Next Article in Journal
Recent Studies on Fluorinated Silica Nanometer-Sized Particles
Next Article in Special Issue
A Porous Carbon with Excellent Gas Storage Properties from Waste Polystyrene
Previous Article in Journal
Towards the Extraction of Radioactive Cesium-137 from Water via Graphene/CNT and Nanostructured Prussian Blue Hybrid Nanocomposites: A Review
Previous Article in Special Issue
Fabrication of Highly Conductive Porous Cellulose/PEDOT:PSS Nanocomposite Paper via Post-Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Preparation and Catalytic Properties of Nanoporous Pt/CeO2 Composites with Nanorod Framework Structures

School of Science, MOE Key Laboratory for Non-equilibrium Synthesis and Modulation of Condensed Matter, Key Laboratory of Shaanxi for Advanced Functional Materials and Mesoscopic Physics, State Key Laboratory for Mechanical Behaviour of Materials, Xi’an Jiaotong University, Xi’an 710049, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(5), 683; https://doi.org/10.3390/nano9050683
Submission received: 26 March 2019 / Revised: 15 April 2019 / Accepted: 23 April 2019 / Published: 2 May 2019

Abstract

:
Pt/CeO2 catalysts with nanoporous structures were prepared by the facile dealloying of melt-spun Al92−XCe8PtX (X = 0.1; 0.3 and 0.5) ribbons followed by calcination. The phase compositions and structural parameters of the catalysts were characterized by X-ray diffraction (XRD), field emission scanning electron microscopy (FESEM) and high-resolution transmission electron microscopy (HRTEM). The specific surface area and pore size distribution were characterized by N2 adsorption–desorption tests. The catalytic properties were evaluated by a three-way catalyst (TWC) measurement system. The results revealed that the dealloyed samples exhibited a nanorod framework structure. The Pt nanoparticles that formed in situ were supported and highly dispersed on the CeO2 nanorod surface and had sizes in the range of 2–5 nm. For the catalyst prepared from the melt-spun Al91.7Ce8Pt0.3 ribbons, the 50% CO conversion temperature (T50) was 91 °C, and total CO could be converted when the temperature was increased to 113 °C. An X-ray photoelectron spectroscopy (XPS) test showed that the Pt0.3/CeO2 sample had a slightly richer oxygen vacancy; and a H2 temperature-programmed reduction (H2-TPR) test demonstrated its superior adsorption ability for reduction gas and high content of active oxygen species. The experiments indicated that the catalytic performance could be retained without any attenuation after 130 h when water and CO2 were present in the reaction gas. The favorable catalytic activities were attributed to the high specific areas and small pore and Pt particle sizes as well as the strong interactions between the CeO2 and Pt nanoparticles. The Pt nanoparticles were embedded in the surface of the CeO2 nanorods, inhibiting growth. Therefore, the catalytic stability and water resistance were excellent.

1. Introduction

The efficient catalytic oxidation of carbon monoxide is a very significant topic because of its widespread applications in CO gas sensors, indoor air purification, and industrial fields, including pollution control devices and automotive exhaust treatment [1,2]. Typically, noble metals and transition metals, such as Au [3,4], Ag [5], Pt [6], Pd [7], Rh [8], Co [9,10] and Cu [11], have been widely investigated as catalysts for CO oxidation. Among these metals, platinum stands out because of its high thermal stability [12]. However, the high cost of Pt greatly restricts its wide application. Furthermore, Pt-based catalysts are susceptible to sintering effects, which can lead to a decrease in the active area of the catalysts because of aggregation and coarsening [13,14,15]. Therefore, it is desirable to reduce the content of Pt in a catalyst and introduce metal oxides to construct Pt- and metal oxide-based heterostructures. Noble metals dispersed on specific metal oxides with high storage/release oxygen capacities are highly active towards catalytic CO oxidation.
Cerium dioxide (CeO2), as a representative and also one of the most abundant rate earth oxides on the planet, has unique heterogeneous catalytic abilities due to its excellent storage/release oxygen capacity and numerous oxygen vacancy defects [16]. CeO2 also provides an oxygen source through the fast and reversible Ce4+/Ce3+ redox reaction [17]. Therefore, based on these advantages, CeO2 has a wide range of potential applications and favorable development prospects in the field of air pollution. The morphology and facets of CeO2-based nanocomposites can greatly influence the formation and migration of surface oxygen vacancies. Also, nanosized structured CeO2 materials, including nanospheres, nanorods [18], nanocubes and nanotubes, have been synthesized [12,19,20,21]. Among these structures, nanorod-shaped CeO2 has received a substantial amount of attention because of its potentially large surface area and abundance of oxygen vacancy defects.
Because of its enhanced catalytic performance and wide range of applications, such as CO oxidation and methane partial oxidation, Pt/CeO2 has attracted increasing attention [22,23]. As the synthetic method has a large influence on the distribution of the oxygen vacancies and the structure as well as the ability to control the activity of a catalyst, a number of preparation methods, including the template, conventional impregnation, deposition–precipitation, and photodeposition methods, have been developed to synthesize Pt/CeO2 catalysts [24,25]. For example, Wei et al. prepared three-dimensional Pt@CeO2 nanoparticles through a colloidal crystal template method, which exhibited a high catalytic activity [26]. However, the impregnation and deposition–precipitation methods cannot effectively control the size and dispersion of the metals; additionally, the high-cost, sophisticated technology and low yield of the template method restricts its further development [27,28]. Dealloying, a facile method for fabricating nanoporous materials, has been attracting increased attention because of its green and effective preparation characteristics. Zhang et al. reported highly active Cu-doped CeO2 nanowires fabricated from a dealloying–annealing method for CO oxidation [29]. Our team has also performed research on CeO2, NiO/CeO2 and Ag/CeO2 nanorod frameworks synthesized by dealloying methods, which showed superior catalytic activity for CO oxidation due to their high porosities and large specific surface areas [30,31,32]. However, only a few works have reported the fabrication of Pt/CeO2 nanorods through a facile dealloying method for catalytic CO oxidation.
In this paper, a Pt/CeO2 nanorod framework catalyst for CO oxidation was fabricated by dealloying a melt-spun Al–Ce–Pt alloy followed by calcination. SEM and TEM characterization demonstrated that the Pt nanoparticles that formed in situ were supported and highly dispersed on the CeO2 nanorod surface, and an enhanced catalytic performance was obtained. For the Pt/CeO2 nanorods prepared from the Al91.7Ce8Pt0.3 precursor alloy, the 50% CO conversion temperature (T50) was 91 °C, and all of the CO could be converted when the temperature was increased to 113 °C. An excellent square curve for the degree of CO conversion was exhibited. The formation mechanism, microstructure and catalytic mechanism of the nanoporous Pt/CeO2 were studied.

