Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Polymer-templated synthesis of hollow Pd–CeO2 nanocomposite spheres and their catalytic activity and thermal stability

Chenhao Du , Yun Guo , Yanglong Guo , Xue-qing Gong and Guanzhong Lu *
Key Laboratory for Advanced Materials, Research Institute of Industrial Catalysis, East China University of Science and Technology, Shanghai 200237, P. R. China. E-mail: gzhlu@ecust.edu.cn; Fax: +86-21-64252923

Received 7th July 2015 , Accepted 8th October 2015

First published on 8th October 2015


Abstract

Developing methods for the fabrication of active and thermally stable noble metal/metal oxide nanomaterials are very important for catalysis and material fields. Herein, we used a polymer-template synthesis approach to prepare hollow Pd–CeO2 nanocomposite spheres (NCSs) with Pd nanoparticles evenly distributed inside the CeO2 shell, in which the aggregation of Pd nanoparticles can be well inhibited with the help of the protection of CeO2 nanocrystallites even after being calcined at 700 °C. The Pd nanoparticles are partially buried in the CeO2 shell and the surface Pd species are highly ionic, which is caused by the electron exchange at the Pd–CeO2 interface during calcination. This hollow structure Pd–CeO2 nanocatalyst shows excellent catalytic activity and stability in the aqueous selective reduction of 4-nitrophenol and gaseous CO oxidation. For the selective reduction of 4-nitrophenol, the reaction rate of this h-Pd–CeO2 NCS catalyst compared to those of the supported Pd/CeO2 and physically mixed Pd + CeO2 catalysts is almost 14 times and 5 times faster, respectively. For the CO oxidation, the larger Pd–CeO2 interface in h-Pd–CeO2 NCSs could facilitate the reaction between the adsorbed CO and O2, thus showing better low temperature activity. This paper emphasizes the advantages of a core–shell hollow structure and provides a new way for obtaining novel functional nanocatalysts.


1. Introduction

Noble metal nanoparticles (NPs) have been extensively researched in the field of catalysis in recent years, due to their highly effective catalytic performance.1–3 It is well known that the catalytic activity of noble metal NPs is closely dependent on their composition, size, shape and crystallinity.4–7 However, owing to their high surface area and surface energy, noble metal NPs tend to aggregate to larger particles to minimize their surface energy, especially during the catalytic reaction at high temperature, which could easily lead to their deactivation or loss of the catalytic activity.8,9 To solve these problems, isolation of noble metal NPs in small cavities is a reasonable solution, because it can efficiently prevent particle migration and its deactivation. Besides, this isolation method can also reduce the leaching of the active components from the catalyst particles in the chemical process, which is another important aspect in the design of new nanocatalysts.10,11

To isolate noble metal NPs, porous inert oxide coatings and core–shell hollow nanostructures are the ideal solution approaches, and both have been widely researched recently.12–20 Lu et al. designed a coking- and sintering-resistant palladium catalyst by an atomic layer deposition method, which can enhance the catalytic activity and stability of the Pd/Al2O3 catalyst for the oxidative dehydrogenation of ethane.18 Zhang et al. reported the graphene nanosheet-supported ultrafine metal nanoparticles encapsulated by thin mesoporous SiO2 layers, which exhibited remarkable high-temperature stability.19 Similarly, porous silica-coated Pt NPs on carbon nanotubes possessed good thermal stability for the hydrogenation of nitrobenzene.20 However, Al2O3 or SiO2 are unreactive components and only play a dispersion role, and the synergistic effect between noble metal NPs and the Al2O3 or SiO2 layer is quite weak. From a viewpoint of heterogeneous catalysis, especially for metal/metal oxide catalysts, the interaction between the metal nanoparticles and the reactive support,21–23 such as CeO2 and TiO2, is quite vital for improving the reactivity and stability of the catalysts.

As one of the most commonly used reducible oxides, ceria possesses a high oxygen storage capacity and redox property, and plays an important role in the field of catalysis and has been researched extensively.24–27 Although various methods for the fabrication of hollow ceria spheres have been reported,28–30 the research on CeO2-containing nanocomposite hollow spheres, especially composites with noble metal nanoparticles, is relatively rare. For example, Liu and co-workers designed a nanoreactor with noble metal NPs decorated in the inner walls of the CeO2 shells and showed that it could enhance the catalytic activity and stability.31 Xu et al. synthesized hollow Pd@CeO2 spheres by a core–shell protocol, which exhibited excellent reusable and higher catalytic performance for the thermocatalytic and photocatalytic selective reduction of aromatic nitro compounds.32 However, these research studies were only involved with the catalytic processes under ambient conditions, and the high-temperature stability of these catalysts was not demonstrated, which is very important for the gas–solid phase catalytic reaction at a higher temperature. Therefore it is quite necessary to develop novel nanocatalysts with good activity and stability for both gas- and liquid-phase reactions.

Herein, we reported the fabrication of hollow Pd–CeO2 nano-composite spheres (NCSs) by a facile hard-template method in an aqueous phase, followed by the elimination of the polymer templates by simple calcination. In these hollow Pd–CeO2 NCSs, the aggregation of Pd nanoparticles can be efficiently avoided by the protection of the CeO2 crystal. The high dispersity of Pd species can be well maintained even after calcination at 700 °C. The synthesized hollow Pd–CeO2 NCSs exhibit high catalytic activity for the selective reduction of aromatic nitro compounds under ambient conditions and CO oxidation. It is most important that hollow Pd–CeO2 NCSs possess high thermal stability and reusability compared to conventionally supported Pd/CeO2 catalysts.

