Continuous synthesis of hedgehog-like Ag–ZnO nanoparticles in a two-stage microfluidic system

Sha Taoab, Mei Yang*a, Huihui Chenab, Mingyue Renab and Guangwen Chen*a
aDalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China. E-mail: yangmei@dicp.ac.cn; gwchen@dicp.ac.cn; Fax: +86-411-8469-1570; Tel: +86-411-8437-9031
bGraduate University, Chinese Academy of Sciences, Beijing 100049, China

Received 8th March 2016 , Accepted 29th April 2016

First published on 3rd May 2016


Abstract

Hedgehog-like Ag–ZnO nanoparticles (NPs) were successfully prepared in a continuous two-stage microfluidic system. The first stage was the synthesis of monodispersed triangular Ag nanoprisms by the reduction of AgNO3 with NaBH4 in the presence of trisodium citrate, H2O2 and sodium dodecyl sulfate. The second stage was the growth of Zn(OH)2 nanorods on triangular Ag nanoprisms to form a hedgehog-like shape. Hedgehog-like Ag–ZnO NPs were finally obtained by the thermal decomposition of as-prepared hedgehog-like Ag–Zn(OH)2 NPs. The structure, morphology and chemical state of the obtained samples were characterized by X-ray diffraction, transmission electron microscopy and X-ray photoelectron spectroscopy. It was found that H2O2, the molar ratio of NaOH to Zn(NO3)2 and the flow rate ratio of AgNO3 to Zn(NO3)2 played a crucial role in the preparation of hedgehog-like Ag–ZnO NPs. In addition, the photocatalytic activity of hedgehog-like Ag–ZnO NPs was evaluated in the degradation of methyl orange. The results showed that the photocatalytic activity was greatly enhanced by hedgehog-like Ag–ZnO NPs in comparison to pure ZnO and spherical-like Ag–ZnO NPs. The reaction rate constant of the hedgehog-like Ag–ZnO NPs was twelve times the value of pure ZnO and twice the value of spherical-like Ag–ZnO NPs.


1. Introduction

As an important semiconductor with a direct wide band gap (3.3 eV) and high exciton binding energy (60 meV), ZnO has great potential for application in the fields of catalysis, sensing, medicine and piezoelectricity.1,2 In particular, ZnO has been explored as an efficient photocatalyst for the degradation of organic compounds because of its high activity, low cost and environmental friendliness. Generally, the photocatalytic activity of ZnO is closely related to the amount of electron–hole pairs excited by the photons whose energy is equal to or higher than the band gap of ZnO. Unfortunately, the proton-induced electron–hole pairs usually recombine quickly and thus decrease the photocatalytic activity of ZnO. To solve this problem, considerable efforts have been made, and one of the most effective ways is to synthesize noble metal–ZnO heterostructures.3–5 In a typical noble metal–ZnO heterostructure, the proton-induced electrons in the conduction band of ZnO can transfer to the noble metal due to the Schottky barrier formed at the noble metal–ZnO interface, which can prevent the electrons and holes from recombination. Therefore, the noble metal can act as an electron sink, improving the charge separation efficiency and thereby enhancing the photocatalytic activity.6,7 In addition, doping amounts of noble metal on ZnO can also result in lattice defect and impurity energy level. Consequently, it can improve the efficiency of photons and expand the range of absorbance to visible light.

From the published literature dedicated to noble metal–ZnO heterostructure, various kinds of noble metal–ZnO nanostructures such as Au–ZnO and Pt–ZnO have been successfully prepared.8–11 Among the noble metals, Ag is most widely used to modify ZnO due to its low cost and high efficiency. In the past decades, Ag–ZnO with different morphologies such as worm-like NPs, nanorods and core–shell NPs have been successfully developed.12–15 The strategies for Ag–ZnO synthesis can be classified as three types: (1) the first step of the formation of ZnO followed by the deposition of Ag NPs on ZnO; (2) the first step of the formation of Ag seeds followed by the deposition of ZnO on Ag seeds; (3) one-pot synthesis of Ag–ZnO. In the first strategy, ZnO with different morphologies were firstly prepared and then Ag NPs were deposited on ZnO by photo-reduction or hydrothermal/solvothermal method. Zheng et al. presented a facile solvothermal method to form Ag–ZnO heterostructure where NaOH/ethanol was utilized as the solvent.16 Liang et al. reported the synthesis of porous 3D flower-like Ag–ZnO heterostructure by two-steps, namely hydrothermal and photo-chemical deposition methods.17 This 3D flower-like Ag–ZnO heterostructure showed an obviously improved activity in comparison with pure ZnO in the degradation of Rhodamine B. Chamjangali et al. prepared Ag–ZnO multipodes via a solution route/photo-reduction process and used them as an efficient photocatalyst in simultaneous methylene blue and methyl orange degradation.14 In the second strategy, Ag seeds were firstly prepared by reducing the precursors containing Ag+ ions with reducing agents in the presence of stabilizers such as PVP and trisodium citrate. Subsequently, ZnO were deposited on Ag seeds by precipitation or hydrothermal method. Mondragón-Galicia et al. synthesized unidirectional Ag–ZnO nanostructured brushes on the basis of the use of unidimensional Ag nanowires as precursors.18 Yin et al. prepared Ag nanoparticle/ZnO nanorods nanocomposites with a core/shell structure via a seed-mediated method.13 In their study, Ag NPs were firstly synthesized by a polyol synthesis and used as seeds for further growth of ZnO nanorods. To simplify the process for Ag–ZnO synthesis, one-pot synthesis was proposed. Lu et al. demonstrated the synthesis of 3D hollow microspheres Ag–ZnO NPs assisted by sodium alginate via one-pot method.19 Huang et al. reported a one-pot synthesis of branched Ag–ZnO heterojunction nanostructures with Ag NPs of 500 nm as cores.20 The branched Ag–ZnO heterojunction nanostructures showed a superior photocatalytic activity over the branched ZnO. Obviously, great efforts have been put in and huge advances have been obtained so far. However, there is still plenty of room to optimize the process for Ag–ZnO synthesis. Conventionally, the preparation of Ag–ZnO including the literatures mentioned above is carried out in a batch reactor. As is well known, it is difficult for batch reactors to precisely control the NPs' particle size distribution and morphology due to the limitation of mixing and heat transfer. Additionally, the discontinuous operation of the batch process always results in a low productivity and poor reproducibility, which limits the large-scale production of nanoparticles.

