Putla Sudarsanam,
Baithy Mallesham,
D. Naga Durgasri and
Benjaram M. Reddy*
Inorganic and Physical Chemistry Division, CSIR – Indian Institute of Chemical Technology, Uppal Road, Hyderabad-500 607, India. E-mail: bmreddy@iict.res.in; mreddyb@yahoo.com; Fax: +91 40 2716 0921; Tel: +91 40 2719 1714
First published on 6th January 2014
The development of an efficient doped CeO2 material is an active area of intense research in environmental catalysis. In this study, we prepared highly promising Ce–Fe nano-oxides by a facile coprecipitation method and their catalytic performance was studied for CO oxidation. Various characterization techniques, namely, XRD, BET surface area, pore size distribution, Raman, FT-IR, TEM, H2-TPR, and XPS were used to correlate the structure–activity properties of the Ce–Fe catalysts. XRD results confirmed the formation of nanocrystalline Ce1−xFexO2−δ solid solution due to doping of Fe3+ into the CeO2 lattice. The BET surface area and lattice strain of CeO2 are significantly improved after the Fe-incorporation. Raman studies revealed the presence of abundant oxygen vacancies in the Ce–Fe sample. TEM images evidenced the formation of nanosized particles with an average diameter of 5–20 nm in the prepared samples. Interestingly, despite the thermal treatment at higher temperatures, the Ce–Fe sample showed remarkable reducible nature compared to pure CeO2 ascribed to existence of strong interaction between the CeO2 and FeOx. The synthesized Ce–Fe nano-oxides calcined at 773 K exhibited excellent CO oxidation performance (T50 = 480 K), with a huge difference of 131 K with respect to pure CeO2 (T50 = 611 K). The outstanding activity of the Ce–Fe catalyst is mainly due to smaller crystallite size, facile reduction, enhanced lattice strain, and ample oxygen vacancies. The superior CO oxidation performance of Ce–Fe nano-oxides with the advantages of low cost and easy availability could make them potential alternatives to noble metal-based oxidation catalysts.
Understanding the influence of various dopant features, such as oxidation state, ionic radius, dopant amount, redox nature, etc. on the structure–activity properties of the CeO2 would provide valuable implications towards the selection of an appropriate dopant.1,13,14 It has been shown that the substitution of Ce4+ by isovalent ions, like Zr4+ and Hf4+ enhances the OSC and redox properties of the CeO2 and thus, improved CO oxidation activity.15 Consequently, the Ce–Zr solid solutions have gradually replaced the pure CeO2 as efficient OSC materials in three-way catalytic converters to reduce the automotive exhaust pollutants (CO, NOx, and hydrocarbons).16 Nonetheless, a considerable disadvantage associated with isovalent dopants is the difficulty of self-generation of oxygen vacancies through charge neutrality criterion.17 Alternatively, the doping of aliovalent cations into the CeO2 lattice is found to be a versatile way of modulating the CeO2 characteristics.18 For example, trivalent dopants (e.g., Gd3+ and Sm3+) can promote the creation of large amount of extrinsic oxygen vacancies through charge compensation mechanism, along with intrinsic counterparts.19
It was reported that the similarity of ionic radii between the host (Ce4+) and guest ions (e.g., Eu3+ and Sm3+) is essential for the development of promising doped CeO2 catalysts.5,20 In contrast, metal ions (e.g., Fe3+ = 0.68 and Mn3+ = 0.72 Å) having huge ionic radii disparity with respect to Ce4+ (0.97 Å) were also studied as plausible dopants.10,21 On the other hand, the role of the dopant amount was also considered in the modification of CeO2 properties, indeed that strongly depends on the dopant nature.22 For instance, the introduction of 50% Zr (balance with Ce) leads to superior lattice oxygen mobility and facile reduction of Ce4+ to Ce3+ compared to other dopant amounts, whereas 20% La-incorporation shows better structural and redox properties than 50% Zr-incorporation.6,22 