to the loss of catalysts with the repeated magnetic separation.
image file: c5ra08243a-f9.tif
Fig. 9 Photocatalytic degradation study of 4-AP on Fe3O4@TiO2@Au MS (a) and ln(C0/C) versus time line study on Fe3O4@TiO2@Au under visible light irradiation (b). The recovery ability of Fe3O4@TiO2@Au MS as a catalyst for the reduction of 4-NP with the presence of NaBH4 (c).

4. Conclusions

In conclusion, we have successfully synthesized the novel multifunctional Fe3O4@TiO2@Au core–shell composites. The Fe3O4@TiO2@Au MSs exhibited excellent catalytic activity for the reduction of 4-NP in the presence of NaBH4 under visible light with convenient magnetic separability and fine recoverability. The synergy of noble metal-catalysis and photocatalysis contribute to the high rate constant of the reduction of 4-NP (k = 1.180 min−1 at 25 °C). Especially the Fe3O4@TiO2@Au MSs can be recycled and reused for at least 8 times with a stable conversion of ∼92%. This approach provides a useful platform based on the synergy of varying components under suitable conditions to optimize the catalytic ability. Additionally, these core–shell magnetic microspheres can be extended to various advanced applications such as environmental protection, chemical biosensors, nanoelectronics and so on.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (21471056, 21236003, 21206042, and 21176083), the Basic Research Program of Shanghai (13NM1400700, 13NM1400701), and the Fundamental Research Funds for the Central Universities.