2. Experimental section

2.1. Preparation of samples

Monodispersed resorcinol-formaldehyde (RF) polymer spheres. RF polymer spheres were synthesized.33 Typically, 0.1 mL of ammonia aqueous solution (NH4OH, 25 wt%) was mixed with a solution containing 8 mL of ethanol and 20 mL of de-ionized water. After stirring for 1 h, 0.1 g of resorcinol was added to this mixed solution and it was continually stirred for 30 min. Then 0.14 mL of formaldehyde solution was added and stirred for 24 h at 30 °C, and subsequently this synthesized solution was transferred to a Teflon-lined stainless autoclave and treated at 100 °C for 24 h. The as-synthesized product was collected by centrifugation and washed with de-ionized water twice. Finally, the solid product was dried in air at 100 °C for 24 h.
PVP-protected Pd nanoparticles. PVP-protected Pd nanoparticles were prepared by the alcohol reduction method.34 H2PdCl4 aqueous solution (2 mM) was prepared by mixing 106.4 mg of PdCl2, 6 mL of 0.2 M HCl and 294 mL of ultrapure water. A mixed solution of 15 mL H2PdCl2 aqueous solution, 21 mL de-ionized water, 14 mL ethanol and 66.7 mg PVP was refluxed in a 100 mL flask at 90 °C for 3 h. The obtained dark-brown PVP-protected Pd colloidal solution was stable at room temperature for months.
Hollow Pd–CeO2 nano-composites (h-Pd–CeO2). 25 mg of as-prepared RF spheres was dispersed in 20 mL of de-ionized water under ultra-sonication to obtain a homogeneous dispersion, followed by adding the Pd colloidal solution (10–15 μmol). This mixture was stirred for 4–6 h to allow Pd NPs fully deposited on the surface of RF spheres to form an RF@Pd structure. Then a mixed solution of 0.15 g of CeCl3, 0.085 g of hexamethylenetetramine (HTMA) and 20 mL of de-ionized water was added into the solution containing RF@Pd particles mentioned-above, and it was continually stirred for 15 min. Then this synthesized solution was transferred into a 100 mL Teflon-lined stainless autoclave and treated at 100 °C for 6 h. The obtained products were separated by centrifugation, washed with de-ionized water 3 times to remove unreacted residues, and then dried at 60 °C overnight and calcined at 500–700 °C for 2 h in a muffle furnace.
The Pd/CeO2 catalyst. The supported Pd/CeO2 catalyst was prepared by the conventional impregnation method (Pd/CeO2-imp), and the Pd loading was ∼2.0 wt%. Commercial CeO2 powder was provided by Aladdin Company. The as-synthesized Pd/CeO2 powder was calcined at 500 °C for 2 h.
The physical mixture of Pd + CeO2. The 2.0 wt% Pd physically mixed sample was prepared by an evaporation of the colloidal solution including CeO2 nanoparticles and Pd nanoparticles and followed by calcination in air at 500 °C for 2 h.

2.2. Catalytic activity testing

The selective reduction of 4-nitrophenol. Catalytic performance of the catalyst for the selective reduction of 4-nitrophenol was tested as follows. Typically, the catalyst (1.5 mg) was added into 40 mL of ultrapure water in a 100 mL beaker to form a homogeneous dispersion under ultrasonication for 10 min, and then 0.5 mL of freshly prepared 0.5 M NaBH4 aqueous solution was added. After stirring for a few seconds at room temperature, 4-nitrophenol (0.012 M, 0.25 mL) was added into the reactor at room temperature (25 °C). The reaction process was monitored by the direct color change of the solution. At different reaction times, 1.5 mL of the solution was withdrawn with a syringe and filtered using a 0.22 μm Millipore film, and the reaction solution was analyzed by UV-Vis spectroscopy.
Recycling of the catalyst. After the reaction ended, the catalyst was separated by centrifugation at 5000 rpm for 10 min, and the catalyst was dried at 60 °C. Then this catalyst was directly reused under the same reaction conditions.
CO oxidation. The catalytic activity of the sample for CO oxidation was evaluated at atmospheric pressure in a quartz tubular reactor. 30 mg of the catalyst was mixed with 200 mg of silica (40–60 mesh). The total flow rate of the feed gas of 1% CO–20% O2/N2 was 50 mL min−1. The catalyst was pretreated in N2 at 150 °C for 1 h and then the temperature was raised at a rate of 2 °C min−1. The gas composition was monitored online using a gas chromatograph (GC 2060 system) with a flame ionization detector (FID) and methanation reactor. The conversion of CO was calculated on the basis of the change of CO concentrations in the inlet and outlet gases.

2.3. Characterization of samples

Elemental analysis of the sample was done by inductively coupled-plasma atomic emission spectroscopy (ICP-AES) on a Varian 710ES instrument (Varian Co. Palo Alto, US). Powder X-ray diffraction (XRD) patterns were recorded on a PANalytical PW 3040/60 X'Pert Pro powder diffractometer with CuKα radiation, which was operated at 40 kV and 40 mA and the scanning speed was 0.5° min−1. Thermogravimetric (TG) analysis was conducted on a Pyris diamond thermal analyzer at a heating rate of 10 °C min−1 from room temperature to 800 °C. The surface areas of samples were measured by N2 adsorption/desorption isotherms at −196 °C on a Micromeritics ASAP 2020 instrument, and calculated by the Brunauer–Emmett–Teller (BET) method.

Transmission electron microscopy (TEM) images and energy dispersive X-ray spectroscopy (EDX) spectra were obtained with JEM-1400 & JEM-2100 transmission electron microscopes, respectively. The sample was ultrasonically suspended in the ethanol solvent, and one or two drops of this slurry was deposited on a copper grid. The liquid phase was evaporated before the grid was loaded into the microscope. Scanning electron microscopy (SEM) images were taken on a Hitachi S-3400 scanning electron microscope operated at 15 kV and using secondary electrons to form the images. The sample was coated with a thin layer of gold before testing.

X-ray photoelectron spectroscopy (XPS) spectra of the samples were obtained on a Kratos Axis Ultra DLD photoelectron spectrometer equipped with AlKα (1486.6 eV) radiation as the excitation source. All binding energies (BE) were determined with respect to the C 1s line (284.6 eV) originating from adventitious carbon.

3. Results and discussion

3.1. Preparation of hollow Pd–CeO2 nanocomposite spheres and their properties

The synthesis process of hollow Pd–CeO2 nanocomposite spheres (NCSs) is illustrated in Scheme 1. The pre-synthesized Pd colloidal nanoparticles (NPs) with PVP as the stabilizer are 2.5–3 nm (Fig. S1), and could be loaded on the surface of monodispersed resin polymer spheres by the π–π interaction between the polymer template and PVP (Fig. S2).32,35 Basically, almost all of the Pd colloidal nanoparticles are successfully deposited on the surface of the polymer, which can be verified by the relative colorless transparent supernatant obtained after centrifugation of the RF@Pd suspension. As is well known, hexamethylenetetramine (HTMA) can hydrolyze in water to give NH3,36 while the basicity of the solution is slowly changed during the reaction, which ensures the slow hydrolysis of CeCl3 and condensation of Ce3+ ions. Meanwhile, with the coordination effect of PVP and surface functional groups on the polymer surface, CeO2 could be well deposited on the template surface to form an RF@Pd@CeO2 core–shell structure. Followed by calcination in air, the RF template and PVP are removed and the hollow CeO2 shell forms simultaneously.
image file: c5ta05092h-s1.tif
Scheme 1 The formation process of hollow Pd–CeO2 nano-composite spheres.