Interestingly, very small dopant amounts (≤10%) have also been employed for the preparation of smaller sized metal ion (Ni2+ and Co2+) doped CeO2 solid solutions.23 Especially, it is indispensable to mention here that the redox character of the dopant could exhibit a favourable role in the improvement of the CeO2 properties. It has been proved that the incorporation of tetravalent reducible dopants, such as Sn (Sn4+/Sn2+) and Ti (Ti4+/Ti3+) into the CeO2 framework significantly enhances its catalytic activity (NO removal by CO) in comparison to non-reducible Zr-dopant.13 Recently, Pintos et al.24 reported that the trivalent metal ion, such as Mn3+ can be easily reduced to lower oxidation state by release of oxygen, which leads to superior oxygen storage and release capacity during CO–O2 cycles. Zhang et al.25 proposed a redox mechanism, in which, the lattice oxygen from the Fe2O3 participates in soot oxidation by the reduction of Fe3+ to Fe2+ and the resulting oxygen vacancies are refilled by lattice oxygen of CeO2 through the reduction of Ce4+ to Ce3+. Therefore, it can be expected that the reducible dopant not only improves the structure–activity properties of the ceria, but also participates itself in the catalytic reactions.
As one of the promising reducible dopants, Fe (Fe3+/Fe2+) has received much attention due to its unique beneficial properties.21,26–29 Particularly, the redox potential of Fe3+/2+ (0.77 V) is very much lower compared to Ce4+/3+ (1.61 V).26 As well, Fe is inexpensive, abundantly available, and environmentally harmless.27 Hence, it can be believed that Fe-doped CeO2 oxides may be able to solve the challenges of environmental catalysis by combining the promising reactivity with the economic cost-effective performance. To understand this, we prepared Ce–Fe nano-oxides by incorporation of 10 mol% Fe into the CeO2 lattice by a facile coprecipitation method. The synthesized mixed oxides were calcined at different temperatures (773 and 1073 K) to explore their structural and textural stability. The catalytic efficiency of the Ce–Fe samples was studied for CO oxidation, a vital reaction in many industrial processes, such as automobile exhaust purification and CO preferential oxidation.1 An extensive physicochemical characterization was undertaken by means of X-ray diffraction (XRD), transmission electron microscopy (TEM) and high resolution electron microscopy (HREM), Brunauer–Emmett–Teller (BET) surface area, Barrett–Joyner–Halenda (BJH) pore size distribution, X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, Fourier transform infrared spectroscopy (FT-IR), and H2-temperature programmed reduction (TPR) techniques. Great efforts have been made to understand the structure–activity properties of the Ce–Fe catalysts. Also, the CO oxidation performance of the Ce–Fe nano-oxides was correlated with various non-reducible metal doped CeO2 oxides prepared by same preparation method.
The BET surface area, pore volume, and pore size distribution measurements were performed using a Micromeritics ASAP 2020 instrument. Prior to analysis, the samples were degassed at 573 K for 3 h to remove any residual moisture. The BET surface areas were calculated from adsorption data in the relative pressure (P/P0) range = 0.04–0.25. Pore size and pore volume were calculated using the BJH method applied to the desorption leg of the isotherms. Williamson–Hall plots (βcos
θ/λ vs. sin
θ/λ) were used to estimate the lattice strain of the catalysts. The TEM studies were carried out on a JEOL JEM-2100F instrument equipped with a slow-scan CCD camera, and the accelerating voltage of the electron beam was 200 kV. The preparation of samples for TEM analysis involved sonication in ethanol for 2–5 min followed by deposition on a copper grid. The FT-IR analysis was performed on a Nicolet 740 FT-IR spectrometer at ambient conditions using KBr discs with a nominal resolution of 4 cm−1 and averaging 100 spectra.
