Xing-chun Zhou,
Xiao-hang Zhu,
Jing-wei Huang,
Xin-zhe Li,
Peng-fei Fu,
Li-xin Jiao,
Hong-fei Huo and
Rong Li*
College of Chemistry and Chemical Engineering Lanzhou University, The Key Laboratory of Catalytic engineering of Gansu Province, Lanzhou 730000, China. E-mail: liyirong@lzu.edu.cn; Fax: +86 0931 891 2582; Tel: +86 0931 891 2311
First published on 16th July 2014
Here, we report a facile synthesis approach to obtain a Pd@hTiO2 hollow mesoporous nanocomposite composed of tiny Pd nanoparticles cores encapsulated within hollow TiO2 mesoporous shells. The core–shell strategy efficiently prevents the aggregation of Pd NPs (Pd nanoparticles) in the high temperature calcination process and the leaching of Pd NPs for the catalytic reaction in a liquid phase. The catalyst was characterized by transmission electron microscopy (TEM), X-ray powder diffraction (XRD), X-ray photoelectron spectroscopy (XPS), N2 adsorption/desorption, elemental analysis and inductively coupled plasma atomic emission spectrometry (ICP-AES). The synthesized catalyst exhibited high catalytic activity in the reduction of p-nitrophenol with NaBH4 aqueous solution at room temperature. Furthermore, Pd@hTiO2 had an excellent recyclability, evidenced by being extensively reused for eight times without any substantial loss of activity.
Many researchers have reported noble metal related core–shell materials (e.g. Pt@C, Pt@SiO2, Pd@SiO2) which showed superior catalytic activity in hydrogenation, oxidation, reforming and coupling reactions.17,19,23,24 However, the unreactive silica or carbon shells only play a dispersion role of the active component, and the synergistic effect between the NMNPs and the shells was quite weak.25 Replacing the shells with reducible oxides, such as CeO226 and TiO2,26,27 to enhance the synergistic effect between the NMNPs and the shells may further improve the catalytic ability. TiO2 is attractive due to its fascinating features such as various polymorphs, good chemical and thermal stability, low cost, low toxicity, and excellent electronic and optical properties.28 In addition, TiO2 has advantageous properties for practical catalytic applications.29 TiO2 hollow nanostructures possess high active surface area, reduced diffusion resistance, and improved accessibility, which provide many new opportunities for the design of a highly active nanostructure.30 Because of these properties, the preparation of metal core/TiO2-shell structures is potentially of great importance. In particular, dispersible TiO2 encapsulated Pd has not been reported to date, although this system is very interesting for its numerous applications especially in catalysis.
Herein, we report the synthesis of Pd@TiO2 core–shell nanocomposites with inner hollow space (denoted as Pd@hTiO2) via a facile sol–gel method followed by calcination at high temperature in air and etched by NaOH to remove the template. And they were confirmed by corresponding characterization means. The hollow nanospheres showed good dispersity of NMNPs, enhanced synergistic effect, and decreased leaching of noble metals, leading to extremely high activity and stability for catalytic application. In addition, Pd@hTiO2 exhibited an excellent reusable performance. Its catalytic activity and recyclability were evaluated through the reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP).
To determine the catalytic recycling properties, the catalyst was separated by centrifugation after the reaction completed, and washed thoroughly with water and ethanol, followed by drying overnight at 30 °C in vacuum oven. Finally, the catalyst was redispersed in a new reaction system for subsequent catalytic experiments under the same reaction conditions.
![]() | ||
Fig. 1 TEM images of (A) SiO2, (B) Pd/SiO2, (C) Pd/SiO2@TiO2, (D–F) Pd@hTiO2 catalysts treated at 298 K, 673 K, 723 K, 773 K, 823 K, 873 K, respectively. |
The elemental composition of the Pd@hTiO2 samples was determined by EDX analysis. The result shown in Fig. 2 reveals that the as-prepared products contain Pd, Cu, Ti, C and O. Among these elements, Cu, C and O are generally influenced by the copper network support films and their degree of oxidation, Ti, O and Pd signals result from the Pd@hTiO2. And from the EDX result, the signal of Si element has not been found, revealed that the SiO2 was totally removed.
Because of the encapsulated Pd nanoparticles can't be visualized very evidently in the TEM images of Pd@hTiO2 hollow core–shell nanocomposites. Therefore, the elemental mapping has been performed to reveal the element distribution in the Pd@hTiO2 spheres. The mapping results [Fig. 3(B–D)] indicate that the elements Ti, O and Pd spread evenly in the whole spheres and the Pd nanoparticles are not located on the outermost surface of the hollow spheres, which confirms the Pd core@TiO2 shell structure as obtained by the step-by-step assembly procedures illustrated in Scheme 1.