As shown in Fig. 1, the diameter of mono-dispersed RF resin polymer spheres is ∼500 nm, and the well-structured hollow spheres and the inner hollow structure can be clearly observed as indicated with the arrows (Fig. 1b), which is further confirmed by the TEM image (Fig. 1c). The CeO2 shells composed of sub-10 nm CeO2 nanocrystallites with the thickness of ∼35 nm, and the cavities in the shell can guarantee the reagents diffusion and contact in the inner and outer surfaces of the shell during the catalytic reaction. No bare Pd NPs are spotted on the surface of hollow spheres and almost all of the Pd NPs are embedded in the CeO2 nanocrystals after calcination, forming Pd–CeO2 interfaces as indicated by the white arrow (Fig. 1d). The TG analysis shows that the hollow Pd–CeO2 NCSs exhibit hardly weight loss at 100–800 °C, which demonstrates the complete elimination of resin polymer spheres during the calcination (Fig. S3).


image file: c5ta05092h-f1.tif
Fig. 1 SEM images of (a) resin polymer template and (b) SEM, (c) TEM, and (d) HR-TEM of hollow Pd–CeO2 NCSs calcined at 773 K.

Owing to the high electron density of polycrystalline CeO2 and the hollow shell structure, it is difficult to identify Pd NPs located on the ceria shell. To visualize the spatial distribution of Pd species, we used the EDX technique to analyze the individual hollow Pd–CeO2 sphere, and the results are shown in Fig. 2. It can be seen that the elements of Pd, Ce and O spread evenly on the whole sphere, which proves that no Pd aggregation appeared on the CeO2 shell. Besides, the composition line-scan profiles across a single h-Pd–CeO2 nanocomposite sphere were obtained by means of SEM-EDX and are shown in Fig. S4, which indicates that all the elements (Ce, O and Pd) are distributed throughout the sphere.


image file: c5ta05092h-f2.tif
Fig. 2 (a) TEM images, (b) EDX spectra of individual hollow Pd–CeO2 NCS calcined at 500 °C, and its elemental mappings of Ce, O and Pd.

As shown in the XRD spectra (Fig. 3d), the hollow Pd–CeO2 NCSs are the polycrystalline structures, the diffraction peaks result from CeO2, and no diffraction peaks of Pd or Pd oxide could be observed, probably due to the relatively low content or high dispersion of Pd species in the sample. With an increase in the calcination temperature from 500 to 700 °C, the hollow structure is well maintained and no Pd NPs or aggregated Pd particles were observed in its TEM images (Fig. 3a–c); the diffraction peaks become slightly narrower, while the size of CeO2 crystallites increases from 5–10 to 10–15 nm and its BET surface area decreases from 59.3 to 33.9 m2 g−1 (Table S1). As shown in Fig. 3b, the shell is more rugged and this hard-template method can effectively inhibit the growth of the crystal size at a higher temperature, whereas isolated CeO2 nanoparticles would severely aggregate as indicated by the white arrow in Fig. 3a. Compared to the supported Pd/CeO2 catalyst (9.1 m2 g−1), the hollow Pd–CeO2 NCSs have a larger surface area (33.9 m2 g−1) after calcination at 700 °C (Table S1).


image file: c5ta05092h-f3.tif
Fig. 3 (a–c) TEM images of hollow Pd–CeO2 NCSs calcined at 700 °C and (d) XRD profiles of CeO2 and h-Pd–CeO2 NCSs.

The hollow Pd–CeO2 NCSs have also been investigated by the XPS technique, and the results are shown in Fig. 4. In the Pd 3d5/2 and 3d3/2 XPS spectra, two peaks are located at ∼337.6 eV and 342.9 eV, respectively. In general, the BE (Binding Energy) values of Pd 3d5/2 in PdO species are in the range of 336–337.2 eV, so that the Pd in the hollow Pd–CeO2 NCS should not be PdO. It was reported that, the fact that the XPS peaks of Pd2+ shifted to a higher BE value in Pd/CeO2, could be explained by the formation of a solid solution-like PdxCe1−xO2 structure, in which the Pd2+ ions replaced the Ce4+ ions in the CeO2 crystalline lattice.37–40 This result demonstrates that the Pd NPs are embedded inside the CeO2 nanocrystals, but not completely buried in the ceria shell. A fraction of Pd atoms located at the Pd–CeO2 interface are exposed and the electron exchange between the Pd atoms and the CeO2 matrix leads to the formation of ionic Pd species. Besides, previous research studies showed that the Pd2+ ions could not be completely reduced and a minor Pd2+ component might exist on the surface of Pd NPs.41,42 Thus, on the basis of XPS data and the Pd–CeO2 interface observed in Fig. 1d, we can consider that the chemical state of surface Pd in the hollow Pd–CeO2 NCSs is mainly ionic Pd2+ but not in the form of PdO, which probably results from the strong metal-support interaction caused by the treatment at high temperature, and forming the solid solution-like PdxCe1−xO2 structure.


image file: c5ta05092h-f4.tif
Fig. 4 XPS spectra of (a) Pd 3d, (b) Ce 3d and (c) O 1s for hollow Pd–CeO2 NCSs.

