Damma Devaiahab,
Gode Thrimurthulua,
Panagiotis G. Smirniotisb and
Benjaram M. Reddy*a
aInorganic and Physical Chemistry Division, CSIR-Indian Institute of Chemical Technology, Uppal Road, Hyderabad-500 007, India. E-mail: bmreddy@iict.res.in; Fax: +91 40 2716 0921; Tel: +91 40 27193510
bChemical Engineering Program, Biomedical, Chemical, and Environmental Engineering, University of Cincinnati, Cincinnati, OH 45221-0012, USA
First published on 20th April 2016
In this work, alumina supported ceria–praseodymia (CP/A) samples were synthesized by a deposition coprecipitation method. The structural, textural, and redox properties of the prepared samples were characterized at different calcination temperatures from 773 to 1073 K and their catalytic activity was assessed in the CO oxidation reaction. In order to determine the promoting effect of the alumina support in the sample, the physicochemical and catalytic properties of CP/A were compared with unsupported ceria–praseodymia (CP) solid solutions. The X-ray diffraction results indicated the formation of ceria–praseodymia solid solutions over the alumina support. The nanocrystalline nature of the samples was confirmed by transmission electron microscopy. The CP/A sample showed an extremely high surface area which remained reasonably high even after calcination at 1073 K. The combined analyses revealed that the CP/A sample had more oxygen vacancies than CP. The H2-temperature programmed reduction results suggested that the active oxygens were significantly improved in CP/A over CP. The characterization results also highlighted the excellent thermal stability of CP/A. The CO oxidation profiles signified that the catalytic activity of CP/A calcined at 773 K was remarkably enhanced in comparison to that of CP. The fine dispersion of ceria–praseodymia solid solutions over the alumina support in the process of deposition coprecipitation and the synergistic effect between ceria–praseodymia and the support, which resulted in very high surface areas, oxygen vacancy concentrations, and active oxygen species, are believed to be responsible for the superior activity of the CP/A sample.
As is well-known, nanostructured ceria (CeO2) has proved to be highly efficient for CO oxidation and many other important catalytic reactions based on its distinctive oxygen vacancies and redox properties.8–10 In general, it is considered to be an essential catalyst support in environmental catalysis.11,12 Thus, the development of CeO2-based catalysts is strongly desirable not only for the reduction of pollutants, but also for other industrial applications. Tremendous efforts have been made to improve the catalytic performance of ceria by doping it with other metal cations.13–16
Especially, doped ceria with multivalent cations is attractive because of its enhanced catalytic properties.17,18 Among the various multivalent elements, praseodymium (Pr) with 3+/4+ states is expected to be the most appropriate for dissolution into the ceria matrix to form a Ce–Pr oxide solid solution, due to its analogous fluorite structural nature (Pr6O11) and because its ionic radius is close to that of Ce4+ ions.18–21 In addition, the presence of Pr in CeO2 is found to be propitious to enhance the formation of oxygen vacancies, resulting in excellent oxygen release ability (redox properties) due to its valence changeability from 3+ to 4+ or from 4+ to 3+. These admirable properties can strengthen CO adsorption on the surfaces and thus significantly activate the CO oxidation reaction from active centers around the vacancies in the Ce–Pr oxide solid solution.21–23 However, noticeable deactivation due to aggregation of these Ce–Pr oxide catalysts is inevitable during the reaction.21
Support materials with high surface areas have proved to be highly beneficial for improving the durability of catalysts, since the high specific surface area of the supports could contribute to the fine dispersion of the active components to avoid sintering.24,25 Among the various supports, alumina is considered to be outstanding, especially for CO oxidation, due to its ideal cost-efficiency, high and thermally stable surface area, relative inertness towards steam, and adhesive properties.26,27 In particular, γ-Al2O3 support has been extensively used to improve the thermal stability of ceria-based materials for three-way catalysis.13,28 Monte et al.29 demonstrated a strong correlation between the γ-Al2O3 support and the oxygen storage capacity (OSC) and thermal stability of supported nanostructured CemZr1−mO2 mixed oxides. Our group also previously reported superior OSC and CO oxidation activity over γ-Al2O3 supported CeO2-based oxides in comparison with unsupported oxides.30,31 Morikawa et al.32 found that the introduction of Al2O3 into Ce1−xZrxO2 primary particles can improve durability at high temperatures and the oxygen release rate of the materials. From these studies, it is clear that the presence of the Al2O3 support is very advantageous for overcoming the potential sintering of ceria-based catalysts during the reaction.