Recently, microreactor has attracted much attention in the synthesis of nanoparticles, such as CdSe, PdS, Ag, TiO2, CaCO3 and layered double hydroxides, etc.21–26 Additionally, some hybrid nanoparticles have been successfully synthesized in the microfluidic system including Ag–Zn2GeO4 nanowires, FeAl@Al(1−x)FexOy and CoZn@Zn(1−x)CoxOy, etc.27–29 Microreactor has been considered as a potential tool to overcome the limitations associated with batch reactors in the nanoparticle synthesis. In comparison with the conventional batch reactor, microreactor has a plenty of advantages such as large surface-to-volume ratio, high heat and mass transfer rate, etc.30–32 Therefore, microreactor can offer a uniform reaction condition which makes the NP's size and shape more controllable. In addition, the continuous operation and numbering-up of microreactor are both beneficial for large-scale production of nanoparticles. In this paper, a facile and efficient method for continuous synthesis of hedgehog-like Ag–ZnO NPs in a microfluidic system was proposed. Hedgehog-like Ag–ZnO NPs were successfully obtained by a rational design and precise parameter control. To the best of our knowledge, no microfluidic-based method has been reported for the preparation of Ag–ZnO heterostructure.

2. Experimental

2.1. Materials

Zinc nitrate hexahydrate (Zn(NO3)2·6H2O), sodium hydrate (NaOH), silver nitrate (AgNO3), sodium borohydride (NaBH4), sodium dodecyl sulfate (C12H25SO4Na, SDS), trisodium citrate (Na3C6H5O7·2H2O), hydrogen peroxide (H2O2) and methyl orange (C14H14N3SO3Na, MO) were purchased with available purity. All chemicals were used as received without further treatment. Deionized water was used in all experiments.

2.2. Synthesis of Ag–ZnO NPs

Ag–ZnO NPs were synthesized in a continuous two-stage microfluidic system. As shown in Fig. 1, the two-stage microfluidic system was comprised of three consecutive microreactors (T-mixer, spiral microreactor and cross-type mixer), which were all fabricated by polytetrafluoroethylene. The geometric dimensions of different microreactors were summarized in Table 1.
image file: c6ra06101j-f1.tif
Fig. 1 Experimental arrangements used for the continuous preparation of Ag–ZnO NPs.
Table 1 Geometric dimensions of different microreactors
Microreactor Cross section Channel size/mm (inner diameter × length)
T-mixer Circle 0.6 × 10
Spiral microreactor 0.6 × 4000
Cross-type mixer 0.6 × 710


In the first stage, monodispersed triangular Ag nanoprisms were prepared in T-mixer and spiral microreactor following a modified protocol reported by Carboni et al.33 AgNO3 and NaBH4 were used as the precursor and reducing agent, respectively. In a typical experiment, two aqueous solutions were prepared firstly. One was 1 mmol L−1 NaBH4 solution with fixed pH (10–12) which was adjusted by adding adequate NaOH. The other was made by adding AgNO3, trisodium citrate, H2O2 and SDS into deionized water. The concentration of AgNO3, trisodium citrate, H2O2 and SDS was 0.5 mmol L−1, 0.2 mmol L−1, 80 mmol L−1 and 10 mmol L−1, respectively. The two as-prepared solutions were continuously fed into the two-stage microfluidic system with the same flow rate (VAgNO3) by two syringe pumps.

In the second stage, the obtained suspension containing triangular Ag nanoprisms was directly fed into cross-type mixer. At the same time, Zn(NO3)2 aqueous solution and NaOH aqueous solution were injected into cross-type mixer with the same flow rate (VZn(NO3)2). The concentration of Zn(NO3)2 was 50 mmol L−1, and the molar ratio of NaOH to Zn(NO3)2 varied from 2[thin space (1/6-em)]:[thin space (1/6-em)]1 to 20[thin space (1/6-em)]:[thin space (1/6-em)]1. The two stages mentioned above were both carried out at room temperature. The precipitation out of the cross-type mixer was collected by centrifugation and washed three times with deionized water and ethanol alternately. Ag–ZnO NPs were finally obtained by drying the precipitation at 100 °C for 5 h. For comparison, the precipitation was also freezing-dried and the obtained sample was denoted as Ag–Zn(OH)2 NPs. Additionally, pure ZnO was also prepared following the same protocol in this two-stage microfluidic system. The NaBH4 and AgNO3 solutions were replaced by deionized water. ZnO was synthesized under a condition of a molar ratio of NaOH to Zn(NO3)2 equal to 2[thin space (1/6-em)]:[thin space (1/6-em)]1.