For the Pd–CeO2 NCS, after being calcined at 700 °C, its valence state of Pd species is unchanged, but the intensities of peaks are stronger than those of the sample calcined at 500 °C (Fig. S5), which indicates that the surface concentration of Pd species increased after calcination at higher temperature (Table 1). This phenomenon is caused probably by two reasons. (1) Small Pd nanoparticles would undergo a structural evolution and the Pd atoms could dissociate from Pd nanoparticles in the process of calcination at 800 °C, but this situation would not occur at 500 °C.43 Thus, in the h-Pd–CeO2 NCSs, some Pd atoms may dissociate and be incorporated into the CeO2 nanocrystallites during the calcination at 700 °C, which causes an increase in the concentration of surface Pd species, though the Pd NPs remain in their original places. (2) After calcination at 700 °C, the growth of CeO2 nanocrystallites leads to the surface shrinkage of the ceria hollow sphere and a decrease in its surface area (from 59.3 to 33.9 m2 g−1, Table S1). Thus the Pd NP number per square meter would increase, resulting in the higher surface concentration of Pd species for the sample calcined at 700 °C than that of the sample calcined at 500 °C.

Table 1 Composition of hollow Pd–CeO2 nanocomposite spheres and supported Pd/CeO2 catalysts
Catalyst Pd (wt%) in solution XPS ICP
Pd (wt%) Pd/Ce (atom) Pd (wt%) Pd/Ce (atom)
h-Pd–CeO2-773 K 1.5 2.00 0.032 2.15 0.035
h-Pd–CeO2-873 K 1.5 2.76 0.045 1.95 0.032
h-Pd–CeO2-973 K 1.5 4.30 0.065 1.95 0.032
Supported Pd/CeO2-773 K 2.0 6.00 0.097 2.00 0.032


Fig. 4b shows the Ce 3d XPS spectrum of hollow Pd–CeO2 NCSs, in which there are the V (BE = 882.5 eV) and U (901.0 eV) peaks, four (*) satellite peaks (BE = 888.8, 898.4, 907.4, and 916.7 eV) and the V′ (885.2 eV) and U′ (903.5 eV) peaks.44 These results show that the main valence state of Ce in the hollow Pd–CeO2 NCSs is +4 (V, U and four satellite peaks) with a small amount of Ce3+ (V′ and U′ peaks). In the O 1s spectra, the peaks at around 529.2 eV and 531.8 eV are ascribed to the lattice oxygen in the hollow Pd–CeO2 NCSs and their surface oxygen species, respectively.

The amount of Pd obtained by ICP-AES in hollow Pd–CeO2 NCSs was ∼2.15%, and obviously higher than the theoretic value of Pd (1.5 wt%) in the synthesized solution (Table 1), indicating that not all Ce3+ ions in the solution were converted to CeO2 in the synthesis process. For the sample calcined at 500 °C, the fact that the Pd content obtained by XPS is similar to that by ICP confirms that Pd evenly distributes in the sample. With an increase in the calcination temperature from 500 to 700 °C, the surface concentration of Pd increases from 2.0 to 4.3 wt% (by XPS), indicating that high temperature enriched Pd on the surface. For the supported Pd/CeO2 catalyst, the surface concentration (6.0 wt%) of Pd is much higher than that of the Pd loading of 2.0 wt.%.

The above-mentioned results show that by using resin polymer spheres as the hard template and with the help of the strong interaction between Pd and CeO2, the sandwich-like RF polymer@Pd NP@CeO2 core–shell structure materials can be efficiently converted to hollow Pd–CeO2 nanocomposite spheres after simple calcination. CeO2 nanocrystals are closely stacked around the well-dispersed Pd NPs on the shell, thus making the surface valence state of Pd highly ionic, rather than metallic. In the high-temperature calcination process, no Pd aggregation was formed, which demonstrates the efficient shielding effect of shell CeO2 crystals at higher temperatures.

3.2. Catalytic performance of hollow Pd–CeO2 nanocomposite spheres

The catalytic activity of hollow Pd–CeO2 NCSs was evaluated for the reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP) by an excess of NaBH4 under ambient conditions and the oxidation of CO, which have been considered as the typical model reactions to investigate the catalytic activities of noble metal catalysts under aqueous or gaseous conditions.

It is well known that the 4-NP solution exhibits a strong absorption peak at 317 nm under a neutral or acid environment. However, with the excess of NaBH4 in the solution, the 4-NP ions quickly become the dominant species due to the alkalinity of the solution and relatively higher acidity of 4-NP compared to water, thus leading to the H+ dissociation from 4-NP to react with NaBH4, leaving the 4-NP anions. This is consistent with the peak shifting to 400 nm.

As shown in Fig. 5a, when a small amount of the catalyst (1.5 mg) was added to the solution, the absorption peak at 400 nm decreased quickly and a new peak at 295 nm appeared simultaneously, which indicates the reduction of 4-NP and formation of the 4-AP product. After ∼2 min of the reaction, the absorption intensity at 295 nm reached a maximum and the absorption peak at 400 nm entirely diminished. Meanwhile, the light yellow solution completely faded and became colorless.


image file: c5ta05092h-f5.tif
Fig. 5 (a) UV-Vis absorption spectra in the reduction of 4-NP to form 4-AP by an excess of NaBH4 under ambient conditions over the hollow Pd–CeO2 NCSs and (b and c) the concentration (C/C0) of 4-NP as a function of the reaction time over (b) the fresh sample and (c) one treated with NaBH4 and then calcined at 650 °C for 4 h in N2.

In the selective reduction of 4-NP to 4-AP, Pd acted as the primary active site to catalyze the reduction reaction by facilitating the electron transfer from BH4 to the reactant, thus leading to the effective reduction of the nitro group.45 In order to clarify the chemical valence of Pd active sites in the reaction process, we performed one experiment without the 4-NP reactant, but only the excess NaBH4, in which other experiment conditions were maintained. The results show that when the hollow Pd–CeO2 NCS catalyst was added to the solution under stirring for a few seconds, the color of this catalyst changed from brown to gray, which is caused by the reduction of Pd2+ to metallic Pd.