Accordingly, the present investigation aimed at preparing γ-alumina supported ceria–praseodymia (CP/A) solid solutions by a deposition coprecipitation method. The structural, textural, and redox properties of the prepared samples were determined by a combination of several characterization techniques, such as X-ray diffraction (XRD), inductively coupled plasma-optical emission spectroscopy (ICP-OES), transmission electron microscopy (TEM), Raman spectroscopy (RS), ultraviolet-visible diffuse reflectance spectroscopy (UV-vis DRS), H2-temperature programmed reduction (H2-TPR), BET surface area (BET SA), and X-ray photoelectron spectroscopy (XPS). The obtained results allowed us to study the effect of alumina addition as a support on the physicochemical properties of ceria–praseodymia; this was correlated with the catalytic activity of CP/A and was evaluated in the CO oxidation reaction.
The chemical analysis of the prepared samples was performed by inductively coupled plasma-optical emission spectroscopy (Thermo Jarrel Ash model IRIS Intrepid II XDL, USA) to confirm the respective concentrations of the elements in the system. For ICP analysis, approximately 50 mg of the sample was dissolved in a solution of 25 mL aqua regia and 475 mL distilled water. Then 10 mL of the above solution was diluted to 250 mL.
Transmission electron microscopy studies were carried out on a JEM-2100 (JEOL) instrument equipped with a slow-scan CCD camera at an accelerating voltage of 200 kV. Samples for TEM analysis were prepared by crushing the materials in an agate mortar and dispersing them ultrasonically in ethyl alcohol. Afterward, a drop of the dilute suspension was placed on a perforated-carbon-coated copper grid and allowed to dry by evaporation at ambient temperature.
The Raman spectra were obtained at room temperature using a LabRam HR800UV Raman spectrometer (Horiba Jobin-Yvon) fitted with a confocal microscope and a liquid-nitrogen cooled charge-coupled device (CCD) detector. The samples were excited either with the emission line at 325 nm from a He–Cd laser (Melles Griot Laser) or with the emission line at 632 nm from an Ar+ ion laser (Spectra Physics) which was focused on the sample under the microscope, with the diameter of the analyzed spot being ∼1 μm. The acquisition time was adjusted according to the intensity of the Raman scattering. The wavenumber values obtained from the spectra were precise to within 2 cm−1. The UV-vis DRS measurements were performed using a GBSCintra 10e UV-vis NIR spectrophotometer with an integration sphere diffuse reflectance attachment. BaSO4 was used as the reference, and the spectra were recorded in the range of 200 to 800 nm.
The reducibility of the catalysts was studied by H2-TPR analysis using a thermal conductivity detector with a gas chromatograph (Shimadzu). Prior to the reduction, approximately 30 mg of the sample was loaded in an isothermal zone of the reactor, pre-treated in a helium gas flow at 473 K and then cooled to room temperature. Then, the sample was heated at a rate of 10 K min−1 from ambient temperature to 1100 K in a 20 mL min−1 flow of 5% H2 in Ar. The hydrogen consumption during the reduction process was estimated by passing the effluent gas through a molecular sieve trap to remove the produced water; it was analyzed by gas chromatography using a thermal conductivity detector. The BET surface areas were determined by N2 physisorption at liquid N2 temperature on a Micromeritics Gemini 2360 instrument using a thermal conductivity detector (TCD). Prior to analysis, the samples were degassed at 393 K for 2 h to remove the surface adsorbed residual moisture.