2. Experimental

2.1. Preparation of the Pt/CeO2 Catalysts

As-cast Al91.9Ce8Pt0.1, Al91.7Ce8Pt0.3 and Al91.5Ce8Pt0.5 alloy button ingots were prepared from pure Al (99.90 wt%), pure Ce (99.90 wt%) and pure Pt (99.99 wt%) by arc-melting six times under a protective atmosphere of high-purity Ar. The ribbons were obtained after the alloy ingots were re-melted and rapidly solidified into foils through a melt-spinning technique under a high-purity Ar atmosphere; the resultant ribbons were 3–4 mm in width and 20–40 μm thick. Subsequently, the precursor alloy ribbons were dealloyed in a 20 wt% NaOH aqueous solution at room temperature for 2 h and then at 80 °C for 10 h. The resulting Pt/Ce–OH precursor nanorod framework-structured samples were carefully rinsed with deionized water several times. Finally, the dealloyed ribbons were calcined in a muffle furnace at 200 °C, 300 °C, 400 °C and 500 °C, respectively, for 2 h.

2.2. Material Characterization

The phase compositions of the dealloyed Al–Ce–Pt ribbons were collected by X-ray diffractometry (XRD-6100, Shinadzu Inc., Kyoto, Japan) with radiation from a Cu target (Kα, λ = 0.3615 nm). The morphologies and microstructures of the dealloyed ribbons were measured by field emission scanning electron microscopy (FESEM, JSM-7000F, JEOL Inc., Tokyo, Japan) and high-resolution transmission electron microscopy (HRTEM, JEM-2100, JEOL Inc., Tokyo, Japan). Energy dispersive spectroscopy (EDS) analysis and mapping were performed using scanning transmission electron microscopy (STEM, JEM-F200, JEOL Inc., Tokyo, Japan) equipped with an Oxford Instruments EDS spectrometer. An X-ray photoelectron spectroscopy (XPS, ESCALAB Xi+, Fisher Inc., Marshalltown, IA, USA) with Al-Ka radiation was used to investigate the element composition and valence state of the sample. The surface areas of the samples were determined by the Brunauer–Emmett–Teller (BET) method, and their pore size distributions were determined by the desorption branch of the isotherm using the Barrett–Joyner–Halenda (BJH) algorithm. The N2 sorption isotherms were collected using an ASAP 2020 (Micromeritics Inc., Norcross, GA, USA) system at a temperature of 77 K. The reduction behaviors and the interaction between the active phase and the support of each sample were examined by using the H2 temperature-programmed reduction (H2-TPR) technique. The H2-TPR measurements were performed with an Auto ChemTM II 2920 Instrument (Micromeritics Inc., Norcross, GA, USA) catalyst characterization system. H2-TPR was performed in a quartz cell, and approximately 50 mg of the sample was used for the measurement. The test temperature ranged from 50 to 850 °C with a heating rate of 10 °C/min.

2.3. Evaluation of the CO Oxidation Performance

The catalytic CO oxidation performance of the Pt/CeO2 catalyst was measured in a fixed-bed flow reactor with an internal diameter of 9 mm configured in a tube furnace. In a typical experiment, approximately 100 mg of the Pt/CeO2 catalyst was placed in the middle of the reactor supported by a quartz wool plug. The reaction gas mixture consisted of 1% CO, 10% O2 and 89% N2 (volume fraction). The total reaction gas flow was 100 mL min−1, and the corresponding space velocity was 60,000 h−1. The outlet gas was analyzed by an 7890B gas chromatograph equipped with a hydrogen flame detector (FID) (Anglit, Palo Alto, CA, USA). The CO conversion for CO oxidation can be calculated as follows:
CO   conversion = C i n C o u t C i n × 100 %
where Cin was the initial concentration of CO at room temperature and Cout was the corresponding concentration of CO at the outlet as the temperature was increased.