2.3. Characterizations

The powder X-ray diffraction patterns (XRD) of the as-prepared samples were recorded by an X-ray diffractometer (X'pert Pro) at a scanning rate of 5° min−1 for 2θ ranging from 30° to 90°. The morphologies of the samples were characterized by transmission electron microscopy (JEOLJEM-2000EX) with the accelerating voltage of 120 kV. X-ray photoelectron spectroscopy (XPS) analysis was performed on an ESCALAB 250Xi system, using Al Ka radiation as the X-ray source, which was used to analyze elemental and chemical states of the samples. Inductively coupled plasma-optical emission spectroscopy (ICP-OES, PerkinElmer ICP-OES 7300DV) was employed to measure the concentration of Zn2+ ions in the aqueous solution.

2.4. Photocatalytic activity test

The photocatalytic performance of as-prepared Ag–ZnO NPs was evaluated by using MO as presentative dye pollutant. A 300 W Xenon lamp (Ushio-CERMAXLX300) was used as the visible light source. In a typical procedure, 30 mg catalyst was dispersed into 100 mL MO solution with the concentration of 20 mg L−1. After stirring in the dark for 3 h, the suspensions were placed under the xenon lamp with a constant stirring rate. During the photoreaction, the samples were withdrawn and the photocatalyst was separated from the suspension by centrifuging. The residual MO in the solution was measured by the UV-vis spectroscopy (UV4802) to evaluate the photocatalytic degradation process.

3. Results and discussion

3.1. The characterization results of hedgehog-like Ag–ZnO NPs

According to the synthesis conditions of the hedgehog-like Ag–ZnO NPs, the residence time of the solution in T-mixer, spiral microreactor and cross-type mixer was calculated to be 0.2 s, 67.8 s and 7.8 s, respectively. The total residence time was equal to 75.8 s, while the conventional batch method always took more than one hour to prepare Ag–ZnO NPs with different shapes. This indicated that the continuous synthesis of Ag–ZnO NPs in the two-stage microfluidic system was highly efficient. Fig. 2 depicts the XRD patterns of hedgehog-like Ag–Zn(OH)2 NPs and Ag–ZnO NPs. Two crystalline phases of Zn(OH)2 and ZnO are detected in the XRD pattern of Ag–Zn(OH)2 NPs. It can be observed that the intensity of the diffraction peaks of ZnO is much smaller than that of Zn(OH)2. This indicates that Ag–Zn(OH)2 NPs are mainly consisted of Zn(OH)2. The existence of Ag cannot be identified due to the overlapping of the characteristic peaks of Zn(OH)2 and Ag at 38.384°, 44.202°, 64.482°, 77.264° and 81.359°. For Ag–ZnO NPs, the diffraction peaks of Ag are clearly observed. In addition, the diffraction peaks at 31.772°, 34.420° and 36.256° can be indexed as the characteristic diffraction peaks of ZnO. No diffraction peaks corresponding to Zn(OH)2 are observed. This implies that Zn(OH)2 can be totally decomposed into ZnO at 100 °C. To visually observe the morphology of the obtained samples, transmission electron microscopy (TEM) was performed. Fig. 3 shows the TEM images of Ag seeds, hedgehog-like Ag–Zn(OH)2 NPs and Ag–ZnO NPs. As shown in Fig. 3A, Ag seeds are mainly consist of triangular Ag nanoprisms with tip sniping. These triangular Ag nanoprisms are uniformly dispersed with an edge length between 20 nm and 50 nm. Ag–Zn(OH)2 NPs and Ag–ZnO NPs both exhibit a hedgehog-like morphology with an average size of 500 nm (Fig. 3B and C). Each individual nanoparticle is composed of several branches, and the size of each branch is 100 nm approximately. Obviously, the obtained Ag–ZnO NPs maintained the shape of Ag–Zn(OH)2 NPs, indicating that this hedgehog-like morphology was stable in the thermal decomposition of Zn(OH)2 to ZnO. A similar morphology was also obtained in the studies of Huang et al.20 and Fan et al.34 They suggested that ZnO nanorods preferred to grow on the basis of Ag (1 1 1) facets because of the good lattice and symmetry match between ZnO and Ag in the corresponding planes. As discussed above, hedgehog-like Ag–Zn(OH)2 NPs and Ag–ZnO NPs were both successfully obtained in this study. It was well known that Ag nanoprisms were oriented along the (1 1 1) facets. Therefore, it can be concluded that Zn(OH)2 nanorods besides ZnO nanorods can also grow on the (1 1 1) facets of triangular Ag nanoprisms.
image file: c6ra06101j-f2.tif
Fig. 2 XRD patterns of hedgehog-like (A) Ag–Zn(OH)2 NPs (B) Ag–ZnO NPs. 2VAgNO3 = VZn(NO3)2 = 1.0 mL min−1, molar ratio of NaOH/Zn(NO3)2 = 20[thin space (1/6-em)]:[thin space (1/6-em)]1.

image file: c6ra06101j-f3.tif
Fig. 3 TEM images of (A) Ag seeds (B) hedgehog-like Ag–Zn(OH)2 NPs (C) hedgehog-like Ag–ZnO NPs. 2VAgNO3 = VZn(NO3)2 = 1.0 mL min−1, molar ratio of NaOH/Zn(NO3)2 = 20[thin space (1/6-em)]:[thin space (1/6-em)]1.