The vis-Raman analysis was carried out using a Horiba Jobin-Yvon HR800 Raman spectrometer equipped with a liquid-nitrogen cooled charge coupled device (CCD) detector and a confocal microscope. The emission line at 638 nm from Ar+ laser (Spectra Physics) 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 reported from the spectra are accurate to within 2 cm−1. The XPS measurements were carried out on a PHI 5400 instrument with a 200 W Mg-Kα probe beam. The analysis was done at room temperature and pressures were typically in the order of less than 10−8 Pa. The spectrometer was configured to operate at high resolution with a passing energy of 100 eV. Prior to the analysis, the samples were evacuated at high vacuum and then introduced into the analysis chamber. Survey and multi-region spectra were recorded at C 1s photoelectron peaks. Each spectral region of the photoelectron of interest was scanned several times to obtain good signal-to-noise ratios. Surface atomic ratios were calculated from peak area ratios normalized by the corresponding atomic sensitivity factors.
The H2-TPR analysis was conducted using an automated catalyst characterization system (Micrometrics model AutoChem 2910) with a thermal conductivity detector (TCD). Approximately, 100 mg of sample was loaded in an isothermal zone of the reactor and heated at a rate of 10 K min−1 to 573 K in 30 mL min−1 He gas, which facilitates desorption of the physically adsorbed water. After the sample was cooled to room temperature, He was switched to 10% H2/Ar with a rate of 20 mL min−1 and the temperature was linearly raised to 1073 K at a continuous heating ramp of 5 K min−1, keeping all other parameters unchanged. The reactor effluent gas was passed through a molecular sieve trap to remove the produced water and was analyzed by gas chromatograph (GC) using the TCD, and the reduction of CuO to metallic copper was used for the calibration purpose.
Noticeably, the XRD peaks of the CF773 sample are very much broadened compared to other samples, which reveal the formation of smaller sized CeO2 crystallites. The determined average crystallite sizes of the Ce–Fe and CeO2 samples by means of Scherrer equation are shown in Table 1. The crystallite size of CeO2 was significantly reduced after the Fe-incorporation, indicating the favourable role of Fe towards inhibition of the CeO2 crystal growth. However, with the increase of calcination temperature from 773 to 1073 K, the crystallite size of Ce–Fe sample was drastically increased. The average crystallite sizes of C773, C1073, CF773 and CF1073 samples were found to be ∼8.9, 32.5, 6.7 and 19.9 nm, respectively. The obtained specific surface area of the Ce–Fe and CeO2 samples are summarized in Table 1. It was found that the introduction of Fe into the CeO2 lattice considerably enhances its specific surface area from 41 to 68 m2 g−1 (Table 1). On the other hand, at elevated thermal treatments, the Ce–Fe sample exhibited very low specific surface area (15 m2 g−1) that might be due to penetration of the segregated FeOx species into the pores of the ceria. It is interesting to note from Table 1 that the Ce–Fe sample shows better textural properties (i.e., smaller crystallite size and higher specific surface area) compared to CeO2 even at higher calcination temperatures. Generally, the decrease of particle size results in the improved surface area of the material.32 Therefore, the higher specific surface area of the CF773 sample is most probably due to the presence of smaller sized CeO2 crystallites (Table 1).