The XPS measurements were performed on a Shimadzu ESCA 3400 spectrometer using Mg-Kα (1253.6 eV) radiation as the excitation source at room temperature. The samples were maintained in a vacuum, typically on the order of less than 10−8 Pa, to avoid a large amount of noise in the spectra from contaminants. The obtained binding energies were corrected by referencing the spectra to the carbon (C 1s) peak at 284.6 eV.
The lattice parameter (a) of the CP/A and CP 773 samples (see Table 1) is slightly larger than that of pure CeO2. The ionic radius of Pr4+ (0.096 nm) is slightly smaller than the Ce4+ radius (0.097 nm); therefore, if Ce4+ is replaced by Pr4+, no significant change in the lattice parameter is expected. The presence of Pr3+ ions in the ceria lattice will cause a lattice expansion because of the larger Pr3+ ionic radius (0.113 nm). The lattice expansion is also ascribed to the increased presence of oxygen vacancies conjugated with Ce3+ ions in the samples. Therefore, it can be concluded that the lattice expansion in the CP/A and CP 773 samples is due to the substitution of the parent Ce4+ cations by larger Pr3+ ions and the increased content of Ce3+ ions in the samples, which is consistent with the (111) peak shift to lower 2θ angles.21,33,34 Moreover, it is interesting to note that the lattice parameter of CP/A 773 is higher than that of the CP 773 sample, signifying the presence of more 3+ (Ce3+ and Pr3+) cations in the sample. Although the lattice parameter of the CP/A sample decreased with increasing calcination temperature, CP/A 1073 has a greater “a” value than CP 1073 and pure CeO2, demonstrating that a comparatively high number of 3+ ions are present in CP/A at 773 K and that they are relatively stable up to 1073 K. Therefore, it can be deduced that the alumina support enhances the number and thermal stability of 3+ ions in the CP/A sample, which is attributed to the synergistic interaction between ceria–praseodymia and the alumina support. These 3+ cations can play a crucial role in catalytic reactions by creating oxygen vacancies. On the other hand, the lattice parameter of the CP 1073 sample is equal to the pure ceria value, which may be due to the inferior interaction between CeO2 and Pr6O11 in CP at a higher calcination temperature.
Sample | SA (m2 g−1) | Da (nm) | aa (Å) | [Ce3+]/[Ce3+ + Ce4+]b | O 1s binding energyc (eV) | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
OL | OH | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
a Calculated from XRD analysis.b The surface relative molar ratio of [Ce3+]/[Ce3+ + Ce4+] calculated from the Ce 3d XPS spectra.c The values represent the O 1s binding energy obtained from the O 1s XPS spectra. | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
CeO2 | 41 | 7.7 | 5.415 | 0.25 | 530.3 | 532.4 | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
CP 773 | 72 | 7.1 | 5.430 | 0.43 | 530.6 | 533.0 | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
CP 1073 | 31 | 16.3 | 5.415 | 0.32 | 530.7 | 532.9 | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
CP/A 773 | 156 | 5.4 | 5.437 | 0.47 | 530.6 | 533.2 | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
CP/A 873 | 138 | 5.7 | 5.429 | 0.45 | 530.6 | 533.1 | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
CP/A 973 | 120 | 6.6 | 5.425 | 0.43 | 530.5 | 533.0 | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
CP/A 1073 | 107 | 8.1 | 5.420 | 0.38 | 530.5 | 532.9 |
The diffraction peaks of the CP/A samples are broader than that of CP, indicating the formation of smaller CeO2 crystallites in the CP/A sample. The crystallite size was calculated from the peak broadening of the (111) diffraction peak using the Debye–Scherrer equation. The average crystallite sizes of CP/A 773 and CP 773 were estimated to be 5.4 and 7.1 nm, respectively. The lower crystallite size of CP/A 773 was likely to be caused by the good dispersion of ceria–praseodymia solid solution over the support. With increasing calcination temperature from 773 to 1073 K, the crystallite size of CP was increased by 2.3 times, while it is enhanced by only 1.5 times in the case of the CP/A sample. This result again clearly indicated that the Al2O3 support serves as the thermal diffusion barrier and effectively inhibits the growth of ceria crystallites in CP/A during calcination, resulting in the lower CeO2 crystallite size for CP/A 1073 compared to CP 1073. These results are in line with the BET surface area values shown in Table 1. The molar ratios of Ce, Pr, and Al were determined by elemental analysis and were found to be very close to the nominal values for all CP/A samples (Table S1, ESI†), indicating the complete precipitation of Ce and Pr cations over the support during the preparation process.