The XPS analysis was carried out to investigate the surface composition and chemical state of the as-prepared hedgehog-like Ag–ZnO NPs. Fig. 4A represents the scan survey spectra for the hedgehog-like Ag–ZnO NPs. All of the peaks on the curve can be attributed to Ag, Zn, O and C elements. The presence of C element is mainly derived from the hydrocarbon contaminants inherently existing in the XPS survey. It can be further confirmed that the sample is merely composed of Ag, Zn and O elements, which is in good agreement with the XRD results. The high-resolution spectrum of Ag element in the hedgehog-like Ag–ZnO NPs is presented in Fig. 4B. The two peaks centered at 365.1 eV and 371.0 eV can be ascribed to Ag 3d5/2 and Ag 3d3/2, respectively. It can be seen that there is a splitting of 5.9 eV between Ag 3d5/2 and Ag 3d3/2, demonstrating the metallic nature of Ag element. In comparison with the bulk Ag (the standard binding energies for Ag 3d5/2 and Ag 3d3/2 are 368.2 eV and 374.2 eV, respectively), the binding energies of Ag 3d5/2 and Ag 3d3/2 shift to the lower value remarkably, indicating a smaller electron density of Ag in hedgehog-like Ag–ZnO NPs than that of bulk Ag. This phenomenon can be attributed to the strong interaction between Ag and ZnO.35,36 Since the work function of ZnO was higher than that of Ag, the electrons would transfer from Ag to ZnO the moment Ag interacted with ZnO to equilibrate the Fermi level. As shown in Fig. 4C, the binding energies of Zn 2p1/2 and Zn 2p3/2 are 1019.3 eV and 1042.2 eV, respectively. These values are similar to those of pure ZnO,37 suggesting that Zn element mainly exists as Zn2+ ions. As Fig. 4D showed, the XPS peak of O 1s is asymmetric, indicating that there are two kinds of oxygen species on the sample surface. The peaks centered at 528.0 eV and 529.6 eV can be ascribed to the lattice oxygen of ZnO and oxygen of the hydroxyl species chemisorbed on the surface.38 The hydroxyl species were found to be of crucial importance for the high efficiency of photocatalysis. They can not only trap the photoinduced electrons and holes but also generate the active hydroxyl radicals.39–42


image file: c6ra06101j-f4.tif
Fig. 4 (A) XPS survey spectrum of the as-prepared hedgehog-like Ag–ZnO NPs (B) Ag 3d spectra (C) Zn 2p spectra and (D) O 1s spectra.