TEM and HREM analyses were performed to explore the structural evolution of the prepared samples at atomic scale. It must be mentioned here that the calculation of particle size from TEM images of the metal oxides and mixed metal oxides, especially for ceria-based materials, prepared by precipitation methods is rather difficult as no protective agents are used like in the present study. As stated, no precise estimation of the particle size from the obtained TEM pictures was attempted in the present study (Fig. S1, ESI†). However, all samples exhibited nanosized particles in the broad range of 5–20 nm. A meticulous examination of the TEM images revealed that the 1073 K calcined samples show slightly larger sized particles in comparison to 773 K calcined samples. On the other hand, the HREM analysis of the CF773 sample (Fig. 2) clearly shows the lattice fringes with the spacing of ∼0.31 nm, which can be assigned to the CeO2 (111) with the XRD peak at 28.6° (Fig. 1). The estimation of lattice strain provides useful information about the defective nature of the CeO2-based materials.33 It was reported that doping of lower-valent cations facilitates defect formation within the CeO2 lattice by generating lattice strain, which eventually improves the extrinsic surface defects, resulting in superior CO oxidation performance.34,35 Williamson–Hall method was employed to determine the lattice strain of the CeO2 and Ce–Fe samples (Fig. S2, ESI†).33 It was found that the CF773 sample exhibits highest lattice strain (Table 1), which might be one of the reasons for its superior CO oxidation performance as discussed in the activity part. Interestingly, a negligible variation in the lattice strain values of the CF1073 and C773 samples was noticed, suggesting that no Fe-incorporation into the CeO2 lattice for the CF1073 sample, which is in consistent with the lattice parameter measurements (Table 1). Fig. 3 shows the N2 adsorption–desorption isotherms of the Ce–Fe samples. The obtained isotherms can be classified as Type IV isotherms with H1-type hysteresis, which are characteristic of the mesoporous materials.36 The determined data from the BJH model indicate narrow pore size distribution with an average pore diameter of 13.5 and 38.2 nm for the CF773 and CF1073 samples (Table 1), respectively (Fig. S3, ESI†). On the other hand, the CF773 sample shows larger pore volume (0.334 cm3 g−1) than CF1073 sample (0.055 cm3 g−1).
Raman spectroscopy was utilized to understand the structural properties of the Ce–Fe mixed oxides (Fig. 4). As can be seen from Fig. 4, all samples show a prominent band at around 445–460 cm−1, corresponding to the F2g vibration of the cubic fluorite type CeO2 that support the observations made from the XRD studies.8 The F2g band of Ce–Fe samples obviously shifted to lower wavenumbers with broadening compared to pure CeO2, which is more pronounced in the case of CF773 sample. The noticed peak shift and its broadening depends on various parameters, including the crystal defects, oxygen vacancies, phonon confinement, and inhomogeneous strain related to the reduced ceria.15,37 Generally, in trivalent metal doped CeO2 two types of oxygen vacancy bands are observed at around 545 and 600 cm−1 attributed to incorporation of trivalent metal ions into the CeO2 lattice and the presence of Ce3+ ions, respectively.8,9,19 It is well-known that the reduction of two adjacent Ce4+ ions to Ce3+ ions leads to generation of one oxygen vacancy in CeO2.38–40 On the other hand, the substitution of Ce4+ by Fe3+ results in the creation of an oxygen vacancy for every two Fe3+ dopants to balance the charge in the CeO2 lattice.9 From XPS studies (Fig. 5), it was clear that the Ce–Fe samples contain the Ce3+ ions, along with the Ce4+ ions. Therefore, the synthesized Ce–Fe samples are supposed to exhibit two types of oxygen vacancy bands due to Fe3+-incorporation and the presence of Ce3+ ions. However, to our surprise, the CF773 sample showed only one oxygen vacancy band at ∼600 cm−1 (Fig. 4, inset), corresponding to Ce3+ ion promoted oxygen vacancies. It has been