![]() | ||
Fig. 2 TEM images of ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP) samples calcined at 773 K and 1073 K (inset: encircled views of selected areas). |
![]() | ||
Fig. 3 HREM images of ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP) samples calcined at 773 K and 1073 K (inset: enlarged views of selected areas). |
The UV and visible Raman spectra of the samples, presented in Fig. 4, show the presence of two characteristic peaks which arise solely from the cubic CeO2 fluorite phase. The most prominent feature in the spectra is the F2g mode characteristic of a cubic fluorite crystal structure, which is positioned around 466.2 and 465.5 cm−1 in the UV and visible Raman spectra of pure CeO2, respectively. Interestingly, the F2g mode in the CP/A and CP samples experiences a red shift and broadening (in both the UV and visible Raman spectra) in comparison with pure ceria. This is due to the lattice expansion with the increased concentration of 3+ (Pr3+ and Ce3+) ions in the samples. Therefore, the shift and broadening of the F2g mode in the samples confirms that Pr has been incorporated into the ceria lattice, forming ceria–praseodymia solid solutions. In addition, the broader F2g peaks in CP/A and CP suggest a lower CeO2 crystallite size in the samples compared to pure ceria. These results are in agreement with the XRD data. In addition to the F2g peak, the other feature at a higher wave number (∼568 to 600 cm−1) is usually ascribed to oxygen vacancies (OV) in the samples. According to the literature, this mode originates from the existence of both intrinsic and extrinsic oxygen vacancies in the samples.19 The intrinsic oxygen vacancies are created in the samples due to the presence of Ce3+ ions, whereas the extrinsic oxygen vacancies are introduced into the ceria in order to maintain charge neutrality when Ce4+ ions are replaced by Pr3+ ions. For every two substituted Pr3+ ions, one O2− leaves the crystal lattice.
![]() | ||
Fig. 4 (A) UV (B) visible Raman spectra of ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP) samples calcined at different temperatures along with pure ceria (CeO2) calcined at 773 K. |
For all the samples, the UV Raman spectral features are quite different from those in the visible Raman spectra in terms of the relative band intensities, which results from the absorption properties of the samples. The F2g band is weak in the UV Raman spectrum due to the strong absorption of the samples at 325 nm, while it is strong in the visible Raman spectra owing to the weak absorption at 632 nm. Considering the characteristic band of the oxygen vacancy (OV), it is strong in the UV Raman spectra but weak in the visible Raman spectra. Evidently, UV Raman spectroscopy is more sensitive to oxygen vacancies present in the samples than visible Raman spectroscopy. This is due to the resonance Raman effect,38,39 since the samples strongly absorb in the UV region. The ratio of the integrated peak area of the oxygen vacancies (AOV) to that of the main peak (AF2g), defined as AOV/AF2g, is used here to characterize the relative amount of oxygen vacancies among the samples.37,38,40 Fig. 5 illustrates the AOV/AF2g values of the samples in the UV and visible Raman spectra. As can be observed in this figure, the AOV/AF2g value is much higher in the UV region than that in the visible region in all samples, which is due to the fact that the vacancies are enriched on the surface of the samples.38,41 In both the spectra, the AOV/AF2g value follows the sequence: CP/A 773 > CP 773 > CP/A 873 > CP/A 973 > CP/A 1073 > CP 1073 > CeO2, suggesting that CP/A and CP samples have higher oxygen vacancies than pure ceria. This could be due to the enhanced 3+ ions in the samples by the synergistic interaction between ceria and praseodymia. Interestingly, CP/A 773 possesses a greater number of oxygen vacancies in comparison to CP 773. It is clear that the 3+ ions were further augmented (as evidenced from the XRD results); therefore, there are larger oxygen vacancies in CP/A 773 than in CP 773, which is mainly attributed to the increased synergistic effect between CeO2 and Pr6O11 over the Al2O3 support. More oxygen vacancies may result in high mobility of activated oxygens, which are crucial to the catalytic performances of the samples. Moreover, it is found that the AOV/AF2g ratio decreases obviously in both the UV and visible