3.2. The effects of the operation parameters

3.2.1. H2O2. To validate the effect of H2O2, Ag seeds and Ag–ZnO NPs were prepared without H2O2, while other conditions remain unchanged. The TEM images of Ag seeds and Ag–ZnO NPs synthesized in the absence of H2O2 are depicted in Fig. 5. It can be clearly seen that no triangular Ag nanoprisms but Ag nanospheres are obtained in the absence of H2O2. Correspondingly, Ag–ZnO NPs exhibit spherical-like shape. A similar result was also found by Huang et al. in their study.20 Obviously, H2O2 played a crucial role not only in the preparation of Ag triangular nanoprisms but also hedgehog-like Ag–ZnO NPs. It was acknowledged that H2O2 was essential for the synthesis of triangular Ag nanoprisms. H2O2 could remove the less stable Ag seeds and induce to generate Ag seeds with stacking faults which finally grew into triangular Ag nanoprisms. In the absence of H2O2, Ag+ ions were prone to grow into Ag nanospheres by the reduction of NaBH4 in the presence of trisodium citrate and SDS. For the synthesis of hedgehog-like Ag–ZnO NPs, it was supposed that H2O2 also worked as an etchant and helped produce triangular Ag nanoprisms with stacking faults, which were beneficial for the growth of Zn(OH)2 nanorods.
image file: c6ra06101j-f5.tif
Fig. 5 TEM images of (A) Ag seeds (B) Ag–ZnO NPs. 2VAgNO3 = VZn(NO3)2 = 1.0 mL min−1, molar ratio of NaOH to Zn(NO3)2 = 20[thin space (1/6-em)]:[thin space (1/6-em)]1 and no H2O2 was used.
3.2.2. The molar ratio of NaOH to Zn(NO3)2. The effect of the molar ratio of NaOH to Zn(NO3)2 on the synthesis of hedgehog-like Ag–ZnO NPs was also investigated. It can be seen from Fig. 6 that when the molar ratio of NaOH to Zn(NO3)2 equals to the stoichiometric ratio of 2, irregular aggregated Ag–ZnO NPs are obtained. As the molar ratio of NaOH to Zn(NO3)2 increases to 15[thin space (1/6-em)]:[thin space (1/6-em)]1 and 20[thin space (1/6-em)]:[thin space (1/6-em)]1, irregular aggregated Ag–ZnO NPs disappear and hedgehog-like Ag–ZnO NPs are found. It is evident that the molar ratio of NaOH to Zn(NO3)2 plays a crucial role in the synthesis of hedgehog-like Ag–ZnO NPs. Generally, the morphology of nanoparticles was dependent on the nucleation and growth kinetics. When the molar ratio of NaOH to Zn(NO3)2 was equal to 2, a large number of Zn(OH)2 nuclei were formed once NaOH and Zn(NO3)2 contacted. These Zn(OH)2 nuclei grew into large particles and agglomerated gradually. Thus, irregular aggregated Ag–ZnO NPs were finally obtained. When the molar ratio of NaOH to Zn(NO3)2 was much larger than 2, Zn(OH)2 quickly dissolved into the solution containing excess NaOH to form Zn(OH)42−. In previous studies, ZnO with different morphologies were successfully synthesized with the molar ratio of NaOH to Zn(NO3)2 much larger than 2. Zn(OH)42− could be transformed to ZnO under the hydrothermal conditions. For example, Li et al. synthesized one-dimensional needle-like ZnO nanowhiskers by aging a solution of Zn(OH)42− and SDS at 85 °C for 5 h.43 Ge et al. reported that sisal-like 3D ZnO nanostructures were fabricated via the assemble of CTA+ (C16H33(CH3)3N+) and Zn(OH)42− in the temperature range from 120 to 220 °C.44 In this study, a high molar ratio of NaOH to Zn(NO3)2 was also found to be essential for the synthesis of hedgehog-like Ag–ZnO NPs. Taking the molar ratio of NaOH to Zn(NO3)2 of 20 as an example, Zn(OH)2 was the major component of the freeze-dried sample according to the XRD results. Evidently, this phenomenon was interesting because Zn(OH)2 could not exist at this strong basic condition. To investigate the mechanism of the transformation of Zn(OH)42− to Zn(OH)2, a transparent solution containing Zn(OH)42− was prepared by adding Zn(NO3)2 into NaOH aqueous solution with the molar ratio of NaOH to Zn(NO3)2 of 20. Trisodium citrate, SDS and NaBH4 were added into this transparent solution respectively and no precipitant was found, implying that the transformation of Zn(OH)42− to Zn(OH)2 was closely related to the presence of Ag seeds. The suspension out of the cross-type mixer was directly centrifuged, and the concentration of Zn2+ ions in the supernatant liquid was measured by ICP-OES. The concentration of Zn2+ ions in the supernatant liquid was 2.72 mmol L−1, which was equal to 5% of initial concentration of Zn2+ ions. This implied that most Zn2+ ions existed in the form of Zn(OH)2. Unfortunately, the reason why Ag seeds could result in the transformation of Zn(OH)42− to Zn(OH)2 was not clear and the relative study was under way. In addition, ZnO was also found in the freeze-dried sample out of the cross-type mixer. As is well known, Zn(OH)2 can transform to ZnO under strong basic condition even at room temperature.45 Therefore, it was anticipated that a small portion of Zn(OH)2 at the outer surface of the nanoparticles transformed to ZnO and thus protected Zn(OH)2 from being dissolved under this strong basic condition.
image file: c6ra06101j-f6.tif
Fig. 6 TEM images of Ag–ZnO NPs synthesized at different molar ratio of NaOH to Zn(NO3)2 (A) 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (B) 5[thin space (1/6-em)]:[thin space (1/6-em)]1 (C) 15[thin space (1/6-em)]:[thin space (1/6-em)]1 (D) 20[thin space (1/6-em)]:[thin space (1/6-em)]1. 2VAgNO3 = VZn(NO3)2 = 1 mL min−1.
3.2.3. The effect of the flow rate ratio of Ag(NO3)2 to Zn(NO3)2. For the two-stage microfluidic system, the Ag content can be easily adjusted by changing the flow rate ratio of Ag(NO3)2 to Zn(NO3)2. Therefore, it was very necessary to investigate the effect of the flow rate ratio of Ag(NO3)2 to Zn(NO3)2 on the morphology of Ag–ZnO NPs. The TEM images of Ag–ZnO NPs synthesized at different flow rate ratio of Ag(NO3)2 to Zn(NO3)2 are shown in Fig. 7. As shown in Fig. 7, hedgehog-like Ag–ZnO can be successfully prepared with the flow rate ratio of Ag(NO3)2 to Zn(NO3)2 increasing from 0.5[thin space (1/6-em)]:[thin space (1/6-em)]1 to 1[thin space (1/6-em)]:[thin space (1/6-em)]1. In addition, the width of branches increases to some extent with the increase in the flow rate ratio of Ag(NO3)2 to Zn(NO3)2.
image file: c6ra06101j-f7.tif
Fig. 7 TEM images of Ag–ZnO NPs synthesized at different flow rate ratio of Ag(NO3)2 to Zn(NO3)2 (A) 0.5[thin space (1/6-em)]:[thin space (1/6-em)]1 (B) 0.75[thin space (1/6-em)]:[thin space (1/6-em)]1 (C) 1[thin space (1/6-em)]:[thin space (1/6-em)]1. Molar ratio of NaOH/Zn(NO3)2 = 20[thin space (1/6-em)]:[thin space (1/6-em)]1.

3.3. Photocatalytic activity test

In order to examine the photocatalytic performance of hedgehog-like Ag–ZnO NPs, the photocatalytic degradation of MO as representative experiment was launched. It is well known that MO is widely used in the textile industry, which is a typical organic pollutant and always discharged into the industrial effluent. Herein, the activity of hedgehog-like Ag–ZnO NPs (prepared in Section 3.1), spherical-like Ag–ZnO NPs (prepared in Section 3.2.1) and pure ZnO for the photocatalytic degradation of MO were compared and the results are shown in Fig. 8, where C represents the concentration of MO remaining in the solution after irradiation and C0 represents the initial concentration of MO. It can be seen that pure ZnO exhibits a low photocatalytic activity for the degradation of MO under this condition. Only 10% MO is degraded within 90 min. The addition of Ag is found to significantly improve the performance of pure ZnO. Nearly 50% and 70% MO are degraded within 90 min over the spherical-like Ag–ZnO NPs and hedgehog-like Ag–ZnO NPs, respectively. Noticeably, the hedgehog-like Ag–ZnO NPs show the highest photocatalytic activity. Generally, the photocatalytic degradation kinetic equation could be described as
 