Notes and references

  1. B. Hvolbæk, T. V. W. Janssens, B. S. Clausen, H. Falsig, C. H. Christensen and J. K. Nǿrskov, Nano Today, 2007, 2, 14 CrossRef.
  2. M. Haruta, et al., Chem. Lett., 1987, 16, 405 CrossRef.
  3. M. Haruta, N. Yamada, T. Kobayashi and S. Lijima, J. Catal., 1989, 115, 301 CrossRef CAS.
  4. A. Orlov, D. A. Jefferson, N. Macleod and R. M. Lambert, Catal. Lett., 2004, 92, 41 CrossRef CAS.
  5. S. Praharaj, S. Nath, S. K. Ghosh, S. Kundu and T. Pal, Langmuir, 2004, 20, 9889 CrossRef CAS PubMed.
  6. J. Zeng, Q. Zhang, J. Chen and Y. Xia, Nano Lett., 2010, 10, 30 CrossRef CAS PubMed.
  7. J. Gong and C. B. Mullins, Acc. Chem. Res., 2009, 42, 1063 CrossRef CAS PubMed.
  8. E. Gross, X. Z. Shu, S. Alayoglu, H. A. Bechtel, M. C. Martin, F. D. Toste and G. A. Somorjai, J. Am. Chem. Soc., 2014, 136, 3624 CrossRef CAS PubMed.
  9. J. Li, C. Liu and Y. Liu, J. Mater. Chem., 2012, 22, 8426 RSC.
  10. K. Kuroda, T. Ishida and M. Haruta, J. Mol. Catal. A: Chem., 2009, 298, 7 CrossRef CAS.
  11. J. Zheng, Y. Dong, W. Wang, Y. Ma, J. Hu, X. Chen and X. Chen, Nanoscale, 2013, 5, 4894 RSC.
  12. F. Lin and R. Doong, J. Phys. Chem. C, 2011, 115, 6591 CAS.
  13. A. A. Ismail, A. Hakki and D. W. Bahnemann, J. Mol. Catal. A: Chem., 2012, 358, 145 CrossRef CAS.
  14. Y. Zhang, S. Liu, W. Lu, L. Wang, J. Tian and X. Sun, Catal. Sci. Technol., 2011, 1, 1142 CAS.
  15. L. Han, C. Zhu, P. Hu and S. Dong, RSC Adv., 2013, 3, 12568 RSC.
  16. Y. C. Chang and D. H. Chen, J. Hazard. Mater., 2009, 165, 664 CrossRef CAS PubMed.
  17. Y. Zhu, J. Shen, K. Zhou, C. Chen, X. Yang and C. Li, J. Phys. Chem. C, 2011, 115, 1614 CAS.
  18. K. Esumi, R. Isono and T. Yoshimura, Langmuir, 2004, 20, 237 CrossRef CAS PubMed.
  19. Y. Zhang, S. Liu, W. Lu, L. Wang, J. Tian and X. Sun, Catal. Sci. Technol., 2011, 1, 1142 CAS.
  20. G. Chang, Y. Luo, W. Lu, X. Qin, A. M. Asiri, A. O. Al-Youbibc and X. Sun, Catal. Sci. Technol., 2012, 2, 800 CAS.
  21. J. Zheng, Y. Dong, W. Wang, Y. Ma, J. Hu, X. Chen and X. Chen, Nanoscale, 2013, 5, 4894 RSC.
  22. S. Linic, P. Christopher and D. B. Ingram, Nat. Mater., 2011, 10, 911 CrossRef CAS PubMed.
  23. E. Marais and T. Nyokong, J. Hazard. Mater., 2008, 152, 293 CrossRef CAS PubMed.
  24. W. Sun, J. Li, G. Yao, M. Jiang and F. Zhang, Catal. Commun., 2011, 16, 90 CrossRef CAS.
  25. T. L. Lai, K. F. Yong, J. W. Yu, J. H. Chen, Y. Y. Shu and C. B. Wang, J. Hazard. Mater., 2011, 185, 366 CrossRef CAS PubMed.
  26. M. Panizza and G. Cerisola, Water Res., 2009, 43, 339 CrossRef CAS PubMed.
  27. Y. Zhang, L. Wu, W. Lei, X. Xia, M. Xia and Q. Hao, Electrochim. Acta, 2014, 146, 568 CrossRef CAS.
  28. Y. Du, H. Chen, R. Chen and N. Xu, Appl. Catal., A, 2004, 277, 259 CrossRef CAS.
  29. S. Saha, A. Pal, S. Kundu, S. Basu and T. Pal, Langmuir, 2010, 26, 2885 CrossRef CAS PubMed.
  30. J. Liu, Z. K. Sun, Y. H. Deng, Y. Zou, C. Y. Li, X. H. Guo, L. Q. Xiong, Y. Gao, Y. F. Li and D. Y. Zhao, Angew. Chem., Int. Ed., 2009, 48, 5875 CrossRef CAS PubMed.
  31. W. Li, J. P. Yang, Z. X. Wu, J. X. Wang, B. Li, S. S. Feng, Y. H. Deng, F. Zhang and D. Y. Zhao, J. Am. Chem. Soc., 2012, 134, 11864 CrossRef CAS PubMed.
  32. J. Kimling, M. Maier, B. Okenve, V. Kotaidis, H. Ballot and P. Plech, J. Phys. Chem. B, 2006, 110, 15700 CrossRef CAS PubMed.
  33. C. Ziegler and A. Eychmuller, J. Phys. Chem. B, 2011, 115, 4502 CAS.
  34. Y. Qiu, Z. Ma and P. Hu, J. Mater. Chem. A, 2014, 2, 13471 CAS.
  35. K. Hayakawa, T. Yoshimura and K. Esumi, Langmuir, 2003, 19, 5517 CrossRef CAS.
  36. S. Praharaj, S. Nath, S. K. Ghosh, S. Kundu and T. Pal, Langmuir, 2004, 20, 9889 CrossRef CAS PubMed.
  37. J. Ge, Q. Zhang, T. Zhang and Y. Yin, Angew. Chem., 2008, 120, 9056 CrossRef.
  38. J. M. Herrmann, Catal. Today, 1999, 53, 115 CrossRef CAS.
  39. K. S. Shin, Y. K. Cho, J.-Y. Choi and K. Kim, Appl. Catal., A, 2012, 413, 170 CrossRef.
  40. S. J. Zhuo, M. W. Shao and S. T. Lee, ACS Nano, 2012, 6, 1059 CrossRef CAS PubMed.
  41. D. M. Dotzauer, S. Bhattacharjee, Y. Wen and M. L. Bruening, Langmuir, 2009, 25, 1865 CrossRef CAS PubMed.
  42. Y. Mei, Y. Lu, F. Polzer and M. Ballauff, Chem. Mater., 2007, 19, 1062 CrossRef CAS.
  43. B. Naik, S. Hazra, V. S. Prasad and N. N. Ghosh, Catal. Commun., 2011, 12, 1104 CrossRef CAS.
  44. X. Du, J. He, J. Zhu, L. Sun and S. An, Appl. Surf. Sci., 2012, 258, 2717 CrossRef CAS.
  45. J. M. Herrmann, Top. Catal., 2006, 39, 3 CrossRef CAS.
  46. J. M. Herrmann, Top. Catal., 2005, 34, 49 CrossRef CAS.
  47. X. Zhang, Y. L. Chen, R.-S. Liu and D. P. Tsai, Rep. Prog. Phys., 2013, 76, 1 CrossRef PubMed.
  48. C. K. Sun, F. Vallée, L. H. Acioli, E. P. Ippen and J. G. Fujimoto, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 15337 CrossRef CAS.
  49. S. Link and M. A. El-Sayed, Annu. Rev. Phys. Chem., 2003, 54, 331 CrossRef CAS PubMed.
  50. P. Christopher, D. B. Ingram and S. Linic, J. Phys. Chem. C, 2010, 114, 9173 CAS.
  51. J. R. Adleman, D. A. Boyd, D. G. David and D. Psaltis, Nano Lett., 2009, 9, 4417 CrossRef CAS PubMed.
  52. C. W. Yen and M. A. El-Sayed, J. Phys. Chem. C, 2009, 113, 19585 CAS.
  53. J. R. Lombardi and R. L. Birke, J. Phys. Chem. C, 2008, 112, 5605 CAS.
  54. K. Awazu, M. Fujimaki, C. Rockstuhl, J. Tominaga, H. Murakami, Y. Ohki, N. Yoshida and T. Watanabe, J. Am. Chem. Soc., 2008, 130, 1676 CrossRef CAS PubMed.
  55. S. Sun, W. Wang, L. Zhang, M. Shang and L. Wang, Catal. Commun., 2009, 11, 290 CrossRef CAS.
  56. X. Zhang, Y. Zhu, X. Yang, Y. Zhou, Y. Yao and C. Li, Nanoscale, 2014, 6, 5971 RSC.
  57. X. D. Wang, T. Dornon, M. Blackford and R. A. Caruso, J. Mater. Chem., 2012, 22, 11701 RSC.
  58. O. Akhavan and E. Ghaderi, Surf. Coat. Technol., 2010, 204, 3676 CrossRef CAS.
  59. X. D. Wang and R. A. Caruso, J. Mater. Chem., 2011, 21, 20 RSC.

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