As shown in Fig. 6a, the XPS spectrum of the in situ reduced sample exhibits only two peaks at about 335.3 eV and 340.2 eV, indicating the formation of metallic Pd species by the reduction of partial Pd2+ ions. Therefore, we presume that the high catalytic activity of the hollow Pd–CeO2 NCSs is attributed to these metallic Pd clusters on the surface of CeO2 crystallites. Besides, the core–shell structural features can enhance the synergistic effect between the Pd NPs and the CeO2 NPs and speed up the charge transfer rate, resulting in an increase in the catalytic activity.46 To test the thermal-resistant stability of these h-Pd–CeO2 NCSs, we further treated the reduced sample at 650 °C in N2 for 4 h, and then its catalytic activity for the 4-NP reduction was tested under the same conditions. Compared to the catalytic activity of the fresh catalyst (Fig. 5b), after the sample was treated with NaBH4 and then calcined at 650 °C in N2 for 4 h, its catalytic activity remained unchanged (Fig. 5b), that is, the conversion of 4-NP to 4-AP could be completed in ∼2 min. As shown in the TEM images of this reduced and thermally treated sample, its structure and shape could be well maintained (Fig. S6), indicating the good chemical and thermal stability of this Pd–CeO2 NSC catalyst.


image file: c5ta05092h-f6.tif
Fig. 6 Pd 3d XPS spectra of the hollow Pd–CeO2 nanocomposite (a) calcined at 500 °C, (b) after being treated with NaBH4, and (c) after CO oxidation.

For a comparison, the catalytic activity of the Pd/CeO2-imp catalyst prepared by the impregnation method and the physical mixture of the Pd + CeO2 catalyst were also tested under the above-mentioned experiment conditions. The 4-NP conversion can be directly seen from the ln(C/C0) (C is the concentration of 4-NP at the reaction time (t) and C0 is the initial concentration of 4-NP) versus the reaction time (t), which is shown in Fig. 7. Evidently, the two reference catalysts possess inferior catalytic performance compared to that of the h-Pd–CeO2 NCSs. For the reduction of 4-NP with an excess of NaBH4, CeO2 exhibits no catalytic activity. Therefore, the catalytic performances of these samples are strongly affected by the size of Pd nanoparticles and their states in the reaction system. As shown in the TEM image of Fig. S7, for the physical mixture of the Pd + CeO2 catalyst calcined at 500 °C without the purposeful design of the core–shell hollow structure, agglomeration of Pd nanoparticles occurs, thus, the available surface area of active Pd would decrease obviously, resulting in the reduction of its catalytic activity. As for the Pd/CeO2-imp catalyst, though no bare Pd particles are seen in the TEM images, the Pd species are not evenly distributed on the whole CeO2 support (Fig. S8). The XPS analysis indeed exhibits a peak at 336.6 eV, which could be attributed to the formation of PdO species (Fig. S9). Besides, most of the Pd/CeO2-imp particles are severely sintered (Fig. S8) and the sample cannot be well dispersed in 4-NP solution. The catalyst powder would quickly deposit at the bottom of the reaction vessel during the catalytic reaction, which limits the efficient contact between the reactants and the Pd sites in the solution, resulting in a remarkable reduction of its catalytic activity.


image file: c5ta05092h-f7.tif
Fig. 7 ln(C/C0) as a function of the reaction time (t) over hollow Pd–CeO2 NCS, Pd/CeO2-imp and Pd + CeO2 physical mixture catalysts for the reduction of 4-NP to 4-AP by an excess of NaBH4 under ambient conditions.

The above results show that the structures of the catalyst and active Pd states indeed affect their catalytic performances. Among the three catalysts, the Pd NPs in hollow Pd–CeO2 NCSs could be well protected by CeO2 nanocrystals in the shell, thereby preserving their high catalytic activity. And the porous CeO2 shell and the hollow structure could favor the diffusion of the reactants and products, and increase the efficient contact between the reactants and the Pd active sites, leading to an enhancement of the catalytic performance. TOF of the h-Pd–CeO2 NCSs was calculated and reached 335 h−1, which shows good catalytic activity compared to the catalysts reported (Table 2).

Table 2 TOF values over Pd-based nanocatalysts for the reduction of 4-NP
Catalyst TOF (h−1) Ref.
Pd/PPy/TiO2 32.3 47
Pd-spherical polyelectrolyte brushes 159 48
@Pd/CeO2 1068 31
Core–shell Pd@h-CeO2 300 32
Pd/polymer/CNT 96 49
h-Pd–CeO2 335 This work


Fig. 8a shows the relationship of ln(C/C0) versus the reaction time (t) over the h-Pd–CeO2 NSC catalyst with different amounts. An almost linear evolution between ln(C/C0) and time (t) indicates a first-order reaction for the concentration of 4-NP. When the catalyst amount is 1.5 mg (0.038 mgcat mL−1), the reduction rate of 4-NP is close to that over 2.2 mg of the catalyst (0.055 mgcat mL−1); decreasing the catalyst concentration would dramatically decrease the catalytic reaction rate, and further increasing the catalyst concentration hardly affects the catalytic reaction rate obviously, so the appropriate catalyst amount is about 0.038 mgcat mL−1. As shown in Fig. 7 and 8, the reduction of 4-NP to 4-AP by an excess of NaBH4 under ambient conditions is of the first order reaction for 4-NP concentration, hence, the rate constant k can be calculated from the rate equation ln(C/C0) = −kt.


image file: c5ta05092h-f8.tif
Fig. 8 Plots of ln(C/C0) vs. the reaction time (t) over the hollow Pd–CeO2 NCS catalyst (a) with 0.8–2.2 mg (calcined at 500 °C), (b) with 1.5 mg (calcined at 500–700 °C), and (c) the conversion of 4-NP (4 min) vs. recycling times over the h-Pd–CeO2 NCSs calcined at 500 °C and Pd-on-resin polymer sphere (RPS) catalysts (1.5 mg) for the reduction of 4-NP to 4-AP by an excess of NaBH4 under ambient conditions.

Fig. 8b shows the effect of the calcination temperature on the catalytic activity of the h-Pd–CeO2 NCS catalyst. As shown in the plots of ln(C/C0) versus the reaction time, the catalytic activity of this catalyst increases slightly after being calcined at 700 °C, this is because the surface concentration of Pd species was increased after calcination at higher temperature (Table 1). It is interesting to note that the Pd nanoparticles in the h-Pd–CeO2 NCSs did not aggregate even after being calcined at 700 °C and their catalytic activity could be well maintained.