ln(C0/C) = kt (1)
where k and t represent the pseudo-first-rate kinetic constant and irradiation time, respectively. Therefore, the slope of linear curve in Fig. 8B could be considered as the reaction rate constant. As summarized in Table 2, the reaction rate constant of the hedgehog-like Ag–ZnO NPs is twelve times the value of pure ZnO and twice the value of spherical-like Ag–ZnO NPs. So far, numerous researches have demonstrated that Ag–ZnO heterostructure showed a superior photocatalytic activity than ZnO by inhibiting the recombination of the photoinduced electrons and holes. The better performance of spherical-like Ag–ZnO NPs and hedgehog-like Ag–ZnO NPs in this study could be ascribed to the similar reason. Furthermore, the hedgehog-like Ag–ZnO NPs could offer a much larger opportunities for the interaction of MO to ZnO than spherical-like Ag–ZnO NPs. As a consequence, the catalytic activity of hedgehog-like Ag–ZnO NPs was higher than that of spherical-like Ag–ZnO NPs.

image file: c6ra06101j-f8.tif
Fig. 8 (A) Photocatalytic activity and (B) kinetics of the as-prepared hedgehog-like Ag–ZnO NPs, spherical-like Ag–ZnO NPs and pure ZnO for the degradation of MO under visible light irradiation.
Table 2 Reaction rate constant (k) for the photocatalytic degradation of MO under visible light irradiation
Catalyst Blank ZnO Spherical-like Ag–ZnO NPs Hedgehog-like Ag–ZnO NPs
k (min−1) 6.7 × 10−4 1.0 × 10−3 6.4 × 10−3 1.2 × 10−2


The mechanism of the photocatalytic degradation of MO over hedgehog-like Ag–ZnO NPs is presented in Fig. 9. Under the visible light irradiation, ZnO could be photoactivated, and the electrons in the valence band (VB) could be excited to the conduction band (CB) of ZnO. Meanwhile, there left the same amount of holes in the VB of ZnO. Since the bottom energy level of the CB of ZnO was higher than the new equilibrium Fermi energy level of Ag–ZnO, the photoinduced electrons in the CB of ZnO transferred to the metallic Ag NPs. As a result, the recombination of electrons and holes was inhibited which prolonged the lifetime of photo-electron pairs. During the photocatalytic degradation of MO, the electrons in the CB of ZnO or trapped by Ag NPs could be captured by the adsorbed oxygen on the surface to form superoxide radicals anion (˙O2), and the holes left in the VB of ZnO could be trapped by the surface hydroxyl to form the hydroxyl radicals (˙OH) which was an extraordinary strong oxidant for the degradation of MO.


image file: c6ra06101j-f9.tif
Fig. 9 Schematic diagram of the electron–hole pairs separation and catalysis mechanism of hedgehog-like Ag–ZnO NPs under visible light irradiation.

4. Conclusion

In summary, hedgehog-like Ag–ZnO NPs were continuously synthesized in a two-stage microfluidic system for the first time. It was found that H2O2, the molar ratio of NaOH to Zn(NO3)2 and the flow rate ratio of Ag(NO3)2 to Zn(NO3)2 had a significant effect on the synthesis of hedgehog-like Ag–ZnO NPs. Furthermore, hedgehog-like Ag–ZnO NPs was found to exhibit a better photocatalytic activity than spherical-like Ag–ZnO NPs and pure ZnO in the degradation of MO, which could be ascribed to the unique morphology as well as the strong interaction between Ag and ZnO. In addition, the mechanism of photocatalytic degradation of MO over hedgehog-like Ag–ZnO NPs was proposed. This brand-new method on basis of the microfluidic technology was continuous, highly efficient and easy to scale up.

Acknowledgements

The authors gratefully acknowledge the financial support from National Natural Science Foundation of China (No. 21406226, 21225627).