![]() | ||
Fig. 3 (A) the high angle annular dark field-scanning transmission electron microscopy (HAADF-STEM) image and mapping results of the elements (B) Ti, (C) O and (D) Pd for the Pd@hTiO2. |
Fig. 4 shows the XRD patterns between 10° and 90° of the samples calcined at various temperatures. The sharp, strong peaks confirm the products are well crystallized. The XRD patterns show that the appearance of the peaks at 2θ = 40.1°, 46.5° and 68.0°corresponding to the reflections of (111), (200) and (220) crystal planes of palladium respectively, and mostly corresponded to face-centered cubic metallic palladium diffraction. For the sample before calcination, diffraction peaks due to the crystalline phase are not observed, suggesting that the sample is still in the amorphous phase. When the sample was calcined at 400 °C, weak and broad peaks at 2θ = 25.5°, 37.9°, 48.2°and 53.8° were observed. These peaks represent the indices of (101), (004), (200) and (105) planes of anatase phase, respectively.34 The average crystallite size of Pd in Pd@hTiO2 calculated using Scherrer's equation, is about 6 nm. When the calcination temperature was increased, the diffraction peaks due to anatase phase became narrow and intense in intensity. This indicates that the crystallinity of the anatase phase is further improved.35 When the sample was calcined at 773 K, the new weak peak was observed at 2θ = 27.6°, which correspond to the indices of (110) planes of rutile phase.36 This indicates that the anatase phase starts to transform into the rutile phase at 773 K.
![]() | ||
Fig. 4 XRD patterns of Pd@hTiO2 catalysts calcined at 673 K, 723 K, 773 K, 823 K and 873 K and dried at 298 K. |
In addition, the surface area and porosity of the Pd@hTiO2 hollow core–shell nanocomposites have been characterized. Fig. 5 displayed the N2 adsorption–desorption isotherms and the corresponding pore-size distribution curve for Pd@hTiO2 calcined at 773 K. The pore size was calculated by using the BJH method, as shown in the inset of Fig. 5. As a result, average pore size was about 5 nm for Pd@hTiO2. The BET surface areas and the cumulative pore volumes of Pd@hTiO2 were 105 m2 g−1 and 0.37 cm3 g−1, respectively.
![]() | ||
Fig. 5 Nitrogen adsorption–desorption isotherm of the Pd@hTiO2 calcined at 773 K (inset: the corresponding pore diameter distribution). |
Fig. 6 shows the XPS spectrum of the synthesized Pd/hTiO2 nanoreactor calcined at 773 K. Peaks corresponding to O, C, Pd, Ti and Sn are observed. However, typical peaks of Sn and Pd elements are not obviously found. This indicates that Sn, and Pd particles have been coated by TiO2, since the analysis depth of XPS is only several nanometers. To ascertain the oxidation state of the Pd, X-ray photoelectron spectroscopy (XPS) studies were carried out. The XPS signature of Pd 3d doublet for the Pd nanoparticles was given in the inset figure. The Pd 3d3/2 and Pd 3d5/2 peaks appeared at 342.6 and 337.2 eV respectively were the characteristic peaks for the metallic Pd. It can be seen that there is no oxidation state of Pd species in the catalyst. In the experiment, the reducing agent is excessive that can reduce the Pd(II) into Pd(0) completely, which is beneficial for the reaction. So the oxidation state of Pd species is hard to find in the figure.
![]() | ||
Fig. 6 XPS wide-scan spectrum of the Pd@hTiO2 calcined at 773 K (inset: high-resolution spectrum of Pd3d). |
To investigate the catalytic characteristics of the new Pd@hTiO2 nanostructures, the reduction of 4-NP to 4-AP by NaBH4 at room temperature was chosen as a model reaction. As is well known, 4-NP can cause water pollution, which has attracted a wide public concern, while its derivative, 4-AP, is very important in synthetic organic chemistry and many industrial applications such as analgesic and anti-pyretic drugs, photographic developers, corrosion inhibitors, and so on.37 The color changes of the solution in the cuvette at different time could be observed visually, which was from bright yellow to colorless (Fig. 7) (Fig. 8).