Fig. 8c shows the effect of recycling use on the catalytic activity of h-Pd–CeO2 NCSs. With the increase of the recycling times, the conversion of 4-NP dropped gradually, and the conversion of 4-NP reached 84% after 6 times of recycling use. Notice that the h-Pd–CeO2 NCS catalyst was reclaimed from the solution by centrifugation before reuse, in which some nanocomposites cannot be reclaimed due to their excellent dispersion in water, and only about 50% of h-Pd–CeO2 NCSs could be obtained after 6 cycles. When this recovered catalyst was replenished to 1.5 mg by adding a fresh catalyst, its catalytic conversion can reach 100% once again in the 7th and 8th run. The TEM image (Fig. S10) shows that the morphology of the Pd–CeO2 NCSs after 5 times of recycling use could be well maintained, though some hollow spheres are broken, and the ICP analysis shows that their Pd amount is about 1.9 wt%, that is to say, almost no leaching of Pd species occurs after the recycling use of 5 times. These phenomena verify that it is the weight loss during the separation process rather than degradation of the catalyst, which is the reason for the decrease of the catalytic conversion with increasing recycling times.

Moreover, we prepared the catalyst of Pd NPs supported on resin polymer spheres (denoted as Pd-on RPSs), and its catalytic performance for the reduction of 4-NP to 4-AP by an excess of NaBH4 under ambient conditions was tested. The results (Fig. 8c) show that after being repeatedly used 3 times, the catalytic activity of Pd-on RPSs decreases dramatically, probably due to the detachment of Pd NPs from the resin polymer spheres.

The Pd supported on CeO2 is one of the most prevalent catalysts in the exhaust emission control, so the catalytic performance of the hollow Pd–CeO2 NCS catalyst for CO oxidation was also evaluated. Typically, 30 mg of h-Pd–CeO2 NCSs was loaded in a U-type quartz tubular reactor, the gas mixture of 1% CO–20% O2/N2 was used as the feed gas, and the space velocity was 100[thin space (1/6-em)]000 mL gcat−1 h−1.

As shown in Fig. 9a, the h-Pd–CeO2 NCSs possess good catalytic activity for CO oxidation in a high space velocity of 100[thin space (1/6-em)]000 mL gcat−1 h−1, and the calcination temperature affects their catalytic activity, for instance, the sample calcined at 700 °C exhibits higher catalytic activity than the sample calcined at 500 or 600 °C. This situation is consistent with the catalytic reduction of 4-NP over this h-Pd–CeO2 NCS catalyst, which should be attributed to the higher surface concentration of Pd species on the sample calcined at 700 °C (Table 1).


image file: c5ta05092h-f9.tif
Fig. 9 CO oxidation over (a) h-Pd–CeO2 NCSs calcined at 500–700 °C, and (b) h-Pd–CeO2 NCSs, Pd/CeO2-im and the physical mixture of Pd + CeO2 catalysts (space velocity of 100[thin space (1/6-em)]000 mL gcat−1 h−1).

A cycling test was conducted to study the stability of the h-Pd–CeO2 NCS catalyst at different temperatures, in which the reaction was tested at 600 °C for 30 min, and then the reaction temperature was cooled to 130 °C to evaluate the CO oxidation for 30 min, and then repeated the cycling operation mentioned above. As seen in Fig. 10, after the alternate reaction at 600 °C and 130 °C three times, the h-Pd–CeO2 NCSs still maintain around 80% CO conversion at 130 °C. As shown in its TEM image (Fig. S11), its original structure can be well retained after being used alternately at 600 °C and 130 °C three times, indicating its high and stable catalytic performance at higher reaction temperature.


image file: c5ta05092h-f10.tif
Fig. 10 CO oxidation over the h-Pd–CeO2 NCS catalyst at alternate changes of 600 °C and 130 °C (space velocity of 100[thin space (1/6-em)]000 mL gcat−1 h−1).

The catalytic activities of the three Pd catalysts for CO oxidation are shown in Fig. 9b. The activity of the h-Pd–CeO2 NCS catalyst is much higher than those of Pd/CeO2-imp and the physical mixture of Pd + CeO2 catalysts. Based on the differences in the textural and physicochemical properties of the three catalysts, the higher activity of the h-Pd–CeO2 NCS catalyst for the CO oxidation might be closely related to the size of Pd NPs, the surface area of the catalyst and the interface between Pd and CeO2.50–52 In the h-Pd–CeO2 NCSs, the Pd NPs are uniformly scattered in the ceria shell and no aggregation occurs after calcination at high temperature. Besides, a comparatively higher surface area of h-Pd–CeO2 NCSs would be beneficial to the improvement of its catalytic activity.

The larger Pd particles detached from the nanosized CeO2 can be observed in the mixture of the Pd + CeO2 catalyst, due to sintering at high temperature. For the Pd/CeO2-imp catalyst, the surface area and TEM testing have confirmed a low surface area and uneven distribution of PdO particles on the sample. These disadvantages result in their poor catalytic activities at low temperature (below 120 °C).

In the h-Pd–CeO2 NCSs, the electron exchange occurs at the interface between Pd NPs and nanosized CeO2, and the surface Pd species is ionic, as evidenced in the Pd 3d XPS profile. To maintain the charge neutrality, the oxygen vacancies should be created at the CeO2 near Pd NPs, leading to the partial reduction of CeO2 to CeO2−x, because of the lower valence state of Pd2+ compared to that of the Ce4+. Therefore, we can conclude that the Pd2+ ions at the interface are the active sites for CO adsorption, and the neighboring oxygen vacancies can facilitate the adsorption and activation of gaseous O2 to form peroxide (O22−) species,53 thus the CO oxidation could occur at lower temperature. And the O 1s XPS spectra in Fig. S12 clearly shows that the h-Pd–CeO2 NCS has a higher ratio of surface chemisorbed oxygen (531.8 eV) to lattice oxygen (529.2 eV) than that of the supported Pd/CeO2-imp catalyst. More chemisorbed oxygen species would be beneficial to the reaction with the absorbed CO molecules on nearby interfacial Pd2+ species, which could well improve the catalytic activity for the CO oxidation. After this catalyst was used in the CO oxidation, the valence state of Pd is mostly still in a form of ionic state (Fig. 6c). Two new tiny peaks at EB = 335.3 eV and 340.5 eV in its Pd 3d XPS spectrum showed the formation of metallic Pd0 species, which might be caused by the reduction of CO.39

4. Conclusions

In summary, we have successfully prepared hollow Pd–CeO2 nanocomposite spheres (NCSs) via a polymer-template method. The Pd nanoparticles (NPs) in the feedstock solution are closely stacked in the surrounding CeO2 nanocrystals after eliminating the template core, that is to say, the Pd NPs are highly dispersed inside the CeO2 shell in these prepared hollow Pd–CeO2 NCS materials, and would not aggregate even after being calcined at 700 °C and their structure and catalytic performance could be well maintained. The confined shield effect of the CeO2 shell can efficiently inhibit the aggregation of Pd NPs and prevents the leaching of Pd species in the solution reaction. Besides, the surface Pd species in the hollow Pd–CeO2 NCSs is mainly ionic, which is caused by the electron exchange between Pd atoms and their surrounding CeO2 nanocrystals in the calcination process. Therefore, the synergistic catalytic effect between Pd NPs and CeO2 can be enhanced in this core–shell like structure.