3. Results

Figure 1 shows the XRD patterns of dealloyed Al91.9Ce8Pt0.1, Al91.7Ce8Pt0.3 and Al91.5Ce8Pt0.5 ribbons calcined at different temperatures. As shown in Figure 1a, the diffraction peaks of the three samples can be indexed to the spinel cubic phase of CeO2 (PDF # 89-8436). However, no diffraction peaks corresponding to Pt were detected in the dealloyed and calcined Al91.9Ce8Pt0.1 sample; the intensity of the Pt peak in the diffraction patterns increased with increasing Pt content of the precursor alloys. The crystallinity of the dealloyed Al91.7Ce8Pt0.3 sample, as shown in Figure 1b, increased with increasing calcination temperature. However, the intensity of the Pt peak did not become stronger as the Pt content in the precursor alloys increases, which might be attributed to the high dispersity of the Pt.
The surface morphologies of the dealloyed and calcined Al91.9Ce8Pt0.1, Al91.7Ce8Pt0.3 and Al91.5Ce8Pt0.5 ribbons are shown in Figure 2. After dealloying and calcining, the CeO2 nanorods were formed and arranged in a framework structure with pores and holes distributed between the nanorods, as shown in Figure 2a,d,g. In this work, the diameter of the nanorods was approximately 20 nm in all the samples, indicating a small effect from the Pt content. At lower magnifications, the Pt/CeO2 appeared as a bulk material (insert of the figure). The TEM images of the Pt/CeO2 catalysts are shown in Figure 2b,e,h. When the Pt content is very low (Pt0.1/CeO2), it is difficult to observe the existence of Pt nanoparticles due to the low Pt content, as shown in Figure 2b. As observed from the SAED image in Figure 2c, only the diffraction ring of CeO2 can be observed, further indicating the low content and poor crystallinity of Pt on the surface of the CeO2 nanorod. When the Pt content in the precursor alloys increased to 0.3%, the size of the Pt nanoparticles was approximately 3–4 nm, as shown in Figure 2e. However, it is difficult to observe the diffraction ring of Pt from the selected area electron diffraction (SAED) image in Figure 2f, implying the existence of amorphous Pt. When the Pt content increased to 0.5%, aggregation of nanoparticles occurred, as observed from the yellow circle in Figure 2h. Furthermore, as clearly observed from Figure 2i, the reflection from (111) plane corresponded closely to Pt, which is consistent with the XRD results in Figure 1.
To further prove the distribution of Pt in the Pt/CeO2 catalysts, STEM imaging and EDS mapping of the dealloyed Al91.7Ce8Pt0.3 ribbon calcined at 300 °C were conducted. The results indicate that the Pt nanoparticles were supported in situ and highly dispersed on the surface of the CeO2 nanorods, as shown in Figure 3. The nanoparticles had an average size of 3.5 nm, which is consistent with the TEM results above. In our previous works [31,32], Au/CeO2 catalysts with nanorod framework structures were prepared, and the formation of the catalysts was analyzed in detail. In this work, Pt/CeO2 catalysts with similar structures were obtained by dealloying and calcining melt-spun Al–Ce–Pt ribbons, and their formation mechanism was similar.
The specific surface area and pore size distribution of Pt/CeO2 nanorod framework were characterized by N2 adsorption–desorption tests. The N2 sorption isotherms and Barrett–Joyner–Halenda (BJH) pore size distributions are presented in Figure 4. The N2 sorption isotherms were type IV isotherms, with a hysteresis loop occurring at relative pressures ranging from 0.6–1.0 P/P0, as shown in Figure 4a, demonstrating the presence of a mesoporous structure [33]. The Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 nanorod frameworks had Brunauer–Emmett–Teller (BET) surface areas of 148.13, 161.95 and 148.44 m2 g−1, respectively, and the corresponding total pore volumes were 0.352, 0.3486 and 0.3264 cm3 g−1, respectively. The pore sizes for the Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 samples were centered around 8.88, 8.73 and 9.19 nm, respectively. Among them, the dealloyed and calcined Al91.7Ce8Pt0.3 sample exhibited the highest specific surface area and lowest pore diameter.
Figure 5 shows the catalytic performance of the Pt/CeO2 nanorod framework-structured catalysts with different Pt contents towards CO catalytic oxidation. For pure CeO2 prepared from dealloying Al92Ce8 alloys, the conversion temperatures for catalytic CO oxidation at 50% (T50) and 99% (T99) levels were 235 and 320 °C, respectively. The catalytic CO-oxidized performances were markedly improved upon the addition of Pt in the Al–Ce precursor alloys. For the Pt0.1/CeO2 catalyst, the temperatures for conversion of CO oxidation at 50% (T50) and 99% (T99) levels were 105 and 120 °C, respectively. As the Pt content was increased to 0.3% in the precursor alloy, the temperatures for conversion of CO oxidation at 50% (T50) and 99% (T99) levels were reduced to 91 °C and 113 °C, respectively. After the Pt content was increased to 0.5% in the precursor alloy, the total conversion temperature (T99) was 128 °C, and the catalytic activity followed a downward trend, as shown in Figure 5a. The temperature difference between T50 and T99 was only 22 °C for the Pt0.3/CeO2 catalyst, exhibiting a high thermal stability and an outstanding conversion rate at high temperatures. This means that the Pt/CeO2 catalyst exhibited a better square curve for the degree of CO conversion compared with that of the Au/CeO2 catalyst in our previous work [32]. Figure 5b shows the effect of the calcination temperature on the CO conversion of the catalysts prepared from the Al91.7Ce8Pt0.3 alloy. As observed, the total conversion temperatures (T99) for Pt0.3/CeO2 without calcination, and calcined at 200 °C, 300 °C, 400 °C, 500 °C were 170 °C, 113 °C, 113 °C, 117 °C, 126 °C, respectively. The conversion rates of the composites improved slightly with increasing calcination temperature from room temperature to 300 °C and then decreased. In comparison, the best catalytic performance was obtained by annealing at 200 °C and 300 °C, which resulted in complete conversion temperatures of approximately 113 °C.
The long-term stability of the Pt0.3/CeO2 catalyst was evaluated, and the results are shown in Figure 6. The Pt0.3/CeO2 catalyst exhibited nearly 98% CO conversion without noticeable deterioration in activity for a holding time of 75 h at 110 °C under a mixed composition of 1% CO, 10% O2 and 89% N2. To examine the long-term stability and the water resistance of the catalysts, CO conversion as a function of the holding time and the environmental effects was measured. The results reveal that the catalyst remained stable for 130 h when the reaction gas was mixed with 10% H2O. Typically, the addition of CO2 will cause adsorption competition with CO on the surface of Pt nanoparticles or the boundaries Pt and oxide, leading to a reduction in adsorption of activated CO per unit time, thus resulting in a decrease in catalytic CO oxidation activity [34,35,36]. However, the CO conservation could still remain stable when 5% CO2 and 10% H2O were introduced into the reaction gas after 65 h. This result implies that the nanorod framework structure in the present work is conducive to maintaining a high stability and excellent water resistance, even with the existence of adsorption competition.
Figure 7 shows an SEM image of the dealloyed Al91.7Ce8Pt0.3 calcined at 300 °C after continuous reaction for 130 h. The results show that the porous structure of the sample surface was well maintained and can clearly be observed after long-term testing. This result further demonstrates that the nanorod framework structure is conducive to maintaining a high stability.
X-ray photoelectron spectroscopy (XPS) was characterized to investigate the surface chemical states of the three catalysts. As shown in Figure 8a, the Ce 3d spectra of the Pt/CeO2 catalysts fitted into eight peaks, which are consistent with the peaks reported in the literature. The six peaks at 881.9 eV, 888.5 eV, 897.6 eV, 900.4 eV, 906.6 eV, and 916.1 eV can be assigned to Ce4+ [27]; while the other two peaks at 883.9 eV and 901.9 eV can be ascribed to Ce3+ [37,38]. Additionally, as calculated from the results of the Ce 3d spectra, the Ce3+ content of Pt0.3/CeO2 was 21.7%, which was a little higher than that of Pt0.1/CeO2 (21.2%) and Pt0.5/CeO2 (20.0%), indicating a higher concentration of oxygen vacancy on the surface of the Pt0.3/CeO2 sample [32,39]. Figure 8b shows that the ratios of Pt0/(Pt0 + Pt2+) for the Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts are 65.8%, 64.4% and 62.1%, respectively. For the Pt 4f spectrum of Pt0.1/CeO2, the peaks at 70.8 eV and 74.1 eV are attributed to the Pt0 species, and the peaks at 71.5 eV and 74.8 eV are attributed to the Pt2+ species. For the Pt 4f spectrum of Pt0.3/CeO2, the peaks at 70.8 eV and 74.1 eV are attributed to the Pt0 species, and the peaks at 72.0 eV and 76 eV are attributed to the Pt2+ species. Similarly, for the Pt 4f spectrum of Pt0.5/CeO2, the peaks at 70.7 eV and 74.0 eV are ascribed to the Pt0 species, and the peaks at 71.7 eV and 76.1 eV are attributed to the Pt2+ species [40].
Temperature-programmed reduction (TPR) by H2 was extensively used to characterize the oxygen reducibility of the dealloyed and calcined Al–Ce–Pt catalysts. Three distinct hydrogen consumption peaks can be observed in Figure 9. For Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2, the first reduction peak is located at 101 °C, 80 °C, 116 °C, respectively. The second peak is located at 486 °C, 403 °C, 481 °C, respectively. The third peak is located at approximately 740 °C.