Raman spectra for the CP/A and CP samples as the calcination temperature increases from 773 to 1073 K. This result implies that the relative concentration of oxygen vacancies decreases in the samples with increasing calcination temperature. Especially, the decreasing trend of the oxygen vacancies is more pronounced in the visible Raman spectra than in the UV Raman spectra during the calcination process, indicating that the vacancies in the surface region are more stable than in the bulk of the samples. However, the CP/A sample showed more vacancies than the CP sample at 1073 K in both the UV and visible Raman spectra. This observation confirms the enhanced thermal stability of the oxygen vacancies in CP/A, which is again due to the synergistic interaction between ceria–praseodymia and the alumina support.
![]() | ||
Fig. 6 UV-vis DR spectra of ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP) samples calcined at 773 and 1073 K along with pure ceria (CeO2) calcined at 773 K. |
![]() | ||
Fig. 7 H2-TPR profiles of ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP) samples calcined at 773 and 1073 K along with pure ceria (CeO2) calcined at 773 K. |
An important observation in the TPR patterns is that the surface reduction temperature of CP/A is significantly lower than that of the CP sample. As the support is hardly reducible in the explored temperature range, the H2 consumption in the CP/A samples is mainly attributed to the reduction of Ce4+ and Pr4+ ions. Hence, the reduced surface reduction temperature of CP/A should be attributed to the increased synergistic interaction between the Ce4+/Ce3+ and Pr4+/Pr3+ redox couples on the surface due to the high dispersion of ceria–praseodymia solid solutions over the alumina support, which is corroborated by the surface area data. Notably, the greatly decreased surface reduction temperature of CP/A 773 indicates that it has a higher number of active oxygen species compared with the other samples. This enhancement of the active oxygen species in the CP/A 773 sample is strongly related to its highest number of oxygen vacancies (confirmed from the Raman and UV-visible DR spectral results) among the samples, which could play a crucial role in catalytic oxidation reactions. Thus, it can be expected that the CP/A 773 sample will have high activity in oxidation catalysis. On the other hand, the CP/A sample has a higher bulk reduction temperature than the CP sample. The strong interaction between ceria–praseodymia and the alumina support in the CP/A sample could be responsible for this result.
![]() | ||
Fig. 8 XPS spectra of (A) Ce 3d, (B) O 1s for ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP) samples calcined at different temperatures, along with pure ceria (CeO2) calcined at 773 K. |
The relative amount of Ce3+ ions in the samples can be estimated by the [Ce3+]/[Ce3+ + Ce4+] ratio, and the obtained values are shown in Table 1. It can be clearly seen that the relative concentration of Ce3+ ions for the samples increased in the order CeO2 < CP 1073 < CP/A 1073 < CP/A 973 < CP/A 873 < CP 773 < CP/A 773. That is to say, the CP/A and CP samples have a higher relative amount of Ce3+ ions than CeO2. This observation demonstrates that the formation of ceria–praseodymia solid solution by the doping of Pr ions into the CeO2 lattice could facilitate the generation of more surface oxygen vacancies in the samples. Notably, the CP/A 773 sample is found to have the highest amount of Ce3+ ions on the surface; therefore, it has the most oxygen vacancies among the samples. This result suggests that not only the substitution of the CeO2 lattice by Pr ions but also the alumina support could further increase the quantity of the oxygen vacancies in the CP/A 773 sample due to its synergistic effect with ceria–praseodymia, which can change the electronic structure of the Ce ions and thus alter the coordination between the Ce and oxygen ligands.48,49 Additionally, with increasing calcination temperature, the content of Ce3+ ions, and therefore the oxygen vacancies, decreased in both the CP/A and CP samples. However, CP/A 1073 had a higher concentration of Ce3+ ions compared to CP 1073. Hence, it could be concluded that the thermal stability of CP/A is obviously improved by the presence of the Al2O3 support. These observations are in good agreement with our previous Raman, UV-vis DRS, and H2-TPR results.