reported that small doping amounts of Fe improve the oxygen vacancy concentration (vacancy compensation mechanism), whereas large amounts of Fe annihilate them (dopant interstitial compensation mechanism).28 Li et al.9 have showed that the oxygen vacancies generated due to the Fe3+-incorporation can be compensated by the presence of interstitial Fe3+ even for smaller doping amounts of Fe (15%). Moreover, it must be understood here that the observation of Raman band with relatively low oxygen vacancy concentration is extremely difficult in the case of CeO2-based oxides.41 Therefore, it can be suggested that the absence of oxygen vacancy band at around 545 cm−1 in the CF773 sample might be either due to the existence of interstitial Fe3+ ions or negligible concentration of the oxygen vacancies. Previous studies also reported only one oxygen vacancy band in the case of Ce–Fe mixed oxides.25,42
![]() | ||
Fig. 4 Visible Raman spectra of pure CeO2 (C773) and CeO2–Fe2O3 (CF) samples calcined at 773 and 1073 K. |
![]() | ||
Fig. 5 Ce 3d XPS profiles of pure CeO2 (C773, inset) and CeO2–Fe2O3 (CF) samples calcined at 773 and 1073 K. |
On the other hand, no oxygen vacancy band was found for the CF1073 sample. Usually, oxygen vacancies are generated due to the release of loosely bound oxygen ions from the octahedral interstitial sites, which are formed through relocation of the oxygen ions from the tetrahedral sites to the octahedral sites.40 At higher temperatures (∼1073 K), the oxygen ions can achieve sufficient energy to overcome the potential barrier to fill up the tetrahedral sites from the octahedral sites, which can hardly be removed from the lattice. As asserted, in the present study also, no visible oxygen vacancy band (Fig. 4, inset) was identified in the 1073 K calcined Ce–Fe sample. In addition to the F2g band, the CF1073 sample exhibited two bands at ∼216 and ∼280 cm−1, indicating the presence of α-Fe2O3 (hematite) phase with the D63d crystal space group.43 Interestingly, no α-Fe2O3 peaks were noticed in the CF773 sample, confirming the formation of CeO2 solid solution, and corroborating the XRD results very well.
XPS analysis was undertaken to determine the oxidation states of the elements present in the Ce–Fe samples as well as to understand the influence of the dopant (Fe) on the chemical environment of the cerium and oxygen. Fig. 5 shows the Ce 3d core level spectra of the CeO2 (inset) and CeO2–Fe2O3 samples. The obtained binding energies (BE) of the Ce 3d3/2 (u′′′), Fe 2p3/2 and O 1s are summarized in Table 2. The Ce 3d XP spectrum of CeO2-based materials is known to be complex due to the hybridization of Ce 4f with ligand orbitals and fractional occupancy of the valence 4f orbitals.6,8 This unnecessary hybridization assists the multiplet splitting of the peaks into doublets, with each doublet showing further structure due to final state effects.6 As shown in the Fig. 5, the curves of Ce 3d spectra are composed of ten peaks corresponding to five pairs of spin-orbit doublets. The peaks labelled ‘v’ indicates Ce 3d5/2 contributions and those of labelled ‘u’ represent the Ce 3d3/2 contributions. The dominant peaks denoted by v, v′′, v′′′, u, u′′and u′′′indicate the 3d104f0 state of Ce4+ ions, whereas those marked by v0, v′, u0 and u′ represent the 3d104f1 initial electronic state of the Ce3+ ions. Identification of these peaks reveals coexistence of Ce3+ and Ce4+ in all samples. Interestingly, the binding energy of the u′′′ peak was considerably decreased for the Ce–Fe samples with respect to the pure CeO2 (Table 2), indicating the influence of the dopant on the chemical environment of the CeO2.10 This noteworthy result suggests the improved reducible nature of the CeO2 in the Ce–Fe samples. Moreover, a meticulous observation of the Fig. 5 reveals that the intensity of the u′′′ peak (a qualitative fingerprint of Ce4+ ions) was increased with increase of calcination temperature, which signifies the presence of higher amount of Ce3+ ions in the CF773 sample compared to CF1073 sample.5,6 However, it must be mentioned here that the conversion of Ce4+ to Ce3+ is also possible under ultra-high vacuum conditions during the XPS measurements.44