4. Discussions

The SEM and TEM images indicate the framework structures of the catalysts, with Pt nanoparticles of 3.5 nm in size being highly dispersed on the surface of the nanorods. The STEM analysis also proved that the Pt nanoparticles are formed in situ, supported and highly dispersed onto the surface of the CeO2 nanorods. This structure not only reflects the high catalytic activity of the Pt nanoparticles (Figure 5) but can also inhibit the agglomeration and growth of Pt particles during the catalytic process (Figure 7), thus contributing to the enhanced catalytic performance. The results of the BET and BJH analysis indicate that the dealloyed and calcined Al91.7Ce8Pt0.3 sample exhibited the highest specific surface area and lowest pore diameter among the three catalysts. Generally, a high specific surface area can provide an abundance of active catalytic sites for CO oxidation, while a good mesoporous structure provides more gas diffusion channels for the catalytic reaction [30]. In comparison, the dealloyed and calcined Al91.7Ce8Pt0.3 sample exhibited the highest specific surface area and lowest pore diameter among the three samples, thus exhibiting the best catalytic performance in the present work. The XPS results confirm that the Pt0.3/CeO2 catalyst has higher proportions of Ce3+ than those of the Pt0.1/CeO2 and Pt0.5/CeO2 catalysts, indicating its higher concentration of oxygen vacancies and active sites. The oxygen vacancies can induce the adsorption of oxygen species with low-temperature adsorption–desorption properties and can promote the transfer of electrons between metal ions as well as the activity of the lattice oxygen; the active sites can provide more reaction paths for catalytic CO oxidation, thereby enhancing the catalytic ability. For the H2-TPR result in Figure 9, the appearance of the first peak for Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts is due to Pt-oxide reduction, which demonstrates that the Pt0.3/CeO2 sample shows better adsorption ability for reduction gas. The occurrence of the second peak is ascribed to the reduction of surface-active oxygen. The reduction peak for the Pt0.3/CeO2 sample stays left, and its integrated area of the peak is much larger than that of the other two samples, implying that the content of surface-active oxygen on CeO2 for Pt0.3/CeO2 is significantly higher than that of the other two samples. The amount of surface-active oxygen is also an important element in the evaluation of the catalytic ability of a catalyst, thus further demonstrating that Pt0.3/CeO2 possesses the best catalytic activity. The third peak at approximately 740 °C is due to the existence of CeO2 in the catalyst [41,42,43]. Durability testing of the Pt0.3/CeO2 catalysts indicated that the catalyst had a good ability, which may be explained by the high catalytic activity of the in situ supported Pt nanoparticles and the thermal and catalytic stability of the nanorod framework structure. The CeO2 catalysts can effectively prevent the deposition of carbon and lead to the formation of fewer by-products during the catalytic process.
A comparison of the nanorod framework-structured Pt/CeO2 catalysts with different reported Pt/CeO2 catalysts is provided in Table 1. The results show that the nanorod framework structure of the Pt/CeO2 catalysts in the present paper exhibited a high catalytic activity and an excellent stability compared to those of previous reports.
The corresponding schematic illustration of the CO oxidation mechanism is shown in Figure 10. The oxygen vacancies, which play a key role in CO catalytic oxidation, are mainly generated at the Pt–CeO2 interface or boundary. Initially, CO is mainly adsorbed onto the Pt surface and reacts with the oxygen species activated by the oxygen vacancies at the Pt–CeO2 boundary, leading to the formation of CO2 and more surface oxygen vacancies. The strongly adsorbed CO reacts with the active oxygen species adsorbed at the surface vacancies [46]. The existence of the Pt particles can accelerate the migration of active oxygen adsorbed onto the CeO2 nanorod. The migration of the generated oxygen vacancies provides access to neighboring surface lattice oxygen for the following cycles [32], contributing to the high and efficient CO conversion. This result conforms to the CeO2-assisted Mars–van Krevelen mechanism [46].