References

  1. G. P. Xian and W. Z. Lin, Mesoporous polyhedral cages and shells formed by textured self-assembly of ZnO nanocrystals, J. Am. Chem. Soc., 2003, 125, 11299–11305 CrossRef PubMed.
  2. X. Wang, C. J. Summers and Z. L. Wang, Large-Scale Hexagonal-Patterned Growth of Aligned ZnO Nanorods for Nano-optoelectronics and Nanosensor Arrays, Nano Lett., 2004, 4, 423–426 CrossRef CAS PubMed.
  3. B. Subramanian, P. Natarajan and S. Meenakshisundaram, Heteroarchitectured Ag–Bi2O3–ZnO as a bifunctional nanomaterial, RSC Adv., 2016, 6, 20247–20257 RSC.
  4. Y.-H. Chin, R. Dagle, J. Hu, A. C. Dohnalkova and Y. Wang, Steam reforming of methanol over highly active Pd/ZnO catalyst, Catal. Today, 2002, 77, 79–88 CrossRef CAS.
  5. Y. Zhang, Q. Xiang, J. Xu, P. Xu, Q. Pan and F. Li, Self-assemblies of Pd nanoparticles on the surfaces of single crystal ZnO nanowires for chemical sensors with enhanced performances, J. Mater. Chem., 2009, 19, 4701–4706 RSC.
  6. S. Y. Baek, S. Y. Chai, K. S. Hur and W. I. Lee, Synthesis of Highly Soluble TiO Nanoparticle with Narrow Size Distribution, Bull. Korean Chem. Soc., 2005, 26, 1333 CrossRef CAS.
  7. M. A. Fox and M. T. Dulay, Heterogeneous photocatalysis, Chem. Rev., 1993, 93, 341–357 CrossRef CAS.
  8. P. Li, Z. Wei, T. Wu, Q. Peng and Y. Li, Au–ZnO hybrid nanopyramids and their photocatalytic properties, J. Am. Chem. Soc., 2011, 133, 5660–5663 CrossRef CAS PubMed.
  9. C. Yuanzhi, Z. Deqian, Z. Kun, L. Aolin, W. Laisen and P. Dong-Liang, Au–ZnO hybrid nanoflowers, nanomultipods and nanopyramids: one-pot reaction synthesis and photocatalytic properties, Nanoscale, 2014, 6, 874–881 RSC.
  10. Y. Changlin, Y. Kai, X. Yu, F. Qizhe, J. C. Yu, S. Qing and W. Chunying, Novel hollow Pt–ZnO nanocomposite microspheres with hierarchical structure and enhanced photocatalytic activity and stability, Nanoscale, 2013, 5, 2142–2151 RSC.
  11. M. Mujahid, Synthesis, characterization and electrical properties of visible-light-driven Pt–ZnO/CNT, Bull. Mater. Sci., 2015, 38, 995–1001 CrossRef CAS.
  12. H. R. Liu, G. X. Shao, J. F. Zhao, Z. X. Zhang, Y. Zhang, J. Liang, X. G. Liu, H. S. Jia and B. S. Xu, Worm-Like Ag/ZnO Core–Shell Heterostructural Composites: Fabrication, Characterization, and Photocatalysis, J. Phys. Chem. C, 2012, 116, 16182–16190 CAS.
  13. X. Yin, W. Que, D. Fei, F. Shen and Q. Guo, Ag nanoparticle/ZnO nanorods nanocomposites derived by a seed-mediated method and their photocatalytic properties, J. Alloys Compd., 2012, 524, 13–21 CrossRef CAS.
  14. M. A. Chamjangali, G. Bagherian, A. Javid, S. Boroumand and N. Farzaneh, Synthesis of Ag–ZnO with multiple rods (multipods) morphology and its application in the simultaneous photo-catalytic degradation of methyl orange and methylene blue, Spectrochim. Acta, Part A, 2015, 150, 230–237 CrossRef PubMed.
  15. M. E. Aguirre, H. B. Rodríguez, E. S. Román, A. Feldhoff and M. A. Grela, Ag@ZnO Core–Shell Nanoparticles Formed by the Timely Reduction of Ag+ Ions and Zinc Acetate Hydrolysis in N,N-Dimethylformamide: Mechanism of Growth and Photocatalytic Properties, J. Phys. Chem. C, 2011, 115, 24967–24974 CAS.
  16. Z. Yuanhui, Z. Lirong, Z. Yingying, L. Xingyi, Z. Qi and W. Kemei, Ag/ZnO heterostructure nanocrystals: synthesis, characterization, and photocatalysis, Inorg. Chem., 2007, 46, 6980–6986 CrossRef PubMed.
  17. Y. Liang, N. Guo, L. Li, R. Li, G. Ji and S. Gan, Fabrication of porous 3D flower-like Ag/ZnO heterostructure composites with enhanced photocatalytic performance, Appl. Surf. Sci., 2015, 332, 32–39 CrossRef CAS.
  18. G. Mondragón-Galicia, C. Gutiérrez-Wing, M. E. Fernández-García, D. Mendoza-Anaya and R. Pérez-Hernández, Ag nanowires as precursors to synthesize Ag–ZnO nanostructured brushes, RSC Adv., 2015, 5, 42568–42571 RSC.
  19. W. Lu, S. Gao and J. Wang, One-Pot Synthesis of Ag/ZnO Self-Assembled 3D Hollow Microspheres with Enhanced Photocatalytic Performance, J. Phys. Chem. C, 2008, 112, 16792–16800 CAS.
  20. Q. Huang, Q. Zhang, S. Yuan, Y. Zhang and M. Zhang, One-pot facile synthesis of branched Ag–ZnO heterojunction nanostructure as highly efficient photocatalytic catalyst, Appl. Surf. Sci., 2015, 353, 949–957 CrossRef CAS.
  21. J. Baumgard, A.-M. Vogt, U. Kragl, K. Jähnisch and N. Steinfeldt, Application of microstructured devices for continuous synthesis of tailored platinum nanoparticles, Chem. Eng. J., 2013, 227, 137–144 CrossRef CAS.
  22. M. Ren, M. Yang, G. Chen and Q. Yuan, High-Throughput Preparation of Monodispersed Layered Double Hydroxides via Microreaction Technology, J. Flow Chem., 2014, 1, 1 CrossRef.