The O 1s core-level spectra of CeO2 and Ce–Fe samples are shown in Fig. 6. Clearly, the O 1s spectrum of all samples is composed of more than one type of oxygen ions.22,45 The observed main peak at ∼530.0 ± 0.5 eV can be assigned to lattice oxygen of the CeO2. The appearance of another peak at ∼532.0 ± 0.3 eV reveals the existence of various hydroxyl and/or carbonyl groups on the catalyst surface, which are further confirmed by FT-IR study (Fig. S4, ESI†). Generally, the surrounding environment of the lattice oxygen in trivalent (e.g., Sm3+ or Gd3+) doped CeO2 could be different from that of pure CeO2 due to difference in the electronegativity of the dopant and cerium.46 In the present study also, the chemical environment of lattice oxygen was significantly changed after the Fe incorporation (CF773 sample) evidenced by shifting of lattice oxygen peak towards lower binding energy (Table 2). As asserted, this fascinating observation can be explained by a huge electronegativity disparity of the dopant (Fe ∼1.83) and Ce (1.12). Owing to lower binding energy, the lattice oxygen becomes more labile, which might be responsible for the generation of oxygen vacancies in the CF773 sample (Fig. 4). On the other hand, a negligible lattice oxygen peak shifting is observed in the case of CF1073 sample (Table 2), suggesting no influence of the Fe on the chemical environment of CeO2 lattice oxygen. This unusual observation is certainly due to Fe segregation at higher calcination temperatures (evidence from XRD and Raman studies). The binding energies of Fe 2p3/2 and Fe 2p1/2 are identified at ∼711.2 and ∼724.2 eV, with two satellite peaks at ∼718.7 and ∼732.4 eV, respectively, confirming the presence of Fe3+ in the Ce–Fe samples (Fig. S5, ESI†).25
![]() | ||
Fig. 6 O 1s XPS profiles of pure CeO2 (C773) and CeO2–Fe2O3 (CF) samples calcined at 773 and 1073 K. |
The surface Ce/Fe atomic ratios for all the samples are estimated from the XPS analyses of the Ce–Fe samples. If the Ce and Fe cations are randomly distributed throughout the sample, the expected Ce/Fe ratio must be 9. However, the obtained Ce/Fe atomic ratio for the CF773 sample was ∼9.356, which strongly deviated from the expected value attributed to segregation of Ce ions at the surface and/or enrichment of Fe ions in the bulk. Interestingly, with the increase of calcination temperature from 773 to 1073 K, there is a considerable decrease in the Ce/Fe atomic ratio (∼9.186), indicating the diffusion of Fe3+ ions to the catalyst surface. Therefore, there is a possibility for the segregation of Fe species at higher calcination temperatures, in agreement with the XRD and Raman studies of the CF1073 sample.
Fig. 7 shows the H2-TPR profiles of pure CeO2 and Ce–Fe samples calcined at different temperatures. Usually, pure CeO2 exhibits two reduction peaks at ∼793 (low-temperature, LT) and ∼1060 K (high-temperature, HT) due to reduction of surface-capping of oxygen and bulk-phase lattice oxygen, respectively.44 Due to lower enthalpy of reduction, the outermost layers of the CeO2 are first reduced at lower temperatures, followed by bulk reduction at higher temperatures.47 It is a well-known fact in the literature that the incorporation of a metal ion into the ceria lattice significantly enhances its reducible nature. As stated, the CF773 sample showed very much lower reduction temperatures (i.e., ∼620 and ∼925 K) in comparison to pristine ceria attributed to doping of Fe3+ ions into the CeO2 lattice. Interestingly, the CF1073 sample exhibited three reduction peaks at ∼656, 765 and 965 K, of which two peaks can be assigned to surface and bulk reduction of the CeO2 and the remaining peak might be due to the Fe2O3 species present in the CF1073 sample (evidence from XRD and Raman studies). Pristine Fe2O3 normally shows two step reduction, i.e., Fe2O3 → Fe3O4 and Fe3O4 → Fe(0) at ∼658 and ∼863 K, respectively.25,31,48 Therefore, by correlating the reduction temperatures of the Fe2O3 with that of the CF1073 sample, the observed reduction peak at ∼656 K in the CF1073 sample can be assigned to the Fe2O3 → Fe3O4 transition. On the other hand, no reduction peak related to the Fe2O3 species was noticed in the CF773 sample that might be due to the absence of Fe2O3 species. Besides having the low specific surface area, the CF1073 catalyst showed a better reducible nature than pure CeO2. This interesting observation can be explained by the existence of a strong interaction between the segregated Fe species and CeO2.49