5. Conclusions

In summary, the nanorod framework-structured Pt/CeO2 catalysts were fabricated by a simple dealloying method. The SEM and TEM analysis indicated the framework structure of the catalysts, with Pt nanoparticles of 3.5 nm in size highly dispersed over the surface. The catalytic tests showed that the T99 temperatures of CeO2, Pt0.1/CeO2, Pt0.3/CeO2, Pt0.5/CeO2 were 320 °C, 120 °C, 113 °C, 128 °C, respectively. Obviously, the Pt/CeO2 catalysts exhibited enhanced catalytic performance towards CO oxidation compared with CeO2, which is ascribed to their high specific surface area and the uniform distribution of Pt particles. XPS and H2-TPR results showed that the Pt0.3/CeO2 sample exhibited a higher Ce3+ cation concentration, richer oxygen vacancy and better adsorption ability for reduction gas, thus showing the best catalytic activity. The Pt0.3/CeO2 sample showed the highest catalytic activity when calcined at 300 °C. Moreover, the catalyst could remain stable for 130 h when water was added to the reaction gas, demonstrating the excellent long-term stability and strong water resistance of the catalyst.

Author Contributions

Conceptualization, H.W. and D.D.; contribution to samples preparation, characterization and measurement, H.W., D.D. and W.S.; writing-original draft preparation, H.W. and M.L.; contribution to discussion and analysis, H.W., M.L., L.W., L.G. and C.M.; writing-review and editing, H.W. and Z.S.; supervision, X.S. and Z.S.; funding acquisition, X.S. All authors read and approved the final version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (Grant No. 51771141 and 51671155).