For the aqueous-phase selective reduction of 4-nitrophenol to 4-aminophenol and gas phase low temperature oxidation of CO, this hollow Pd–CeO2 NCS material shows excellent catalytic activity and reusability, compared to that of the supported Pd/CeO2 catalyst and the physically mixed catalyst of Pd + CeO2. The synthetic strategy herein might provide a novel guidance for the designing and fabrication of mutli-core@oxide shell core–shell nanocomposites that can act as effective multifunctional nanoreactors in a variety of research fields.

Acknowledgements

This project was financially supported by the National Natural Science Foundation of China (21273150), the National Basic Research Program of China (2013CB933201), the National High Technology Research and Development Program of China (2011AA03A406, 2012AA062703), and the Fundamental Research Funds for the Central Universities.

References

  1. K. Serp and K. Philippot, Nanomaterials in Catalysis, Wiley-VCH, 2012 Search PubMed.
  2. D. Astruc, F. Lu and J. R. Aranzaes, Angew. Chem., Int. Ed., 2005, 44, 7852–7872 CrossRef CAS PubMed.
  3. W. Yu, M. D. Porosoff and J. G. Chen, Chem. Rev., 2012, 112, 5780–5817 CrossRef CAS PubMed.
  4. G. H. Hutchings and C. J. Kiely, Acc. Chem. Res., 2013, 46, 1759–1772 CrossRef CAS PubMed.
  5. S. Linic, P. Christopher, X. Hong and A. Marimuthu, Acc. Chem. Res., 2013, 46, 1890–1899 CrossRef CAS PubMed.
  6. Y. Xia, X. Xia, Y. Wang and S. Xie, MRS Bull., 2013, 38, 335–344 CrossRef CAS.
  7. Q. L. Zhu, J. Li and Q. Xu, J. Am. Chem. Soc., 2013, 135, 10210–10213 CrossRef CAS PubMed.
  8. T. W. Hansen, A. T. DeLaRiva, S. R. Shalla and A. K. Dayte, Acc. Chem. Res., 2013, 46, 1720–1730 CrossRef CAS PubMed.
  9. R. Ouyang, J. Liu and W. Li, J. Am. Chem. Soc., 2013, 135, 1760–1771 CrossRef CAS PubMed.
  10. M. Pérez-Lorenzo, B. Vaz, V. Salgueiriño and M. A. Correa-Duarte, Chem.–Eur. J., 2013, 19, 12196–12211 CrossRef PubMed.
  11. H. C. Zeng, Acc. Chem. Res., 2013, 46, 226–235 CrossRef CAS PubMed.
  12. Y. Deng, Y. Cai, Z. Sun, J. Liu, C. Liu, J. Wei, W. Li, C. Liu, Y. Wang and D. Zhao, J. Am. Chem. Soc., 2010, 132, 8466–8473 CrossRef CAS PubMed.
  13. J. Ge, Q. Zhang, T. Zhang and Y. Yin, Angew. Chem., Int. Ed., 2008, 47, 8924–8928 CrossRef CAS PubMed.
  14. Y. Dai, B. Lim, Y. Yang, C. M. Cobley, W. Li, E. C. Cho, B. Grayson, P. T. Fanson, C. T. Campbell, Y. Sun and Y. Xia, Angew. Chem., Int. Ed., 2010, 49, 8165–8168 CrossRef CAS PubMed.
  15. P. Lu, C. T. Campbell and Y. Xia, Nano Lett., 2013, 13, 4957–4962 CrossRef CAS PubMed.
  16. Q. Zhang, I. Lee, J. B. Joo, F. Zaera and Y. Yin, Acc. Chem. Res., 2013, 46, 1816–1824 CrossRef CAS PubMed.
  17. G. Li and Z. Tang, Nanoscale, 2014, 6, 3995–4011 RSC.
  18. J. Lu, B. Fu, M. C. Kung, G. Xiao, J. W. Elam, H. H. Kung and P. C. Stair, Science, 2012, 335, 1205–1208 CrossRef CAS PubMed.
  19. L. Shang, T. Bian, B. Zhang, D. Zhang, L. Wu, C. Tung, Y. Yin and T. Zhang, Angew. Chem., Int. Ed., 2014, 53, 250–254 CrossRef CAS PubMed.
  20. Z. Sun, H. Zhang, Y. Zhao, C. Huang, R. Tao, Z. Liu and Z. Wu, Langmuir, 2011, 27, 6244–6251 CrossRef CAS PubMed.
  21. C. T. Campbell, Nat. Chem., 2012, 4, 597–598 CrossRef CAS PubMed.
  22. A. Bruix, J. A. Rodriguez, P. J. Ramírez, S. D. Senanayake, J. Evans, J. B. Park, D. Stacchiola, P. Liu, J. Hrbek and F. Illas, J. Am. Chem. Soc., 2012, 134, 8968–8974 CrossRef CAS PubMed.
  23. P. Hu, Z. Huang, Z. Amghouz, M. Makkee, F. Xu, F. Kapteijn, A. Dikhtiarenko, Y. Chen, X. Gu and X. Tang, Angew. Chem., Int. Ed., 2014, 53, 3418–3421 CrossRef CAS PubMed.
  24. N. Ta, J. Liu, S. Chenna, P. A. Crozier, Y. Li, A. Chen and W. Shen, J. Am. Chem. Soc., 2012, 134, 20585–20588 CrossRef CAS PubMed.
  25. J. Paier, C. Penschke and J. Sauer, Chem. Rev., 2013, 113, 3949–3985 CrossRef CAS PubMed.
  26. N. Acerbi, S. Golunski, S. H. Tsang, H. Daley, C. Hardacre, R. Smith and P. Collier, J. Phys. Chem. C, 2012, 116, 13569–13583 CAS.
  27. C. Sun, H. Li and L. Chen, Energy Environ. Sci., 2012, 5, 8475–8505 CAS.
  28. Y. Jiao, F. Wang, X. Ma, Q. Tang, K. Wang, Y. Guo and L. Yang, Microporous Mesoporous Mater., 2013, 176, 1–7 CrossRef CAS PubMed.
  29. N. C. Strandwitz and G. D. Stucky, Chem. Mater., 2009, 21, 4577–4582 CrossRef CAS.
  30. C. Y. Cao, Z. M. Cui, C. Q. Chen, W. G. Song and W. Cai, J. Phys. Chem. C, 2010, 114, 9865–9870 CAS.
  31. B. Liu, S. Yu, Q. Wang, W. Hu, P. Jing, Y. Liu, W. Jia, Y. Liu, L. Liu and J. Zhang, Chem. Commun., 2013, 49, 3757–3759 RSC.
  32. N. Zhang and Y. J. Xu, Chem. Mater., 2013, 25, 1979–1988 CrossRef CAS.
  33. J. Liu, S. Z. Qiao, H. Liu, J. Chen, A. Orpe, D. Zhao and G. Q. Lu, Angew. Chem., Int. Ed., 2011, 50, 5947–5951 CrossRef CAS PubMed.
  34. T. Teranishi and M. Miyake, Chem. Mater., 1998, 10, 594–600 CrossRef CAS.
  35. S. Guo, S. Dong and E. Wang, ACS Nano, 2010, 4, 547–555 CrossRef CAS PubMed.
  36. H. Sun, J. He, J. Wang, S. Y. Zhang, C. Liu, T. Sritharan, S. Mhaisalkar, M. Y. Han, D. Wang and H. Chen, J. Am. Chem. Soc., 2013, 135, 9099–9110 CrossRef CAS PubMed.
  37. K. Noack, H. Zbinden and R. Schlögl, Catal. Lett., 1990, 4, 145–155 CrossRef CAS.
  38. S. Hinokuma, H. Fujii, M. Okamoto, K. Ikeue and M. Machida, Chem. Mater., 2010, 22, 6183–6190 CrossRef CAS.
  39. K. R. Priolkar, P. Bera, P. R. Sarode, M. S. Hegde, S. Emura, R. Kumashiro and N. P. Lalla, Chem. Mater., 2002, 14, 2120–2128 CrossRef CAS.
  40. M. S. Hegde, G. Madras and K. C. Patil, Acc. Chem. Res., 2009, 42, 704–712 CrossRef CAS PubMed.
  41. B. Basu, S. Das, P. Das, B. Mandal, D. Banerjee and F. Almqvist, Synthesis, 2009, 7, 1137–1146 CrossRef.
  42. B. Basu and S. Paul, Appl. Organomet. Chem., 2013, 27, 588–594 CAS.
  43. S. Zhang, C. Chen, M. Cargnello, P. Fornasiero, R. J. Gorte, G. W. Graham and X. Pan, Nat. Commun., 2015, 6, 7778–7784 CrossRef PubMed.
  44. G. J. Wang, Y. Guo and G. Z. Lu, Fuel Process. Technol., 2015, 130, 71–77 CrossRef CAS PubMed.
  45. S. Praharaj, S. Nath, S. K. Ghosh, S. Kundu and T. Pal, Langmuir, 2004, 20, 9889–9892 CrossRef CAS PubMed.
  46. T. Yu, J. Zeng, B. Lim and Y. Xia, Adv. Mater., 2010, 22, 5188–5192 CrossRef CAS PubMed.
  47. X. Liu, X. Bian, G. Nie, C. Zhang, C. Wang and Y. Wei, J. Mater. Chem., 2012, 22, 12723–12730 RSC.
  48. Y. Mei, Y. Liu, F. Polzer, M. Ballauff and M. Drechsler, Chem. Mater., 2007, 19, 1062–1069 CrossRef CAS.
  49. H. Li, L. Han, J. Cooper-White and I. Kim, Green Chem., 2012, 14, 586–591 RSC.
  50. A. I. Boronin, E. M. Slavinskaya, I. G. Danilova, R. V. Gulyaev, Y. I. Amosov, P. A. Kuznetsov, I. A. Polukhina, S. V. Koscheev, V. I. Zaikovskii and A. S. Noskov, Catal. Today, 2009, 144, 201–211 CrossRef CAS PubMed.
  51. R. V. Gulyaev, A. I. Stadnichenko, E. M. Slavinskaya, A. S. Ivanova, S. V. Koscheev and A. I. Boronin, Appl. Catal., A, 2012, 439, 41–50 CrossRef PubMed.
  52. M. Cargnello, V. V. T. Doan-Nguyen, T. R. Gordon, R. E. Diaz, E. A. Starch, R. J. Gorte, P. Fornasiero and C. B. Murray, Science, 2013, 341, 771–773 CrossRef CAS PubMed.
  53. H. Y. Li, H. F. Wang, X. Q. Gong, Y. L. Guo, Y. Guo, G. Z. Lu and P. Hu, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79(19), 193401–193404 CrossRef.

Footnote

Electronic supplementary information (ESI) available: TEM images: Pd NPs and their size distribution (Fig. S1); Pd NPs on resin polymer spheres (Fig. S2); h-Pd–CeO2 NCSs treated with NaBH4 and then N2 at 650 °C (Fig. S6), after being used in reduction of 4-NP (Fig. S10) and CO oxidation (Fig. S11); a mixture of Pd + CeO2 (Fig. S7); Pd/CeO2-imp (Fig. S8). XPS spectra: h-Pd–CeO2 NCSs (Pd 3d, Fig. S5) and Pd/CeO2-imp (Pd 3d, Fig. S9); h-Pd–CeO2 NCSs and Pd/CeO2-imp (O 1s, Fig. S12). TG curves and EDS line-scanning profiles of h-Pd–CeO2 NCSs (Fig. S3 and S4). Textural properties and crystallite sizes of catalysts (Table S1). See DOI: 10.1039/c5ta05092h

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