  23. G. Luo, L. Du, Y. Wang, Y. Lu and J. Xu, Controllable preparation of particles with microfluidics, Particuology, 2011, 9, 545–558 CrossRef CAS.
  24. T. Maki, J. I. Kitada and K. Mae, Preparation and Control of the Size Distribution of Zirconia Nanoparticles in a Concentric-Axle Dual-Pipe Microreactor, Chem. Eng. Technol., 2013, 36, 1027–1032 CrossRef CAS.
  25. Y. Song, J. Hormes and C. S. Kumar, Microfluidic synthesis of nanomaterials, Small, 2008, 4, 698–711 CrossRef CAS PubMed.
  26. C.-X. Zhao, L. He, S. Z. Qiao and A. P. Middelberg, Nanoparticle synthesis in microreactors, Chem. Eng. Sci., 2011, 66, 1463–1479 CrossRef CAS.
  27. J. Wang, Solution-assembled nanowires for high performance flexible and transparent solar-blind photodetectors, J. Mater. Chem. C, 2015, 3, 596–600 RSC.
  28. R. Wang, W. Yang, Y. Song, X. Shen, J. Wang, X. Zhong, S. Li and Y. Song, A General Strategy for Nanohybrids Synthesis via Coupled Competitive Reactions Controlled in a Hybrid Process, Sci. Rep., 2014, 5, 9189 CrossRef PubMed.
  29. X. Shen, Y. Song, S. Li, R. Li, S. Ji, Q. Li, H. Duan, R. Xu, W. Yang, K. Zhao, R. Rong and X. Wang, Spatiotemporal-resolved nanoparticle synthesis via simple programmed microfluidic processes, RSC Adv., 2014, 4, 34179–34188 RSC.
  30. K. F. Jensen, Microreaction engineering—is small better?, Chem. Eng. Sci., 2001, 56, 293–303 CrossRef CAS.
  31. H. Cao, G. Chen and Q. Yuan, Testing and Design of a Microchannel Heat Exchanger with Multiple Plates, Ind. Eng. Chem. Res., 2009, 48, 132–143 Search PubMed.
  32. Y. Zhao, G. Chen and Q. Yuan, Liquid–liquid two-phase mass transfer in the T-junction microchannels, AICHE J., 2007, 53, 3042–3053 CrossRef CAS.
  33. M. Carboni, L. Capretto, D. Carugo, E. Stulz and X. Zhang, Microfluidics-based continuous flow formation of triangular silver nanoprisms with tuneable surface plasmon resonance, J. Mater. Chem. C, 2013, 1, 7540–7546 RSC.
  34. F. Feng-Ru, D. Yong, L. De-Yu, T. Zhong-Qun and W. Zhong Lin, Facet-selective epitaxial growth of heterogeneous nanostructures of semiconductor and metal[thin space (1/6-em)]:[thin space (1/6-em)]ZnO nanorods on Ag nanocrystals, J. Am. Chem. Soc., 2009, 131, 12036–12037 CrossRef PubMed.
  35. R. K. Sahu, K. Ganguly, T. Mishra, M. Mishra, R. Ningthoujam, S. Roy and L. Pathak, Stabilization of intrinsic defects at high temperatures in ZnO nanoparticles by Ag modification, J. Colloid Interface Sci., 2012, 366, 8–15 CrossRef CAS PubMed.
  36. C. Chen, Y. Zheng, Y. Zhan, X. Lin, Q. Zheng and K. Wei, Enhanced Raman scattering and photocatalytic activity of Ag/ZnO heterojunction nanocrystals, Dalton Trans., 2011, 40, 9566–9570 RSC.
  37. C. Chongqi, Z. Yuanhui, Z. Yingying, L. Xingyi, Z. Qi and W. Kemei, Enhanced Raman scattering and photocatalytic activity of Ag/ZnO heterojunction nanocrystals, Dalton Trans., 2011, 40, 9566–9570 RSC.
  38. D.-M. Tang, G. Liu, F. Li, J. Tan, C. Liu, G. Q. Lu and H.-M. Cheng, Synthesis and photoelectrochemical property of urchin-like Zn/ZnO core–shell structures, J. Phys. Chem. C, 2009, 113, 11035–11040 CAS.
  39. F. Zhang, Y. Zheng, Y. Cao, C. Chen, Y. Zhan, X. Lin, Q. Zheng, K. Wei and J. Zhu, Ordered mesoporous Ag–TiO2–KIT6 heterostructure: synthesis, characterization and photocatalysis, J. Mater. Chem., 2009, 18, 2771–2777 RSC.
  40. S. Gao, X. Jia, S. Yang, Z. Li and K. Jiang, Hierarchical Ag/ZnO micro/nanostructure: Green synthesis and enhanced photocatalytic performance, J. Solid State Chem., 2011, 184, 764–769 CrossRef CAS.
  41. R. Manoj, R. Md Harunar, T. K. Paira, D. Enakshi and T. K. Mandal, Ascorbate-assisted growth of hierarchical ZnO nanostructures: sphere, spindle, and flower and their catalytic properties, Langmuir, 2010, 26, 8769–8782 CrossRef PubMed.
  42. D. Lin, H. Wu, R. Zhang and W. Pan, Enhanced Photocatalysis of Electrospun Ag–ZnO Heterostructured Nanofibers, Chem. Mater., 2009, 21, 15 CrossRef.
  43. P. Li, Y. Wei, H. Liu and X. Wang, A simple low-temperature growth of ZnO nanowhiskers directly from aqueous solution containing Zn(OH)42− ions, Chem. Commun., 2004, 24, 2856–2857 RSC.
  44. J. Ge, B. Tang, L. Zhuo and Z. Shi, A rapid hydrothermal route to sisal-like 3D ZnO nanostructures via the assembly of CTA+ and Zn(OH)42−: growth mechanism and photoluminescence properties, Nanotechnology, 2006, 17, 1316 CrossRef CAS.
  45. P. Li, Y. Wei, H. Liu and X. Wang, A simple Low-temperature growth of ZnO nanowhiskers directly from aqueous solution containing Zn(OH)42− ions, Chem. Commun., 2004, 24, 2856–2857 RSC.

This journal is © The Royal Society of Chemistry 2016