The catalytic efficiency of the Ce–Fe samples was investigated for CO oxidation and the obtained results are shown in Fig. 8. It was found that the CO conversion increases with the increase of reaction temperature for all samples. The relative order of the catalyst activity can be evaluated by comparing the T50 values (T50 = temperature at which 50% CO conversion is achieved). Amongst, the CF773 sample exhibited remarkable CO oxidation performance at much lower temperature (T50 = 480 K), with a huge difference of 131 K in comparison to the pure ceria (T50 = 611 K). The enhanced lattice strain, facile reducible nature, ample oxygen vacancies, and improved BET surface area are found to be the decisive factors for high catalytic activity of the CF773 sample (Table 3). Despite its low surface area, the CF1073 sample showed comparable CO oxidation activity (T50 = 608 K) with that of CeO2 (T50 = 611 K). Ma et al.50 reported that FeOx species present on the Pt surface provide sufficient active oxygen species for CO oxidation. As mentioned in the introduction part, the lattice oxygen of the Fe2O3 also participates in soot oxidation through facile reduction of Fe3+ to Fe2+ and the generated vacancy can be refilled by CeO2 lattice oxygen.25 Hence, it can be expected that the Fe2O3 species present on the surface of CF1073 sample provides adequate active oxygen species for CO oxidation. The formed oxygen vacancies can be refilled by migration of the oxygen from the CeO2 lattice due to existence of strong interaction between CeO2 and Fe2O3 species (evidence from H2-TPR studies). Although no oxygen vacancies were found in the pure CeO2, its catalytic activity could be attributed to facile generation of oxygen vacancies under reaction conditions, followed by rapid and repeatable redox cycles of Ce4+/Ce3+.20
![]() | ||
Fig. 8 Oxidation of CO over pure CeO2 (C773) and CeO2–Fe2O3 (CF) samples calcined at 773 and 1073 K. |
Sample | BET surface area (m2 g−1) | Reduction peaksa | Lattice strainb | T50 (K)c | |
---|---|---|---|---|---|
LT | HT | ||||
a Determined from H2-TPR results (LT – low temperature and HT – high temperature).b Estimated from Williamson–Hall plots.c Obtained from CO oxidation activity results. | |||||
C773 | 41 | 793 | 1060 | 0.025 | 611 |
CF773 | 68 | 620 | 925 | 0.046 | 480 |
CF1073 | 15 | 765 | 965 | 0.028 | 608 |
Finally, we have made an attempt to correlate the performance (T50 values) of the Ce–Fe nano-oxides with that of Ce–Hf,15 Ce–Sm,5 Ce–La,44 and Ce–Zr15 catalysts prepared by same copreparation method (Fig. 9). It was found that the Ce–Fe sample exhibits 50% CO conversion at much lower temperatures compared to other doped CeO2 samples. An interesting observation that can be noted from Fig. 9 is that only Fe exhibits reducible nature, whereas Hf, Sm, La, and Zr are non-reducible dopants. Therefore, it can be suggested that the reducible nature of the dopant plays a beneficial role in the CO oxidation activity of the CeO2. The superior CO oxidation activity of Ce–Fe nano-oxides is expected to bring new opportunities in the design of efficient CeO2-based catalysts to replace the noble metal-based catalysts for industrially important oxidation reactions.
![]() | ||
Fig. 9 Correlation of CO oxidation performance (T50 values) of Ce–Fe, Ce–Hf, Ce–Sm, Ce–La, and Ce–Zr mixed oxides. |
Footnote |
† Electronic supplementary information (ESI) available: Pore size distribution, TEM, Williamson–Hall plots, XPS, and FT-IR profiles of Ce–Fe samples. See DOI: 10.1039/c3ra45778h |
This journal is © The Royal Society of Chemistry 2014 |