The XPS spectra of the CP/A, CP, and CeO2 samples in the O 1s region are displayed in Fig. 8(B), and the corresponding binding energy values are demonstrated in Table 1. The O 1s spectra for all samples are asymmetric due to the non-equivalence of the surface oxygen ions, which causes the spectra to split into two peaks. This result clearly illustrates that the two kinds of oxygen species are present on the surface of all the samples. The peak labeled with OL at a lower binding energy (530.3 to 530.7 eV) is characteristic of the lattice oxygen (O2−) for metallic oxides in the samples, while the other sub-band labeled with OH at a high binding energy (532.4 to 533.2 eV) is attributed to the surface chemisorbed oxygen, such as O22− or O− attributed to oxide defects, surface oxygen ions with low coordination or surface hydroxyl and/or carbonate impurities.50,51 Meanwhile, it can be noted from Table 1 that the binding energies (OH peak) of CP/A and CP are slightly shifted to higher values compared with CeO2. The increase in the binding energy of the OH peak in the CP/A and CP samples is ascribed to their greater surface oxygen vacancy concentration compared to pure CeO2.52,53
Fig. 9(A) illustrates the Pr 3d core level XPS spectra of the CP/A and CP samples. Usually, the Pr 3d spectrum is characterized by two sets of 3d3/2 and 3d5/2 spin–orbit multiplets. However, the peaks corresponding to the 3d3/2 sublevel are very complex in nature due to the multiplet effect, whereas the 3d5/2 sublevel exhibits only two peaks, which are related to the possible 3+ and 4+ oxidation states of Pr.54,55 As a result, we studied only the 3d5/2 region of the Pr 3d spectrum to determine the oxidation state of Pr in the samples. The Pr 3d5/2 spectra of the samples show two prominent peaks at ∼930.7 and ∼935.3 eV, which correspond to Pr3+ and Pr4+ ions, respectively.23,55 This clearly indicates that the Pr ions display both 3+ and 4+ oxidation states in both the CP/A and CP samples at the surface region. Moreover, the intensity of the Pr4+ ion peak is greater than that of Pr3+, indicating the greater concentration of Pr4+ ions on the surface of all samples. However, the figure clearly demonstrates that the calcination temperature has no significant effect on the intensity of the Pr 3d5/2 spectrum in both the CP/A and CP samples. It is interesting to note that the binding energies of the CP/A samples are higher than that of CP, which is probably due to the interaction between ceria–praseodymia and the alumina support in the samples.
![]() | ||
Fig. 9 XPS spectra of (A) Pr 3d5/2 for ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP). (B) Al 2p for ceria–praseodymia/alumina (CP/A) samples calcined at different temperatures. |
The Al 2p XPS spectra of the CP/A samples are shown in Fig. 9(B). The Al 2p peaks for all samples are observed to be centered at ∼76 eV, which is slightly higher than the values previously reported for Al2O3.56,57 This shift could be ascribed to the strong interaction between ceria–praseodymia and the alumina support, which is also supported by the XPS spectra of Ce 3d and Pr 3d5/2 as discussed earlier.58 In addition, the binding energy values of Al 2p do not vary with calcination temperature, suggesting the presence of one type of aluminium oxide (γ-alumina) with an oxidation state of 3+ at all calcination temperatures.