Acknowledgments

The authors also thank Jiamei Liu and Gang Chang at the Instrument Analysis Center of Xi’an Jiaotong University for their assistance with XPS analysis and H2-TPR characterization.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Min, B.K.; Friend, C.M. Heterogeneous Gold-Based Catalysis for Green Chemistry: Low-Temperature CO Oxidation and Propene Oxidation. Chem. Rev. 2007, 107, 2709–2724. [Google Scholar] [CrossRef]
  2. Zhang, J.S.; Di, L.B.; Yu, F.; Duan, D.Z.; Zhang, X.L. Atmospheric-Pressure Cold Plasma Activating Au/P25 for CO Oxidation: Effect ofWorking Gas. Nanomaterials 2018, 8, 742. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, H.; Shen, J.; Huang, J.; Xu, T.; Zhu, J.; Zhu, Y.; Li, C. Atomically dispersed Au catalysts supported on CeO2 foam: Controllable synthesis and CO oxidation reaction mechanism. Nanoscale 2017, 9, 16817–16825. [Google Scholar] [CrossRef] [PubMed]
  4. Najafishirtari, S.; Guardia, P.; Scarpellini, A.; Prato, M.; Marras, S.; Manna, L.; Colombo, M. The effect of Au domain size on the CO oxidation catalytic activity of colloidal Au-FeOx dumbbell-like heterodimers. J. Catal. 2016, 338, 115–123. [Google Scholar] [CrossRef]
  5. Zhang, J.; Li, L.; Huang, X.; Li, G. Fabrication of Ag-CeO2 core-shell nanospheres with enhanced catalytic performance due to strengthening of the interfacial interactions. J. Mater. Chem. 2012, 22, 10480–10487. [Google Scholar] [CrossRef]
  6. Bera, P.; Patil, K.C.; Jayaram, V.; Subbanna, G.N.; Hegde, M.S. Ionic Dispersion of Pt and Pd on CeO2 by Combustion Method: Effect of Metal-Ceria Interaction on Catalytic Activities for NO Reduction and CO and Hydrocarbon Oxidation. J. Catal. 2000, 196, 293–301. [Google Scholar] [CrossRef]
  7. Vogel, D.; Spiel, C.; Suchorski, Y.; Trinchero, A.; Schlögl, R.; Grönbeck, H.; Rupprechter, G. Local Catalytic Ignition during CO Oxidation on Low-Index Pt and Pd Surfaces: A Combined PEEM, MS, and DFT Study. Angew. Chem. Int. Ed. 2012, 51, 10041–10044. [Google Scholar] [CrossRef]
  8. Huang, H.; Xu, Y.; Feng, Q.; Leung, D.Y.C. Low temperature catalytic oxidation of volatile organic compounds: A review. Catal. Sci. Technol. 2015, 5, 2649–2669. [Google Scholar] [CrossRef]
  9. Hu, P.D.; Long, M.C. Cobalt-catalyzed sulfate radical-based advanced oxidation: A review on heterogeneous catalysts and applications. Appl. Catal. B Environ. 2016, 181, 103–117. [Google Scholar] [CrossRef]
  10. Tang, Y.; Ma, L.; Dou, J.; Andolina, C.M.; Li, Y.; Ma, H.; House, S.D.; Zhang, X.; Yang, J.; Tao, F.F. Transition of surface phase of cobalt oxide during CO oxidation. Phys. Chem. Chem. Phys. 2018, 20, 6440–6449. [Google Scholar] [CrossRef] [PubMed]
  11. Lykakia, M.; Pachatouridou, E.; Carabineiroc, S.A.C.; Iliopouloub, E.; Andriopouloud, C.; Kallithrakas-Kontose, N.; Soghomon, B.; Konsolakis, M. Ceria nanoparticles shape effects on the structural defects and surface chemistry: Implications in CO oxidation by Cu/CeO2 catalysts. Appl. Catal. B Environ. 2018, 230, 18–28. [Google Scholar] [CrossRef]
  12. Singhania, N.; Anumol, E.A.; Ravishankar, N.; Madras, G. Influence of CeO2 morphology on the catalytic activity of CeO2-Pt hybrids for CO oxidation. Dalton Trans. 2013, 42, 15343. [Google Scholar] [CrossRef]
  13. Tan, H.; Wang, J.; Yu, S.; Zhou, K. Support Morphology-Dependent Catalytic Activity of Pd/CeO2 for Formaldehyde Oxidation. Environ. Sci. Technol. 2015, 49, 8675–8682. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, K.; Fu, X.P.; Yu, W.Z.; Wang, W.W.; Jia, C.J.; Du, P.P.; Si, R.; Wang, Y.H.; Li, L.D.; Zhou, L.; et al. Pt-Embedded CuOx-CeO2 Multicore-Shell Composites: Interfacial Redox Reaction-Directed Synthesis and Composition-Dependent Performance for CO Oxidation. ACS Appl. Mater. Interfaces 2018, 10, 34172–34183. [Google Scholar] [CrossRef] [PubMed]
  15. Zhou, A.; Wang, J.; Wang, H.; Li, H.; Wang, J.; Shen, M. Effect of active oxygen on the performance of Pt/CeO2 catalysts for CO oxidation. J. Rare Earths 2018, 36, 257–264. [Google Scholar] [CrossRef]
  16. Medina, O.E.; Gallego, J.; Arias-Madrid, D.; Cortés, F.B.; Franco, C.A. Optimization of the Load of Transition Metal Oxides (Fe2O3, Co3O4, NiO and/or PdO) onto CeO2 Nanoparticles in Catalytic Steam Decomposition of n-C7 Asphaltenes at Low Temperatures. Nanomaterials 2019, 9, 401. [Google Scholar] [CrossRef]
  17. Lin, K.S.; Chowdhury, S. Synthesis, characterization, and application of 1-D cerium oxide nanomaterials: A review. Int. J. Mol. Sci. 2010, 11, 3226–3251. [Google Scholar] [CrossRef]
  18. Zhou, K.; Wang, X.; Sun, X.; Peng, Q.; Li, Y. Enhanced catalytic activity of ceria nanorods from well-defined reactive crystal planes. J. Catal. 2005, 229, 206–212. [Google Scholar] [CrossRef]
  19. Pan, C.; Zhang, D.; Shi, L.; Fang, J. Template-Free Synthesis, Controlled Conversion, and CO Oxidation Properties of CeO2 Nanorods, Nanotubes, Nanowires, and Nanocubes. Eur. J. Inorg. Chem. 2008, 2008, 2429–2436. [Google Scholar] [CrossRef]
  20. Tana; Zhang, M.; Li, J.; Li, H.; Li, Y.; Shen, W. Morphology-dependent redox and catalytic properties of CeO2 nanostructures: Nanowires, nanorods and nanoparticles. Catal. Today 2009, 148, 179–183. [Google Scholar] [CrossRef]
  21. Mai, H.X.; Sun, L.D.; Zhang, Y.W.; Si, R.; Feng, W.; Zhang, H.P.; Liu, H.C.; Yan, C.H. Shape-Selective Synthesis and Oxygen Storage Behavior of Ceria Nanopolyhedra, Nanorods, and Nanocubes. J. Phys. Chem. B 2005, 109, 24380–24385. [Google Scholar] [CrossRef]
  22. Yu, W.; Porosoff, M.D.; Chen, J.G. Review of Pt-based bimetallic catalysis: From model surfaces to supported catalysts. Chem. Rev. 2012, 112, 5780–5817. [Google Scholar] [CrossRef]