The reduction process was monitored by UV-vis absorption spectroscopy. The reduction reaction did not proceed in the absence of Pd nanoparticles, which was evidenced by a constant absorption peak at 400 nm. The intensity of the characteristic absorption peak of 4-NP at 400 nm quickly decreases, and the characteristic absorption of 4-AP at around 300 nm also appears rapidly. This indicates that 4-NP is reduced to 4-AP quickly with discoloration of the solution from bright yellow to colorless. Fig. 9 showed the ln(Ct/C0) versus reaction time for the reduction of 4-NP over the Pd@hTiO2 composite catalyst used for the first time, where Ct and C0 are the concentrations of 4-NP at intervals and the initial stage, respectively. (At/A0) was defined as the ratio of maximal absorbance At of 4-NP at 400 nm at time t to its value.
![]() | ||
Fig. 9 The corresponding plot of ln(Ct/C0) against the reaction time of Pd@hTiO2 calcined at (B) 673 K, (C) 723 K, (D) 773 K, (E) 823 K, (F) 873 K and dried at (A) 298 K. |
A0 at t = 0, which could be directly interpreted as the ratio of the respective concentrations Ct/C0. Therefore, the reaction conversion at time t was calculated according to eqn (1):
Conversion (%) = (1 − Ct/C0) × 100 = (1 − At/A0) × 100 | (1) |
In this reaction system, the concentration of NaBH4 was much higher than that of 4-NP and could be considered as constant during the reaction, thus pseudo-first-order kinetics could be applied to evaluate the kinetic rate constant (k) of the current reaction [eqn (2)]:
dCt/dt = −kt, or ln(Ct/C0) = ln(At/A0) = −kt | (2) |
Meanwhile, the expected linear relationship between ln(Ct/C0) and reaction time (t) confirmed the pseudo-first-order kinetics. Since the rate constant k is influenced by the amount of the used catalyst, two kinds of the reaction rate constant were used to compare the catalytic activity.
The apparent rate constant k and k/(μmol Pd) (the reaction rate constant per the Pd content of the used catalyst) were calculated and showed in Table 1. The catalyst that calcined at 500 °C showed the highest catalytic activity under the present conditions (Table 1, entry 4), because the TiO2 have completely transformed into the anatase phase and the anatase shows higher catalytic activity than other crystalline phase.38 Compared with the literature,39–41 this synthesized catalyst Pd@hTiO2 exhibited comparable or much better catalytic activity for 4-NP reduction.
Entry | Calcination temperature T (K) | K (s−1) | k per Pd content(s−1 μmol Pd−1) |
---|---|---|---|
1 | 298 | 4.77 × 10−4 | 0.0825 |
2 | 673 | 1.32 × 10−3 | 0.229 |
3 | 723 | 3.23 × 10−3 | 0.56 |
4 | 773 | 4.77 × 10−3 | 0.826 |
5 | 823 | 4.02 × 10−3 | 0.696 |
6 | 873 | 9.03 × 10−4 | 0.156 |
For comparing the catalytic activity, the corresponding Pd/SiO2, hollow TiO2 and Pd/TiO2 catalysts were also prepared and evaluated using model reaction of the reduction of 4-NP to 4-AP with NaBH4 aqueous solution at room temperature as showed in Fig. 10. The Pd@hTiO2 hollow structure that calcined at 773 K exhibit better catalytic activity than the corresponding Pd/SiO2, hollow TiO2 and Pd/TiO2, due to its unique hollow structures with active NMNPs uniformly dispersed inside, which favor the quick diffusion of reactants, decreased depletion of catalytic active species, and effective contact of reactants with catalytic active species, eventually helping to improve their catalytic performance.
The separation and recovery of the Pd nanoparticles from the reaction system is one of the most important issues in the Pd-catalysis practical applications, the reusability of Pd@hTiO2 was further investigated. The catalyst was recycled in the reduction of 4-NP with NaBH4 and the catalytic process was achieved within 10 minutes under each cycle. As illustrated in Fig. 11, the catalyst was reused for at least eight times without a significant loss of activity.
The catalyst Pd@hTiO2 demonstrating excellent activity and reusability in the reduction of 4-NP was attributed to several factors. Firstly, the catalysts with a thin mesoporous shell and the desired large inner space favors reactant diffusion and exposure of active sites, leading to the improvement of catalytic activity. Secondly, the election affinity and electron transfer of neutral Pd particles largely depend on their size and small Pd nanoparticles encapsulated in the support behave more like individual Pd atoms, which are highly active. In our case, the size of Pd nanoparticles was mostly smaller than 6 nm; besides, they were excellent accessibility, good dispersion and high stability, so the catalyst exhibited very efficient. Moreover, the hollow TiO2 shell not only avoided Pd nanoparticles leaching from the support, but also improved the homogenous distribution of Pd nanoparticles on the support and their catalytic activity for the reduction.
This journal is © The Royal Society of Chemistry 2014 |