![]() | ||
Fig. 10 Catalytic activity of ceria–praseodymia/alumina (CP/A) and ceria–praseodymia (CP) samples calcined at 773 and 1073 K along with pure ceria (CeO2) calcined at 773 K for CO oxidation. |
The difference in the CO oxidation activity of the samples is better understood from the characterization results. Usually, the surface area of the catalyst plays an important role in the oxidation of CO; a larger surface area can expose more active sites for the oxidation reaction, which is advantageous to increase the catalytic activity.42 Obviously, the CP/A 773 sample with the highest surface area of 156 m2 g−1 demonstrated the best activity among the samples. However, the CP 1073 sample shows higher catalytic activity by 25 K, even though its surface area is smaller than that of pure CeO2. This finding indicates that the surface area does not solely determine the catalytic performance, but that additional factors such as oxygen vacancies and active oxygen species may play a vital role in the activity.15,59 Interestingly, the order of the AOV/AF2g (Raman results) values, absorption edges (UV-vis DRS), and relative quantities of Ce3+ ions (XPS spectra) in the samples exactly coincided with the order of the CO oxidation activity. This reflects that the oxygen vacancies play a major role in the CO oxidation reaction. Generally, the CO oxidation occurs by CO + [M − O]* → CO2 + M, where [M − O]* is an active oxygen species. The active oxygen site is notably favoured by the presence of oxygen vacancies that improve the surface dynamics.60 According to the H2-TPR results, the CP/A 773 sample had more active oxygen species than the other samples. Hence, it should be concluded that the CP/A 773 sample with highest number of oxygen vacancies, which in turn increases the number of active oxygen species, showed the most remarkable activity for CO oxidation among the investigated samples. Overall, the combined results illustrate that the surface area, oxygen vacancies, and active oxygen species are the key factors that influence the catalytic performance of the present samples.
Further, the catalytic activity of the present CP/A 773 sample was also compared with our earlier reports of Al2O3 supported samples. Surprisingly, CP/A 773 shows a lower T50 than Ce0.8Tb0.2O2/Al2O3 (CT/A),31 Ce0.8Hf0.2O2/Al2O3 (CH/A),31 Ce0.5Zr0.5O2/Al2O3 (CZ/A),31 and Ce0.8La0.2O2/Al2O3 (CL/A)48 samples calcined at 773 K. In particular, at the T50 of CP/A 773, the CT/A, CH/A, CL/A, and CZ/A samples exhibit only 31%, 27%, 23%, and 20% conversions, respectively (Table 2). Evidently, CP/A 773 showed the best performance for CO oxidation among the various alumina supported ceria-based mixed oxides.
To examine the stability of the CP/A 773 catalyst in the CO oxidation reaction, cycling tests were performed over the sample. Fig. S1 (ESI†) illustrates the T50 values over the CP/A 773 sample for five successive runs. It can be clearly observed that the T50 of the sample remained almost constant up to five cycles. This result indicates the good stability of the CP/A 773 sample for long-term CO oxidation reactions. Moreover, a UV Raman spectrum was also acquired for the CP/A 773 sample after the cycling test; this is shown in Fig. S2 (ESI†). Quite impressively, no significant loss of oxygen vacancies was observed over the spent CP/A 773 catalyst. This finding provides further evidence for the stability of the CP/A 773 sample. Therefore, it can be assumed that the present CP/A 773 sample is the most promising catalyst for CO oxidation and may be envisioned for use in various other catalyzed oxidation reactions; thus, its ultimate industrial applications are also promising.
Footnote |
† Electronic supplementary information (ESI) available: Chemical composition, cycling test of CP/A 773 for CO oxidation, and UV Raman spectra of spent CP/A 773 sample. See DOI: 10.1039/c6ra06679h |
This journal is © The Royal Society of Chemistry 2016 |