  23. Bera, P.; Gayen, A.; Hegde, M.S.; Lalla, N.P.; Spadaro, L.; Frusteri, F.; Arena, F. Promoting Effect of CeO2 in Combustion Synthesized Pt/ CeO2 Catalyst for CO Oxidation. J. Phys. Chem. B 2003, 107, 6122–6130. [Google Scholar] [CrossRef]
  24. Wang, X.; Liu, D.; Song, S.; Zhang, H. Pt@ CeO2 multicore@shell self-assembled nanospheres: Clean synthesis, structure optimization, and catalytic applications. J. Am. Chem. Soc. 2013, 135, 15864–15872. [Google Scholar] [CrossRef]
  25. Wu, K.; Zhou, L.; Jia, C.J.; Yan, C.H. Pt-embedded-CeO2 hollow spheres for enhancing CO oxidation performance. Mater. Chem. Front. 2017, 1, 1754–1763. [Google Scholar] [CrossRef]
  26. Wei, Y.; Zhao, Z.; Liu, J.; Xu, C.; Jiang, G.; Duan, A. Design and synthesis of 3D ordered macroporous CeO2-supported Pt@CeO2 core-shell nanoparticle materials for enhanced catalytic activity of soot oxidation. Small 2013, 9, 3957–3963. [Google Scholar] [CrossRef]
  27. Cao, F.X.; Zhang, S.; Gao, W.; Cao, T.; Qu, Y.Q. Facile synthesis of highly-dispersed Pt/CeO2 by a spontaneous surface redox chemical reaction for CO oxidation. Catal. Sci. Technol. 2018, 8, 3233–3237. [Google Scholar] [CrossRef]
  28. Cao, S.; Tao, F.F.; Tang, Y.; Li, Y.; Yu, J. Size- and shape-dependent catalytic performances of oxidation and reduction reactions on nanocatalysts. Chem. Soc. Rev. 2016, 45, 4747–4765. [Google Scholar] [CrossRef]
  29. Kou, T.; Si, C.; Pinto, J.; Ma, C.; Zhang, Z. Dealloying assisted high-yield growth of surfactant-free <110> highly active Cu-doped CeO2 nanowires for low-temperature CO oxidation. Nanoscale 2017, 9, 8007–8014. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, X.; Li, G.; Yang, S.; Song, X.; Sun, Z. Nanoporous CuO ribbons modified by Au nanoparticles through chemical dealloying and calcination for CO oxidation. Microporous Mesoporous Mater. 2016, 226, 61–70. [Google Scholar] [CrossRef]
  31. Zhang, X.; Li, K.; Shi, W.; Wei, C.; Song, X.; Yang, S.; Sun, Z. Baize-like CeO2 and NiO/CeO2 nanorod catalysts prepared by dealloying for CO oxidation. Nanotechnology 2017, 28, 045602. [Google Scholar] [CrossRef]
  32. Zhang, X.; Duan, D.; Li, G.; Feng, W.; Yang, S.; Sun, Z. Monolithic Au/CeO2 nanorod framework catalyst prepared by dealloying for low-temperature CO oxidation. Nanotechnology 2018, 29, 095606. [Google Scholar] [CrossRef]
  33. Wang, H.; Liang, M.; Duan, D.; Shi, W.; Song, Y.; Sun, Z. Rose-like Ni3S4 as battery-type electrode for hybrid supercapacitor with excellent charge storage performance. Chem. Eng. J. 2018, 350, 523–533. [Google Scholar] [CrossRef]
  34. Liang, F.; Zhu, H.; Qin, Z.; Wang, G.; Wang, J. Effects of CO2 on the stability of Pd/CeO2–TiO2 catalyst for low-temperature CO oxidation. Catal. Commun. 2009, 10, 737–740. [Google Scholar] [CrossRef]
  35. Aguilar-Guerrero, V.; Gates, B.C. Kinetics of CO oxidation catalyzed by highly dispersed CeO2-supported gold. J. Catal. 2008, 260, 351–357. [Google Scholar] [CrossRef]
  36. Ojala, T.; Pietikainen, M.; Harwood, D. A comparative study of texture measures with classification based on feature distributions. Pattern Recoonit. 1996, 29, 51–59. [Google Scholar] [CrossRef]
  37. Peng, R.; Li, S.; Sun, X.; Ren, Q.; Chen, L.; Fu, M.; Wu, J.; Ye, D. Size effect of Pt nanoparticles on the catalytic oxidation of toluene over Pt/ CeO2 catalysts. Appl. Catal. B Environ. 2018, 220, 462–470. [Google Scholar] [CrossRef]
  38. Peng, R.; Sun, X.; Li, S.; Chen, L.; Fu, M.; Wu, J.; Ye, D. Shape effect of Pt/CeO2 catalysts on the catalytic oxidation of toluene. Chem. Eng. J. 2016, 306, 1234–1246. [Google Scholar] [CrossRef]
  39. Wang, B.; Chen, B.; Sun, Y.; Xiao, H.; Xu, X.; Fu, M.; Wu, J.; Chen, L.; Ye, D. Effects of dielectric barrier discharge plasma on the catalytic activity of Pt/CeO2 catalysts. Appl. Catal. B Environ. 2018, 238, 328–338. [Google Scholar] [CrossRef]
  40. Stadnichenko, A.I.; Muravev, V.V.; Koscheev, S.V.; Zaikovskii, V.I.; Aleksandrov, H.A.; Neyman, K.M.; Boronin, A.I. Study of active surface centers of Pt/CeO2 catalysts prepared using radio-frequency plasma sputtering technique. Surf. Sci. 2019, 679, 237–283. [Google Scholar] [CrossRef]
  41. Luo, M.F.; Zhong, Y.J.; Yuan, X.X.; Zheng, X.M. TPR and TPD studies of CuO/ CeO2 catalysts for low temperature CO oxidation. Appl. Catal. A Gen. 1997, 162, 121–131. [Google Scholar] [CrossRef]
  42. Zhu, H.Q.; Qin, Z.; Shan, W.J.; Shen, W.J.; Wang, J.G. Pd/ CeO2-CeO2 catalyst for CO oxidation at low temperature: A TPR study with H2 and CO as reducing agents. J. Catal. 2004, 225, 267–277. [Google Scholar] [CrossRef]
  43. Si, G.; Yu, J.; Xiao, X.; Guo, X.; Huang, H.; Mao, D.; Lu, G. Boundary role of Nano-Pd catalyst supported on ceria and the approach of promoting the boundary effect. Mol. Catal. 2018, 444, 1–9. [Google Scholar] [CrossRef]
  44. Chen, G.Z.; Yang, Y.; Guo, Z.; Gao, D.; Zhao, W.; Yan, H.; Wang, W.; Jia, C.J.; Sun, G.X. Thermally stable and highly active Pt/CeO2@SiO2 catalysts with a porous/hollow structure. Catal. Sci. Technol. 2018, 8, 4413–4419. [Google Scholar] [CrossRef]
  45. Chen, J.; Wanyan, Y.; Zeng, J.; Fang, H.; Li, Z.; Dong, Y.; Qin, R.; Wu, C.; Liu, D.; Wang, M.; et al. Surface Engineering Protocol to Obtain an Atomically Dispersed Pt/ CeO2 Catalyst with High Activity and Stability for CO Oxidation. ACS Sustain. Chem. Eng. 2018, 6, 14054–14062. [Google Scholar] [CrossRef]
  46. Liu, B.; Liu, J.; Li, T.; Zhao, Z.; Gong, X.Q.; Chen, Y.; Duan, A.; Jiang, G.; Wei, Y. Interfacial Effects of CeO2-Supported Pd Nanorod in Catalytic CO Oxidation: A Theoretical Study. J. Phys. Chem. C 2015, 119, 12923–12934. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) dealloyed Al91.9Ce8Pt0.1, Al91.7Ce8Pt0.3 and Al91.5Ce8Pt0.5 ribbons calcined at 300 °C and (b) dealloyed Al91.7Ce8Pt0.3 ribbons calcined at 200, 300, 400 and 500 °C.
Figure 1. XRD patterns of (a) dealloyed Al91.9Ce8Pt0.1, Al91.7Ce8Pt0.3 and Al91.5Ce8Pt0.5 ribbons calcined at 300 °C and (b) dealloyed Al91.7Ce8Pt0.3 ribbons calcined at 200, 300, 400 and 500 °C.
Nanomaterials 09 00683 g001
Figure 2. SEM, TEM and selected area electron diffraction (SAED) images of (ac) dealloyed Al91.9Ce8Pt0.1, (df) Al91.7Ce8Pt0.3 and (gi) Al91.5Ce8Pt0.5 ribbons calcined at 300 °C. The insert images of (a,d,g) are the low magnification SEM images of the samples.
Figure 2. SEM, TEM and selected area electron diffraction (SAED) images of (ac) dealloyed Al91.9Ce8Pt0.1, (df) Al91.7Ce8Pt0.3 and (gi) Al91.5Ce8Pt0.5 ribbons calcined at 300 °C. The insert images of (a,d,g) are the low magnification SEM images of the samples.
Nanomaterials 09 00683 g002
Figure 3. (a) Scanning transmission electron microscopy (STEM) image and element mapping of (b) Ce and (c) Pt in the Pt/CeO2 catalyst prepared from the dealloyed Al91.7Ce8Pt0.3 ribbon calcined at 300 °C.
Figure 3. (a) Scanning transmission electron microscopy (STEM) image and element mapping of (b) Ce and (c) Pt in the Pt/CeO2 catalyst prepared from the dealloyed Al91.7Ce8Pt0.3 ribbon calcined at 300 °C.
Nanomaterials 09 00683 g003
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) the Barrett–Joyner–Halenda (BJH) pore size distributions of dealloyed Al91.9Ce8Pt0.1, Al91.7Ce8Pt0.3 and Al91.5Ce8Pt0.5 nanorod frameworks calcined at 300 °C.
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) the Barrett–Joyner–Halenda (BJH) pore size distributions of dealloyed Al91.9Ce8Pt0.1, Al91.7Ce8Pt0.3 and Al91.5Ce8Pt0.5 nanorod frameworks calcined at 300 °C.
Nanomaterials 09 00683 g004
Figure 5. (a) CO conversion as a function of reaction temperature over the CeO2, Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts; (b) the Pt0.3/CeO2 catalyst prepared from dealloyed Al91.7Ce8Pt0.3 calcined at different temperatures. The inset image of (a) is the complete conversion curve for CeO2.
Figure 5. (a) CO conversion as a function of reaction temperature over the CeO2, Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts; (b) the Pt0.3/CeO2 catalyst prepared from dealloyed Al91.7Ce8Pt0.3 calcined at different temperatures. The inset image of (a) is the complete conversion curve for CeO2.
Nanomaterials 09 00683 g005
Figure 6. Long-term stability of the Pt0.3/CeO2 catalyst tested at 110 °C. (a) Successive reaction under a mixed composition of 1% CO, 10% O2 and 89% N2 for 75 h; (b) Successive reaction under mixed composition of 1% CO, 10% O2, 79% N2 and 10% H2O for 130 h, with addition of 5% CO2 after reaction for 70 h.
Figure 6. Long-term stability of the Pt0.3/CeO2 catalyst tested at 110 °C. (a) Successive reaction under a mixed composition of 1% CO, 10% O2 and 89% N2 for 75 h; (b) Successive reaction under mixed composition of 1% CO, 10% O2, 79% N2 and 10% H2O for 130 h, with addition of 5% CO2 after reaction for 70 h.
Nanomaterials 09 00683 g006
Figure 7. SEM image of the Pt0.3/CeO2 catalyst after reaction for 130 h.
Figure 7. SEM image of the Pt0.3/CeO2 catalyst after reaction for 130 h.
Nanomaterials 09 00683 g007
Figure 8. X-ray photoelectron spectroscopy (XPS) spectra of (a) the Ce 3d and (b) the Pt 4f peaks for the Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts.
Figure 8. X-ray photoelectron spectroscopy (XPS) spectra of (a) the Ce 3d and (b) the Pt 4f peaks for the Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts.
Nanomaterials 09 00683 g008
Figure 9. H2 temperature-programmed reduction (H2-TPR) profile of Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts.
Figure 9. H2 temperature-programmed reduction (H2-TPR) profile of Pt0.1/CeO2, Pt0.3/CeO2 and Pt0.5/CeO2 catalysts.
Nanomaterials 09 00683 g009
Figure 10. CO oxidation reaction mechanism on the Pt/CeO2 nanorod catalysts.
Figure 10. CO oxidation reaction mechanism on the Pt/CeO2 nanorod catalysts.
Nanomaterials 09 00683 g010
Table 1. Comparison of the catalytic performance of the nanorod framework-structured Pt/CeO2 catalysts with those of previous reports.
Table 1. Comparison of the catalytic performance of the nanorod framework-structured Pt/CeO2 catalysts with those of previous reports.
CatalystReaction Temperature (T99) (°C)Durability Test (h)Space Velocity (h−1)Reference
Porous/hollow structured Pt/CeO2@SiO2162Not stated60,000[44]
Nanorods CeO2-IMP-Pt901030,000[45]
Pt/CeO2 hollow sphere1551280,000[25]
Pt/CeO2 mesoporous sphere951280,000[14]
Nanorod framework-structured Au/CeO290Not stated60,000[32]
Nanorod framework-structured Pt0.3/CeO211313060,000This work

Share and Cite

MDPI and ACS Style

Wang, H.; Duan, D.; Ma, C.; Shi, W.; Liang, M.; Wang, L.; Song, X.; Gao, L.; Sun, Z. The Preparation and Catalytic Properties of Nanoporous Pt/CeO2 Composites with Nanorod Framework Structures. Nanomaterials 2019, 9, 683. https://doi.org/10.3390/nano9050683

AMA Style

Wang H, Duan D, Ma C, Shi W, Liang M, Wang L, Song X, Gao L, Sun Z. The Preparation and Catalytic Properties of Nanoporous Pt/CeO2 Composites with Nanorod Framework Structures. Nanomaterials. 2019; 9(5):683. https://doi.org/10.3390/nano9050683

Chicago/Turabian Style

Wang, Haiyang, Dong Duan, Chen Ma, Wenyu Shi, Miaomiao Liang, Liqun Wang, Xiaoping Song, Lumei Gao, and Zhanbo Sun. 2019. "The Preparation and Catalytic Properties of Nanoporous Pt/CeO2 Composites with Nanorod Framework Structures" Nanomaterials 9, no. 5: 683. https://doi.org/10.3390/nano9050683

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop