Next Article in Journal
Photocatalytic H2O2 Generation Using Au-Ag Bimetallic Alloy Nanoparticles loaded on ZnO
Next Article in Special Issue
Fenton-like Remediation for Industrial Oily Wastewater Using Fe78Si9B13 Metallic Glasses
Previous Article in Journal
Late Transition Metal Catalysts with Chelating Amines for Olefin Polymerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Temperature Abatement of N2O over FeOx/CeO2-Al2O3 Catalysts: The Effects of Oxygen Mobility

1
Boreskov Institute of Catalysis SB-RAS, 5, Lavrentiev ave, Novosibirsk 630090, Russia
2
Laboratory of Optical Materials and Structures, Institute of Semiconductor Physics, SB-RAS, Novosibirsk 630090, Russia
3
Research and Development Department, Kemerovo State University, Kemerovo 650000, Russia
4
Department of Industrial Machinery Design, Novosibirsk State Technical University, Novosibirsk 630073, Russia
5
R&D Center “Advanced Electronic Technologies”, Tomsk State University, Tomsk 634034, Russia
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(9), 938; https://doi.org/10.3390/catal12090938
Submission received: 4 August 2022 / Revised: 19 August 2022 / Accepted: 22 August 2022 / Published: 24 August 2022

Abstract

:
CeO2-Al2O3 oxides prepared by co-precipitation (Ce+Al) or CeOx precipitation onto Al2O3 (Ce/Al) to obtain dispersed CeO2 and samples with further supported FeOx (2.5–9.9 weight% in terms of Fe) were characterized by XRD, XPS, DDPA and Raman. Fe/Ce/Al samples with lower surface concentrations of Fe3+ were substantially more active in N2O decomposition at 700–900 °C. It was related to higher oxygen mobility, as estimated from 16O/18O exchange experiments and provided by preferential exposing of (Fe-)Ce oxides. Stabilization of some Ce as isolated Ce3+ in Fe-Ce-Al mixed oxides dominating in the bulk and surface layers of Fe/(Ce + Al) samples retards the steps responsible for fast additional oxygen transfer to the sites of O2 desorption.

Graphical Abstract

1. Introduction

Nitric acid is a key chemical with annual production at above 65 million tons mainly driven by the fertilizer industry. Nitrous oxide (N2O) is produced as an unwanted, but inevitable by-product at HNO3 industrial manufacture by ammonia oxidation. Its emissions by nitric acid plants are growing from 5 to 12 kg N2O/ton HNO3 according to the tendency to high-pressure processes [1]. Meanwhile, nitrous oxide is a potent greenhouse gas characterized by about 300 times the Global Warming Potential of CO2 and contributes to the greenhouse effect and ozone layer depletion. Therefore, 330–780 thousand tons of N2O annually emitted at HNO3 production are equivalent to 100–240 million tons/y CO2e. It can be deleted by placing the catalysts in several points of the ammonia processing flowsheet [2], including that immediately downstream the Pt–Rh gauzes operating at 800–900 °C. A tendency of recent years to increase weight space velocity and reduce the charge of platinum-group metals in the reactor increased the risk of ammonia slipping at gauze ageing or damage. Therefore, efficient conversion of NH3 to NO + NO2, while N2O is low is one of the necessary features for N2O decomposition catalysts operating in these conditions as well.
Fe-Al-O-based catalytic compositions are efficient both in selective ammonia oxidation to NO and NO2 [3,4,5] and in selective N2O decomposition under the conditions of the nitric acid plant [4,5,6,7,8] without decrease in NOx yield [6,7]. The activity in both reactions was shown to increase by using CeO2 as the support instead of alumina [4]. It was due to the appearance of the additional pathway of the oxygen supply through oxygen vacancies located in the near subsurface layers of ceria to Fe active sites facilitating the rate-determining step of O2 desorption. An increase in both ceria and Co(Fe)Ox dispersion could additionally promote oxygen diffusion to Fe. In the first case it is because both distributions of oxygen vacancies and oxygen storage and release capability of CeO2 are related to its particle size [9,10]. The positive effect of the dispersion of Co oxide spinel phase [11,12] or FeOx species over ceria and alumina [4,5,8] onto activity in both reactions was already shown and related to enhanced oxygen diffusion through the active component/support interface periphery [4,11,12]. However, for individual CeO2 all positive effects resulting from fast oxygen transfer are restricted by its low surface area under the reaction conditions [10]. Several methods to stabilize dispersed CeO2 particles over Al2O3 after calcination at above 800 °C were reported. They include the sol–gel method using aluminum tri-sec-butoxide and Ce(III) nitrates as precursors [13], impregnation of alumina with an aqueous solution of (NH4)2[Ce(NO3)6] [14] or stabilization of a complex of the CeO2 precursor in a homogeneous polymeric matrix (Pechini route) followed by supporting onto alumina [15,16].
In the current study, we used CeO2 precipitation onto Al2O3 or CeO2 and Al2O3 co-precipitation methods to increase CeO2 dispersion in the final supports compared with individual CeO2. The reasons for the principally different activity of FeOx/CeO2-Al2O3 samples prepared using these supports have been elucidated.

2. Results and Discussion

2.1. Phase Composition (XRD)

XRD patterns of Ce/Al and (Ce + Al) samples revealed a mixture of CeO2, θ-Al2O3 and γ-Al2O3, phases (Figure S1 in Supplementary Materials). Lattice parameters of CeO2 in both samples were exactly the same (Table 1) and practically coincided with those for pure CeO2 calcined at 900 °C (a = b = c = 5.412 Å), while the values of coherent scattering area dXRD calculated by the Scherrer equation were lower (40 and 11 nm for Ce/Al and (Ce + Al), respectively, compared with 68 nm for pure CeO2 [4]). Lattice parameters of γ(+θ)-Al2O3 phases cannot be calculated correctly because of low intensity and substantial broadening of corresponding overlapping signals. Therefore, some penetration of Ce ions into their lattices cannot be excluded. At the same time, higher dispersion of alumina species in the (Ce + Al) sample follows from larger widths of the peaks at 45.5° and 67.0° (Figure S1).
Fe deposition resulted in a decrease in SBET value and a consistent increase in DXRD of CeO2 species; the value of effect depended on the quantity of Fe (Table 1). Although CeO2 lattice parameters did not change for all Fe content ranges, the formation of Fe-Ce-O solid solution cannot be excluded. Ceria lattice shrinkage was measured in Au/Ce-Fe-O (Ce/Fe = 9) [17], Fe/CeO2 (Ce/Fe = 4–9) [18] or CeO2–Fe2O3 (Ce/Fe = 2.3) [19] samples prepared by procedures favoring homogeneous distribution of Fe and Ce in the samples before their high temperature treatment. It was due to the replacement of Ce4+ ions in the typical cubic fluorite lattice (radius 0.097–0.101 nm) by Fe3+ ions (0.049–0.078 nm, depending on coordination and spin state). However, at smaller Fe content in FexCe1-xO2 (x ≤ 0.05, or Ce/Fe ≥ 20) even a slight increase in lattice parameter compared with pure CeO2 was observed [18] and related to partial Ce4+ reduction to the larger (0.123 nm) Ce3+ ion [20]. Although in our case Ce/Fe ratio varies within 1.1–6.2, these differently directed effects could compensate for each other because Fe was supported onto support already calcined at 900 °C diminishing the efficiency of its incorporation into CeO2 lattice, while incorporation of some Fe cations into the alumina component is possible as well.
Nevertheless, neither relative intensity and dispersion of γ-,θ-Al2O3 changed (Figure S1, inserts), nor α-Al2O3 appeared in both Fe/Ce/Al and Fe/(Ce+Al) samples (Figure S1), unlike Fe/Al2O3 samples with reasonably close SBET and Fe content values [4]. Hence, the formation of Fe-Al-O solid solutions in Ce-containing samples is hardly ever possible.

2.2. Catalytic Activity in N2O Decomposition and NH3 Oxidation

The activity of Ce/Al and (Ce + Al) based catalysts expressed in terms of N2O conversion at 800 °C versus Fe content (Figure 1A) is typical for temperature range 700–900 °C (Figure S2). For every type of support, it increases with Fe content. O2 and H2O expectedly retard the reaction (Figure S3). Conversion values measured for Fe/(Ce + Al) samples are evidently lower than for Fe/Ce/Al. Therefore, better dispersion of CeO2 species at close Ce content in the sample does not guarantee higher activity even at higher SBET values (Table 1).
For all samples, the conversion of NH3 and yield of (NO + NO2) (YNO+NO2) increase, while that of N2O (YN2O) decreases with temperature; YNO+NO2 = 74–80% and YN2O = 0.5–1.8% at T ≥ 800 °C (Figure S4A,B, and Figure 1B). High stability for 8 h of operation, at least, was shown as well (Figure S4C). We estimated that at an NH3 slip of 1% (although such a value is practically impossible) the increment of N2O concentration after the Fe-Ce-Al catalyst at 800 °C will be 65–90 ppm. At the average N2O concentration downstream of the fresh gauzes of 1500 ppm [2], such an addition will diminish N2O conversion only by 4–6%.

2.3. Microstructure and Morphology

2.3.1. DDPA

Detailed analysis of dissolution kinetics of Fe, Ce and Al in different samples has been presented in corresponding section of Supporting Information. Herein, we summarized and discussed the main facts that are important for understanding the composition and spatial distribution of different compounds in Fe/(Ce + Al) and Fe/Ce/Al samples.
Mixed oxides of different compositions dissolving consecutively in either HCl or HF (Figure S6) represent most of the 2.5 Fe/(Ce + Al) sample (Table 2). The most abundant compound is Fe-Ce-Al mixed oxide with Fe0.03Ce0.09Al stoichiometry which (Figure 2A): encapsulates the particles of Fe-Ce mixed oxide (Fe0.02Ce stoichiometry); is partially screened by more disordered Fe-Ce-Al mixed oxide with higher content of Fe and Ce (Fe0.05Ce0.26Al stoichiometry) and Fe-Al-O compounds dissolving in HCl.
The decreasing Fe/Al ratio in Fe-Al-O compounds points to the formation of mixed oxide with most disordered near-surface layers enriched by Fen+ (Figure S6B). At the same time, the protocol of Fe supporting includes using of acidic solutions giving rise to the dissolution of some poorly crystallized alumina detected in the initial support (Figure S5A, Table 2). In this case, the formation of a γ-FeAlO3-like compound with spinel structure is quite possible after following drying and calcination at T < 400 °C [21]. Since the mutual Fe2O3 and Al2O3 solubility in such structure decreased prominently with calcination temperature rise, its decomposition could happen in our case as well leading to the formation of highly dispersed FeOx species (0.06 mmol/g) on the surface of Fe-Al mixed oxide with Fe0.05Al stoichiometry. Its quantity (1.3 mmole/g) is reasonably close to that of alumina dissolving in HCl in the initial support (1.15 mmole/g) (Table 2). Calculation showed that identified soluble compounds include all Al and Fe in the sample, while about 0.7 mmol/g of Ce, or 25% of its general content in the sample, represent insoluble oxides (Table S1 in Supplementary Materials, Table 2) which are, more probably, small sized CeO2 species with unmodified fluorite lattice (Table 1). Their spatial location in the sample is not clear, but encapsulation by Ce-Al mixed oxide preventing Fe incorporation into the lattice is quite reasonable because the same was detected for the particles of Fe-Ce mixed oxide.
The overall quantity of soluble Ce-containing compounds in the 3.8 Fe/Ce/Al sample is substantially lower as compared with the 2.5 Fe/(Ce + A) sample (Table 2), which can be related to a larger size of CeO2 particles (Table 1). Most of the dissolved part of the sample is represented by Fe-Al mixed oxide partially screened by Fe-Ce-Al one with Fe0.04Al and Fe0.08Ce0.2Al stoichiometries, correspondingly (Figure S7C and Figure 2B, Table 2). Surface layers are composed of more dispersed/disordered binary Fe-Al and Fe-Ce oxides and some FeOx dissolved in HCl (Figure S7, Table 2). Completely independent kinetics of their dissolution can be in the case dispersed FeOx species are located on the surface of any of binary oxides and even non-dissolved large CeO2 species with unmodified fluorite lattice including 2.0 mmol/g of Ce, or more than 70% of its content in the sample (Figure 2B, Table 2). A substantially higher quantity of Fe in the surface located Fe-Al-O compounds (Fe0.29Al stoichiometry) compared with that in 2.5 Fe/(Ce + Al) could be due to other mechanisms of their formation compared with that in 2.5 Fe/(Ce + Al) sample (see below).

2.3.2. Raman

The band at 462.9 cm1 in the FT-Raman spectra of Ce/Al sample (Figure 3A) is typical for the F2g vibration mode of the fluorite structure of CeO2 [22,23]. Its width at ~24.5 cm−1 (Table 3) should correspond to a size of CeO2 species at about 10 nm [23,24] which is substantially less than the estimated value of DXRD at 40 nm (Table 1). At the same time, the particles of 10 nm in a size should be characterized by the Raman band at ~448 cm−1, as follows from experimental dependence obtained for CeO2 and CeO2-Al2O3 composites [15]. Spanier et al. [24] showed that lattice strains in the smaller nanoparticles have the largest contribution to both Raman peak position shift and its broadening compared with other reasons (phonon confinement, particle size distribution, defects, phonon relaxation). Tsunekawa et al. [25] attributed the formation of additional strains to the reduction of Ce4+ ions to Ce3+ ions caused by an increasing molar fraction of oxygen vacancies. We supposed that in our case stabilization of additional Ce3+ within fluorite lattice and appearance of strains can be due to insertion of lower sized Al3+ cations into the interstitial sites of c-CeO2 lattice resulting in charge redistribution, such as observed in c-CeO2/α-Al2O3 composites with close Ce/Al ratio [26]. The intensity of the CeO2 band in the spectra of Fe/Ce/Al samples does not decrease prominently (Table 3), while its position and especially the width become very close to those in the single crystals [23]. Such changes could be due to better ordering of CeO2 structure resulting from Al3+ “extraction” from the lattice after impregnation by acidic Fe-containing solutions that proceed with temperature increase and, to a less extent, some enlargement of crystallized CeO2 species (Table 1). It is the stabilization of small Fe-Al-O clusters in the points of Al insertion into CeO2 lattice after their extraction, but not γ-FeAlO3 phase decomposing at T > 400 °C, that can be responsible for the substantially higher content of Fe in the Fe-Al mixed oxide (Fe0.29Al) identified by DDPA compared with that in 2.5 Fe/(Ce + Al) sample (Table 2). Bands at 231 (not shown), 301, 419 and 622 cm−1 in the spectrum of 9.9 Fe/Ce/Al sample are usually related to lattice vibrations in α-Fe2O3 which agrees with XRD data (Table 1).
The absence of the band at ~377 cm−1 related to α-Al2O3 phases [27] in the Raman spectra giving information on near subsurface layers of all Ce/Al-based samples (Figure 3B) agrees with XRD and DDPA data. F2g mode of CeO2 at 461.5 cm−1 in the spectrum of support is supplemented by the bands at ~590, ~260 and ~830 cm−1 ascribed to the vacancy-interstitial (Frenkel-type) oxygen defects, surface mode, and peroxide (O22−) stretching vibration at the defective ceria surfaces, respectively [28,29,30]. These bands disappear in the spectrum of the 3.8 Fe/Ce/Al sample because the doping element can annihilate oxygen defects by dopant interstitial compensation mechanism [31]. The low-frequency shift of the CeO2 band and its broadening (Table 3) typical for smaller particles was detected in the spectra of Fe/Ce/Al samples which contradicts the slight increase in averaged sizes of crystalline CeO2 particles (Table 1). Therefore, the formation of highly dispersed particles of Fe-Ce mixed oxide in the surface layers of the 3.8 Fe/Ce/Al samples was detected by DDPA (Table 2) with inhomogeneous strains due to Ce3+ ions therein [15,28] can be responsible for such changes. Asymmetrical broadening with a low-energy shoulder of CeO2 band, especially prominent in the 9.9 Fe/Ce/Al sample, can result from a higher concentration of such compounds or higher content of α-Fe2O3 therein. In line with this, bands at ~260 and ~590 cm−1 appear again.
The band at 464 cm−1 which is responsible for F2g mode of fluorite lattice attenuates drastically in Fe/(Ce + Al) samples compared with that in the initial support (Figure 4A, Table 3). It is, more probably, due to the noticeable additional inclusion of Ce into alumina or Ce-Al mixed oxides with low Ce content where it exists preferentially as isolated Ce3+ ions in the presence of Fe. Indeed, Ce/Al = 0.05 in Ce-Al mixed oxide detected by DDPA in the (Ce + Al) sample but increased up to 0.09–0.26 after Fe supporting (Table 2). The absolute value of mixed oxides increased as well, while the quantity of insoluble CeO2 decreased from 2.4 mmol/g to 0.7 mmol/g (Table 2). At the same time, the position of CeO2 F2g mode in Fe/(Ce + Al) samples changed non-substantially pointing to the absence or low degree of ceria lattice modification by foreign atoms compared with the support. In this case, a slightly higher FWHM compared with that in Fe/Ce/Al samples is due to the particle size effect, which is true for crystallized particles, at least (Table 1) [15].
The band at ~377 cm−1 related to α-Al2O3 phases [27] was absent in the Raman spectra of near subsurface layers of (Ce + Al) based samples as well which agrees with XRD data (Figure S1, Table 1). After iron supporting, they undergo changes that are very similar to those in the bulk of the sample (Figure 4B, Table 3). An even more prominent drop in the intensity of the Ce band compared with that for the bulk can be related to the additional formation of Fe-Ce-Al mixed oxide with high Ce content (Fe0.05Ce0.26Al stoichiometry) (Table 2). Nevertheless, the close CeO2 band position and FWHM values in FT-Raman and Raman spectra point to a more uniform distribution of elements in the bulk and surface layers of these samples as compared to those in Ce/Al-based samples. This agrees with DDPA data on the preferential formation of Fe-Ce-Al-O mixed oxides in the 2.5 Fe/(Ce + Al) sample, while obviously spatially divided Fe-Al-O, Fe-Ce-O and more CeO2 were found in the 3.8 Fe/Ce/Al sample.

2.4. Surface Composition (XPS)

The normal complex formed due to shake-down satellites from an O1s to Ce 4f electron transfer in the Ce 3d spectra of all CeO2 containing samples (Figure 5A) was supplemented by the features marked as v’ and u’ due to the presence of Ce3+ [10]. An increase in Ce3+ fraction from 8% of total surface Ce concentration in CeO2 (spectrum not shown for shortness) up to 20% in 0.86 Fe/CeO2 sample was related to the formation of Fe-Ce–O solid solution or Fe-Ce mixed oxide [4,32]. Similar compounds of Fe0.19Ce stoichiometry have been detected in the 3.8 Fe/Ce/Al sample (Table 2, Figure 2). Therefore, there are two reasons that explain the further increase in Ce3+ fraction in 3.8 Fe/Ce/Al (25%) and especially Fe/(Ce + Al) (31-33%) samples (Table 4): stabilization of both isolated Ce cations in alumina-based structure of Fe-Ce-Al mixed oxides (Table 2) and Ce located on CeOx-AlOx boundary in the same oxides as Ce3+; lower CeO2 particle sizes (12 nm (2.5 Fe/(Ce + Al)) and 41 nm (3.8 Fe/Ce/Al), Table 1) compared with that in Fe/CeO2 (55 nm [4]) favoring reduction in Ce4+ [25].
Surface Ce concentration (Ces) and Ce/Al ratios in all samples (Table 4) are substantially lower than should follow from its integral composition (Ce/Al ≈ 0.3, Table S1), which agrees in general with higher dispersion of Al-containing compounds. The highest Ces and Ce/Al values were measured in the 3.8 Fe/Ce/Al sample (Table 2). However, the concentration of exposed Ce (Ceexp) in the Fe/Ce/Al and Fe/(Ce + Al) samples differ not so prominently because of higher SBET values in the last (Table 1). It is problematic to estimate Ces in the Fe-Ce-Al-O compounds of the samples. This value should be slightly higher in 2.5 Fe/(Ce + Al) sample (Ce/Al = 0.26 instead of Ce/Al = 0.2 for 3.8 Fe/Ce/Al sample, Table 2). Surface areas of crystallized CeO2 particles in Fe/(Ce + Al) and Fe/Ce/Al samples as formally estimated from their absolute content (Table S1) and sizes (Table 1) are quite close. However, encapsulation of about 29% of Ce composing Fe-Ce-O mixed oxides (Fe/Ce = 0.02, 0.8 mmol/g, Table 2) by Fe-Ce-Al mixed oxide was found for 2.5 Fe/(Ce + Al) sample (Figure 2A). This allows supposing bulk location of some CeO2 species as well thus decreasing Ces value therein. In addition, preferential exposing of highly dispersed Fe-Ce-O compound with Fe0.19Ce stoichiometry can additionally contribute to higher Ces value in the 3.8 Fe/Ce/Al sample. Hence, a substantial part of Ces in the Fe/(Ce + Al) samples composes Fe-Ce-Al mixed oxides with a high concentration of Ce+, while the fraction of CeO2-like structures is higher and seems to be responsible for greater Ces value in Fe/Ce/Al samples.
The 3.8 Fe/Ce/Al sample is characterized by the lowest Fe surface concentration (Fes) and the quantity of exposed Fe as related to a weight unit (Feexp) (Table 4). BE values of the Fe2p3/2 spectra ranging from 710.9 eV to 711.7 eV (Figure 5B) evidence presence of Fe3+ in oxyhydroxide environment (like in α- or γ-FeOOH), and in the oxide (like Fe2O3) structures [33,34,35], correspondingly dominating in 6.6 Fe/Al2O3 and 0.86 Fe/CeO2 samples [4]. Although the positions of the Fe 2P3/2 peak for Fe2+ in Fe0.94O, 2FeO∙SiO2 or Fe0.01Mg0.99O(100) single crystal were characterized by the values between 709.0 and 710.4 eV, its absence or low content in our samples follows from the absence of distinct satellite at 714.6–716 eV [36,37]. FeOOH usually undergoes a dehydroxylation reaction at 250–300 °C resulting in Fe2O3 [38]. At the same time, tetrahedral Fe3+ cations coordinating both terminal and bridging OH groups were found on the surface of both Fe-Al-O solid solution and highly dispersed FeOx species resulting from the decomposition of γ-FeAlO3–like structures [21]. In accordance with this, the Fe2p3/2 band at 711.7 eV (dominates in 3.8 Fe/Ce/Al sample) can be subscribed to: Fe3+ ions in Fe-Al mixed oxides, including those with Fe0.29Al stoichiometry (Table 2) formed by “extraction” of Al from CeO2 lattice after Fe supporting; highly dispersed FeOx species resulted from the decomposition of Fe-Al-O.
The band at 710.4 eV can characterize FeOx species on the surface of CeO2 or Fe3+ ions in the Fe-Ce-O compounds (Table 2, Figure 2). Its high abundance in Fe/(Ce + Al) samples (45–53% of total Fes) contradicts lower Ces values compared with that in 3.8 Fe/Ce/Al sample. However, the substantial part of the surface Fe in Fe/(Ce + Al) samples composes Fe-Ce-Al mixed oxides with reasonably high Ce content (Ce/Al = 0.26, Table 2). It resulted from the inclusion of additional Ce from CeO2 into the surface layers of alumina or Ce-Al mixed oxide of the initial support with Fe participation. Therefore, preferential formation of Fe-Ce-O oxide-like compounds is quite possible in Fe-Ce-Al mixed oxides. The contribution of Fe-Ce-Al mixed oxides into the surface composition in the 3.8 Fe/Ce/Al sample is not high which follows from the preferential dissolution of FeOx, binary Fe-Al-O and Fe-Ce-O compounds in HCl (Table 2) and substantially less prominent decrease in integral intensity of F2g vibrational mode of CeO2 in the near-surface layers of Fe/Ce/Al samples compared with corresponding support (Table 3).

2.5. O2 Adsorption/Desorption

O2 transients recorded during the He/0.5 vol.% O2 switch following 60 s stay in He flow revealed the delay of O2 response compared with Ar (Figure 6) corresponding to adsorption of reasonably close quantities of oxygen (3 × 1019–5 × 1019 oxygen atom/g) by reduced sites in the 2.5 Fe/(Ce + Al) and 3.8 Fe/Ce/Al samples.
Differently paired Fe2+ and Ce3+ ions resulting from the reduction in the oxygen absence are the most evident sites for O2 dissociative adsorption as coordinatively unsaturated surface metal cations. They can form in the surface layers of FeOx clusters, including those resulted from the decomposition of α-FeAlO3; Fe-Al-O mixed oxide of Fe0.29Al stoichiometry formed in 3.8 Fe/Ce/Al sample by Al “extraction” from CeO2 lattice during calcination; highly dispersed Fe-Ce mixed oxides (Fe0.19Ce stoichiometry (Table 2); Fe-Al-Ce-O mixed oxides with high Ce content (Fe0.05Ce0.26Al or Fe0.08Ce0.2Al stoichiometry).
Reduction of isolated Fe3+ in the Fe-Al-(Ce-)O mixed oxides with of Fe0.03Ce0.09Al stoichiometry (most of Ce stabilized as Ce3+ and unlikely can participate in any redox processes) requires transport of second oxygen atom through the occasional vacancies on the surface or in the bulk of alumina and thus be slow.

2.6. Oxygen Mobility (18O SSITKA)

The rates of oxygen exchange in Fe/Ce-Al-O samples in the period of 0–60 s after the 16O/18O switch can be ordered as 3.8 Fe/Ce/Al > 6.6 Fe/(Ce + Al)>2.5 Fe/(Ce + Al), as follows from the slope of NO(t)/g dependencies (Figure 7A). It does not correlate with SBET values of the samples (Table 1) and Fes or Feexp values (Table 4). Moreover, the difference between 3.8 Fe/Ce/Al and Fe/(Ce + Al) samples becomes even more prominent when related to one surface Fe site (Figure 7B). Fast exchange proceeding with reasonably close to initial rate value in these samples during about 20–25 s includes a higher quantity of oxygen atoms (NO = 1 × 1021–2 × 1021 O atoms/g) than can be related to Fe, even with an account of that located in the bulk (~1.0 × 1021 O atoms/g in the 6.6 Fe/(Ce + Al) sample with the highest Fe content). Therefore, efficient exchange in the near-surface layers of Ce-containing compounds contributes to NO as well. Indeed, substantially more profound and faster oxygen exchange took place in Fe/CeO2 samples compared with Fe/Al2O3 [4]. However, both Ce content and Ceexp values are reasonably close in all samples (Table 4). We consider that the higher rate of exchange in the 3.8 Fe/Ce/Al sample can be due to preferential exposure of Fe-Ce mixed oxide or crystallized CeO2, including that bonded with clusters of FeOx or Fe-Al mixed oxides, while in 2.5 Fe/(Ce + Al) sample substantial part of Fe and Ce are included as preferentially isolated ions into Fe-Ce-Al mixed oxides that still retain alumina-like structure. Although Fe-Ce mixed oxide was detected in the 2.5 Fe/(Ce + Al) sample (Table 2, Figure S6), it obviously locates in bulk (Figure 2), like part of CeO2, and thus does not contribute to the fast oxygen exchange.

2.7. Discussion

In accordance with the simplified redox scheme of N2O decomposition proposed by Kapteijn et al. [39]:
N2O + * → N2 + O*
2O* ↔ 2* + O2
one active (reduced) site is sufficient for N2 to evolve, but two neighboring O* are necessary for the desorption of O2. Factually, two neighboring active sites for N2O adsorption are desired for efficient reaction running, especially provided oxygen desorption from the surface is the rate-controlling step of this reaction [39]. Potential quantities of such reduced sites under reaction conditions in 2.5 Fe/(Ce + Al) and 3.8 Fe/Ce/Al samples are quite close (Figure 6). At the same time, the isotopic transient experiment 18O2/N216O performed on LaMnO3 at 900 °C revealed that there is direct oxygen transfer to the catalysts’ bulk from N2O molecule, and O2 formed involves lattice oxygen [40]:
2N216O + 218Olatt18O2 + 216Olatt + 2N2
This means that catalytic activity should depend on oxygen exchange properties.
In CeO2-containing compounds a simplified redox scheme realized in Fe oxides or FeOx/Al2O3 systems:
Fe2+ + N2O → N2 + Fe3+-O
2Fe3+-O ↔ 2Fe2++ O2
is supplemented by analogous steps for Ce3+/Ce4+ redox pair:
Ce3+ + N2O → N2 + Ce4-O
2Ce4+-O ↔ 2Ce3++ O2
And step corresponding to the additional pathway of oxygen species recombination provided by the presence of neighboring Fe3+ and Ce4+:
Fe3+O + Ce4+O ↔ Fe2+ + Ce3+ + O2
In addition, reoxidation of Ce3+ due to fast oxygen diffusion from CeO2 lattice (Olat) by extended oxygen vacancies arising after insertion of Fe3+ ions into the fluorite lattice:
Olat + Ce3+ ↔ Ce4+O
Can become an alternative pathway for the supply of the second oxygen atom which is necessary for O2 desorption and enhancing recombination of oxygen adspecies by steps 2B and 2C without adsorption of the second N2O molecule. Pure CeO2 revealed insubstantial activity as related per surface area unit compared with FeOx/CeO2 and even Fe/Al2O3 samples [4]. Therefore, the contribution of this pathway is high in Fe/Ce/Al samples containing substantial quantities of both Fe-Ce-O mixed oxide and bulky CeO2 bonded with FeOx or Fe-Al-O clusters. Stabilization of a substantial fraction of Ce as isolated Ce3+ cations composing mixed Fe-Ce-Al oxides in the Fe/(Ce + Al) samples diminishes substantially the efficiency of oxygen transfer to Fe3-O sites.

3. Materials and Methods

3.1. Catalysts Preparation

Two procedures were used to prepare the supports with a CeO2:Al2O3 weight ratio of 1:1. In the first of them, the precursor of CeO2 was deposited from 0.75 M Ce(NO3)3·6H2O (99.0%, Vekton, Saint-Petersburg, Russia) water solution using 1 M (NH4)2CO3 (99.3 %, Vekton, Saint-Petersburg, Russia) as a precipitation agent onto Al2O3 (the product of thermochemical activation of hydrargillite calcined at 500 °C) characterized by BET surface area at 210 m2g−1). After washing by H2O unless the pH of the filtrate was 7 and drying at 60 °C overnight, the sample was calcined at 900 °C for 5 h and denoted below as Ce/Al. In the second procedure, the water solution of (NH4)2CO3 was added dropwise to a water solution of Al(NO3)3·9H2O (98.5%, Ecros, Saint-Petersburg, Russia) and Ce(NO3)3·6H2O mixed in the necessary proportion unless (Ce + Al)/CO3 (mole) = 2.5 value was reached. Thus, the obtained gel was dried at 60 °C for 22 h followed by calcination at 500 °C for 2 h and at 900 °C for 5 h resulting in a sample denoted below as (Ce + Al).
FeOx was supported by incipient wetness impregnation of Ce/Al and (Ce + Al) by water solution of Fe(NO3)3·9H2O (98.5%, Vekton, Saint-Petersburg, Russia) with necessary concentration with added citric acid (99.8%, Vekton, Saint-Petersburg, Russia) in 10 wt.% excess to the stoichiometric amount and ethyleneglycole (99.0%, Ecros, Saint-Petersburg, Russia), dried in air at 150 °C for 3 h and then calcined at 900 °C for 4 h. In the commonly used abbreviation n Fe/support, “n” corresponded to Fe weight % concentration in the sample.

3.2. Characterization

XRD patterns were recorded using D8 diffractometer (Bruker, Germany) with CuKα monochromatic radiation. Each sample was scanned in the range of 2θ from 10° to 70° with a step 0.05°. The surface composition of the samples was investigated by X-ray photoelectron spectroscopy (XPS) using spectrometer SPECS (SPECS, Germany) with Al Kα irradiation (hν = 1486.6 eV). The positions of the peaks of Au 4f7/2 (84.0 eV) and Cu 2p3/2 (932.67 eV) core levels were used for calibration of the binding energy (BE) scale. In Raman spectroscopy depth of light penetration depends on the wavelength of monochromatic radiation provided by different sources. We believe that green light (514.5 nm line of an Ar+ laser with 2 mW power reaching the sample) provides information about the structure of preferentially near-surface layers of the samples, while data on the bulk composition are obtained with near-infrared (NIR) radiation at 1064 nm line (provided by an Nd-YAG laser with 100 mW power output) (Bruker Optik GmbH, Ettlingen, Germany). FT-Raman spectra (3700–100 cm−1, 300 scans, resolution 4 cm−1, 180° geometry) were collected using an RFS 100/S spectrometer (Bruker Optik GmbH, Ettlingen, Germany). Horiba Jobin Yvon T64000 spectrometer (HORIBA Scientific, Palaiseau, France) with micro-Raman setup and backscattering geometry for experimental spectra collection was used to measure the Raman spectra. The spectral resolution was not worse than 1.5 cm−1. The detector was a silicon-based CCD matrix, cooled with liquid nitrogen. The band at 520.5 cm−1 of Si single crystal was used to calibrate the spectrometer.
The method of differential dissolution phase analysis (DDPA) [41] was used to reveal the composition and, in some cases, morphology and particle depth distribution of the compounds and phases (including X-Ray amorphous ones) formed in the samples. For this about 10 mg of the sample loaded in a quartz microreactor was dissolved in the flow (3.6 mL/min) of water-based solution with the composition changing from HCl (pH = 2) to 3M HCl (with continuous temperature increase from 20 °C to 90 °C), and finally to 3.6 M HF. Compounds dissolving in milder conditions (HCl (pH = 2) and 1–3M HCl) were reasonably supposed to form in the surface layers of the samples. Better crystallized structures, as a rule, dissolve in HF or even remain insoluble. Change of the outlet mixture composition in time was analyzed by ICP AES 262477-364A spectrometer (BAIRD, Zoeterwoude, The Netherlands) using the spectral lines at 238.2, 308.2, and 413.8 nm which are characteristic for Fe, Al and Ce, respectively.

3.3. Kinetic Measurements and Catalytic Tests

For 16O/18O exchange experiments, the sample was first heated to 800 °C in 0.58 vol %16O2 + He flow and kept at this temperature for 30 min. After, this gas mixture was replaced stepwise by the same one containing 18O2 and Ar (1 vol.%) as an inert tracer. All responses were analyzed using QMS 200 gas analyzer (Stanford Research Systems, Sunnyvale, USA) as time variation of the 18O atomic fraction in the gas phase α g ( t ) = ( O 16 O 18 + 2 O 18 O 18 ) 2 ( O 16 O 18 + O 18 O 18 + O 16 O 16 ) . Time dependencies of exchanged oxygen for different samples as related to the mass unit (NO) were calculated using the formulae N o ( t ) = N A 2 C O 2 U g 0 t ( α g input α g ) dt , where αginput: isotope fraction in the inlet mixture (0.95), CO2: inlet O2 concentration (mol/mol), U: flow rate of the reaction mixture (mole/s), NA: Avogadro number. Dynamics of oxygen adsorption/desorption at 800 °C was elucidated from the experiments on the stepwise replacement of He by 0.5 vol.% O2 + 1 vol.% Ar + He mixture and vice versa flowing the reactor. (O2 + He)/He switch was performed after at least 30 min sample stay in O2 containing flow, while He/(O2 + He) one followed in about 60 s. All kinetic measurements were performed with the sample (g = 0.025 g, particles of 250–500 μm in a size) loaded into a reactor (quartz tube, i.d. = 3 mm). The gas flow rates of all mixtures amounted to 16.7 cm3/s.
The catalytic activity for samples with particles of 250–500 μm in a size was measured in a fixed-bed U-shaped reactor (3 mm i.d. quartz tube) at ambient pressure in the temperature range 700–900 °C. For NH3 oxidation, mixture 1% NH3 + 20% O2 in N2 was fed to the reactor charged by 0.015 g of the sample with a flow rate of 6.9 cm3/s. Concentrations of NH3 and NOx (x = 0.5–2) in the outlet mixture were measured by infrared spectroscopy. For N2O decomposition, a gas mixture of 0.15 vol% N2O in He flowed the reactor charged with 0.038 g of the sample with a flow rate of 16.7 cm3/s. In some experiments about 3 vol.% O2 (+3 vol.% H2O) were added to the inlet mixture. Outlet mixture composition was analyzed by gas chromatograph equipped with Porapack T (i.d. = 3mm, l = 3 m, for N2O analysis) and NaX (i.d. = 3mm, l = 2 m, for N2 analysis) columns. N2O or NH3 conversion (XN2O(NH3)) and yield (Yi) values calculated as X N 2 O = ( C N 2 O o C N 2 O )   C N 2 O o 100 % , X NH 3 = ( C NH 3 o C NH 3 )   C NH 3 o 100 % , and Y i = n C i C NH 3 o 100 % , ( C i and C i o : outlet and inlet concentrations of ith compound, n: number of N atoms in the ith molecule) were considered as a measure of samples activity.

4. Conclusions

FeOx (2.5–9.9 weight.% in terms of Fe) was supported by impregnation of mixed CeO2-Al2O3 support prepared either by co-precipitation (Ce + Al) or CeO2 precipitation onto Al2O3 (Ce/Al). Preferentially Fe-Ce-Al mixed oxides both in the surface layers and in the bulk were formed in Fe/(Ce + Al) samples, while the substantial spatial division of ceria and alumina-based compounds remains intact in Fe/Ce/Al samples. The stabilization of Ce3+ in Fe-Ce-Al mixed oxides was shown to inhibit oxygen mobility in the near-surface layers of Fe/(Ce + Al) samples. It results in retardation of the additional pathway of oxygen supply to the sites responsible for O2 desorption and explains the lower activity of Fe/(Ce + Al) samples in N2O decomposition.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12090938/s1, Figure S1: XRD patterns of Ce/Al (A) and (Ce + Al) (B) based samples with different content of Fe, Figure S2: Dependence of N2O conversion on the temperature for Ce/Al (A) and (Ce + Al) (B) based samples, Figure S3: Effect of O2 and H2O presence in the inlet flow on N2O conversion over 3.8 Fe/Ce/Al and 2.5 Fe/(Ce + Al) samples. T = 800 °C, Figure S4: Temperature dependence of the conversion (A) yield of (NO + NO2) and N2O (B) at NH3 oxidation over different Fe/Ce-Al-O based samples. Stability tests for 9.9 Fe/Ce/Al sample (C) at 800 °C, Figure S5: Differential dissolution curves of Ce and Al (A) and identified compounds (B) at stepwise consecutive change of flow composition from HCl (pH = 2) to 1-3M HCl (1) and then to 3.6 M HF (2) over (Ce + Al) sample, Figure S6: Differential dissolution curves of Fe, Ce and Al (A) and identified compounds (B) at stepwise consecutive change of flow composition from HCl (pH = 2) to 1-3M HCl (1) and then to 3.6 M HF (2) over 2.5 Fe/(Ce + Al) sample, Figure S7: Differential dissolution curves of Fe, Ce and Al (A) and initially (B) and finally (C) identified Fe-Ce-Al compounds at stepwise consecutive change of flow composition from HCl (pH = 2) to 1–3M HCl (1) and then to 3.6 M HF (2) over 3.8 Fe/Ce/Al sample; Table S1: Quantity of Fe, Ce, Al (as prepared, dissolved and insoluble) in different samples (mmol/g). Quantities of insoluble compounds were calculated from balance equations.

Author Contributions

Conceptualization, L.P.; investigation, L.P., I.P., Y.C.; writing—original draft preparation, L.P.; visualization, L.P., I.P., Y.C; writing—review and editing L.P., V.A., supervision, L.P., V.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation within the governmental order for Boreskov Institute of Catalysis (project AAAA-A21-121011390010-7), granted by the Government of the Russian Federation (075-15-2022-1132).

Acknowledgments

The authors appreciate the performing of DDPA measurements and discussion of the results by Dovlitova L.S.

Conflicts of Interest

The authors declare that they have no known competing interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Groves, M.C.E.; Sasonow, A. Uhde EnviNOx® technology for NOX and N2O abatement: A contribution to reducing emissions from nitric acid plants. J. Integr. Environ. Sci. 2010, 7, 211–222. [Google Scholar] [CrossRef]
  2. Pérez-Ramírez, J.; Kapteijn, F.; Schöffel, K.; Moulijn, J.A. Formation and control of N2O in nitric acid production. Where do we stand today? Appl. Catal. B Environ. 2003, 44, 117–151. [Google Scholar] [CrossRef]
  3. Sadykov, V.A.; Isupova, L.A.; Zolotarskii, I.A.; Bobrova, L.N.; Noskov, A.S.; Parmon, V.N.; Brushtein, E.A.; Telyatnikova, T.V.; Chernyshev, V.I.; Lunin, V.V. Oxide catalysts for ammonia oxidation in nitric acid production: Properties and perspectives. Appl. Catal. A Chem. 2000, 204, 59–87. [Google Scholar] [CrossRef]
  4. Pinaeva, L.G.; Prosvirin, I.P.; Dovlitova, L.S.; Danilova, I.G.; Sadovskaya, E.M.; Isupova, L.A. MeOx/Al2O3 and MeOx/CeO2 (Me = Fe, Co, Ni) catalysts for high temperature N2O decomposition and NH3 oxidation. Catal. Sci. Technol. 2016, 6, 2150–2161. [Google Scholar] [CrossRef]
  5. Pinaeva, L.G.; Dovlitova, L.S.; Isupova, L.A. Monolithic FeOx/Al2O3 Catalysts for Ammonia Oxidation and Nitrous Oxide Decomposition. Kinet. Catal. 2017, 58, 167–178. [Google Scholar] [CrossRef]
  6. Giecko, G.; Borowiecki, T.; Gac, W.; Kruk, J. Fe2O3/Al2O3 catalysts for the N2O decomposition in the nitric acid industry. Catal. Today 2008, 137, 403–409. [Google Scholar] [CrossRef]
  7. Kruk, J.; Stołecki, K.; Michalska, K.; Konkol, M.; Kowalik, P. The influence of modifiers on the activity of Fe2O3 catalyst for high temperature N2O decomposition (HT-deN2O). Catal. Today 2012, 191, 125–128. [Google Scholar] [CrossRef]
  8. Sádovská, G.; Tabor, E.; Bernauer, M.; Sazama, P.; Fíla, V.; Kmječ, T.; Kohout, J.; Závěta, K.; Tokarová, V.; Sobalík, Z. FeOx/Al2O3 catalysts for high-temperature decomposition of N2O under conditions of NH3 oxidation in nitric acid production. Catal. Sci. Technol. 2018, 8, 2841–2852. [Google Scholar] [CrossRef]
  9. Imagawa, H.; Suda, A.; Yamamura, K.; Sun, S. Monodisperse CeO2 Nanoparticles and Their Oxygen Storage and Release Properties. J. Phys. Chem. C 2011, 115, 1740–1745. [Google Scholar] [CrossRef]
  10. Chen, L.; Fleming, P.; Morris, V.; Holmes, J.D.; Morris, M.A. Size-Related Lattice Parameter Changes and Surface Defects in Ceria Nanocrystals. J. Phys. Chem. C 2010, 114, 12909–12919. [Google Scholar] [CrossRef]
  11. Iwanek, E.; Krawczyk, K.; Petryk, J.; Sobczak, J.W.; Kaszkur, Z. Direct nitrous oxide decomposition with CoOx-CeO2 catalysts. Appl. Catal. B Environ. 2011, 106, 416–422. [Google Scholar] [CrossRef]
  12. Grzybek, G.; Stelmachowski, P.; Gudyka, S.; Indyka, P.; Sojka, Z.; Guillén-Hurtado, N.; Rico-Pérez, V.; Bueno-López, A.; Kotarba, A. Strong dispersion effect of cobalt spinel active phase spread over ceria for catalytic N2O decomposition: The role of the interface periphery. Appl. Catal. B Environ. 2016, 180, 622–629. [Google Scholar] [CrossRef]
  13. Ferreira, A.P.; Zanchet, D.; Rinaldi, R.; Schuchardt, U.; Damyanova, S.; Bueno, J.M.C. Effect of the CeO2 content on the surface and structural properties of CeO2–Al2O3 mixed oxides prepared by sol–gel method. Appl. Catal. A Chem. 2010, 388, 45–56. [Google Scholar] [CrossRef]
  14. Damyanova, S.; Perez, C.A.; Schmal, M.; Bueno, J.M.C. Characterization of ceria-coated alumina carrier. Appl. Catal. A Chem. 2002, 234, 271–282. [Google Scholar] [CrossRef]
  15. Boullosa-Eiras, S.; Vanhaecke, E.; Zhao, T.; Chen, D.; Holmen, A. Raman spectroscopy and X-ray diffraction study of the phase transformation of ZrO2–Al2O3 and CeO2–Al2O3 nanocomposites. Catal. Today 2011, 166, 10–17. [Google Scholar] [CrossRef]
  16. Ge, C.; Liu, L.; Liu, Z.; Yao, X.; Cao, Y.; Tang, C.; Gao, F.; Dong, L. Improving the dispersion of CeO2 on γ-Al2O3 to enhance the catalytic performances of CuO/CeO2/γ-Al2O3 catalysts for NO removal by CO. Catal. Commun. 2014, 51, 95–99. [Google Scholar] [CrossRef]
  17. Laguna, O.H.; Romero Sarria, F.; Centeno, M.A.; Odriozola, J.A. Gold supported on metal-doped ceria catalysts (M = Zr, Zn and Fe) for the preferential oxidation of CO (PROX). J. Catal. 2010, 276, 360–370. [Google Scholar] [CrossRef]
  18. Zhang, Z.; Han, D.; Wei, S.; Zhang, Y. Determination of active site densities and mechanisms for soot combustion with O2 on Fe-doped CeO2 mixed oxides. J. Catal. 2010, 276, 16–23. [Google Scholar] [CrossRef]
  19. Li, K.; Wang, H.; Wei, Y.; Yan, D. Direct conversion of methane to synthesis gas using lattice oxygen of CeO2–Fe2O3 complex oxides. Chem. Eng. J. 2010, 156, 512–518. [Google Scholar] [CrossRef]
  20. Gupta, A.; Kumar, A.; Waghmare, U.V.; Hegde, M.S. Origin of activation of Lattice Oxygen and Synergistic Interaction in Bimetal-Ionic Ce0.89Fe0.1Pd0.01O2−δ Catalyst. Chem. Mater. 2009, 21, 4880–4891. [Google Scholar] [CrossRef]
  21. Escribano, V.S.; Amores, J.M.G.; Finocchio, E.; Daturi, M.; Busca, G. Characterization of α-(Fe,Al)2O3 solid-solution powders. J. Mater. Chem. 1995, 5, 1943–1951. [Google Scholar] [CrossRef]
  22. McBride, J.R.; Hass, K.C.; Poindexter, B.D.; Weber, W.H. Raman and x-ray studies of Ce1−xREXO2−y where RE=La, Pr, Nd, Eu, Gd, and Tb. J. Appl. Phys. 1994, 76, 2435–2441. [Google Scholar] [CrossRef]
  23. Kosacki, I.; Suzuki, T.; Anderson, H.U.; Colomban, P. Raman scattering and lattice defects in nanocrystalline CeO2 thin films. Solid State Ion. 2002, 149, 99–105. [Google Scholar] [CrossRef]
  24. Spanier, J.E.; Robinson, R.D.; Zhang, F.; Chan, S.W.; Herman, I.P. Size-dependent properties of CeO2−y nanoparticles as studied by Raman scattering. Phys. Rev. B 2001, 64, 245407. [Google Scholar] [CrossRef]
  25. Tsunekawa, S.; Ishikawa, K.; Li, Z.Q.; Kawazoe, Y.; Kasuya, Y. Origin of Anomalous Lattice Expansion in Oxide Nanoparticles. Phys. Rev. Lett. 2000, 85, 3440–3443. [Google Scholar] [CrossRef] [PubMed]
  26. Ojha, A.K.; Ponnilavan, V.; Kannan, S. Structural, morphological and mechanical investigations of in situ synthesized c-CeO2/α-Al2O3 composites. Ceram. Int. 2017, 43, 686–692. [Google Scholar] [CrossRef]
  27. Porto, S.P.S.; Krishnan, R.S. Raman Effect of Corundum. J. Chem. Phys. 1967, 47, 1009–1012. [Google Scholar] [CrossRef]
  28. Sudarsanama, P.; Malleshama, B.; Reddy, P.S.; Großmann, D.; Grünert, W.; Reddy, B.M. Nano-Au/CeO2 catalysts for CO oxidation: Influence of dopants (Fe, La and Zr) on the physicochemical properties and catalytic activity. Appl. Catal. B Environ. 2014, 144, 900–908. [Google Scholar] [CrossRef]
  29. Schilling, C.; Hofmann, A.; Hess, C.; Ganduglia-Pirovano, M.V. Raman Spectra of Polycrystalline CeO2: A Density Functional Theory Study. J. Phys. Chem. C 2017, 121, 20834–20849. [Google Scholar] [CrossRef]
  30. Reina, T.R.; Ivanova, S.; Centeno, M.A.; Odriozola, J.A. Boosting the activity of a Au/CeO2/Al2O3 catalyst for the WGS reaction. Catal. Today 2015, 253, 149–154. [Google Scholar] [CrossRef] [Green Version]
  31. Laguna, O.H.; Centeno, M.A.; Boutonnet, M.; Odriozola, J.A. Fe-doped ceria solids synthesized by the microemulsion method for CO oxidation reactions. Appl. Catal. B Environ. 2011, 106, 621–629. [Google Scholar] [CrossRef]
  32. Perez-Alonso, F.J.; Melián-Cabrera, I.; López Granados, M.; Kapteijn, F.; Fierro, J.L.G. Synergy of FexCe1−xO2 mixed oxides for N2O decomposition. J. Catal. 2006, 239, 340–346. [Google Scholar] [CrossRef]
  33. Suzuki, S.; Yanagihara, K.; Hirokawa, K. XPS study of oxides formed on the surface of high-purity iron exposed to air. Surf. Interface Anal. 2000, 30, 372–376. [Google Scholar] [CrossRef]
  34. Grosvenor, A.P.; Kobe, B.A.; Biesinger, M.C.; McIntyre, N.S. Investigation of multiplet splitting of Fe 2p XPS spectra and bonding in iron compounds. Surf. Interface Anal. 2004, 36, 1564–1574. [Google Scholar] [CrossRef]
  35. Atuchin, V.V.; Vinnik, D.A.; Gavrilova, T.A.; Gudkova, S.A.; Isaenko, L.I.; Jiang, X.; Pokrovsky, L.D.; Prosvirin, I.P.; Mashkovtseva, L.S.; Lin, Z. Flux Crystal Growth and the Electronic Structure of BaFe12O19 Hexaferrite. J. Phys. Chem. C 2016, 120, 5114–5123. [Google Scholar] [CrossRef]
  36. Yamashita, T.; Hayes, P. Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide materials. Appl. Surf. Sci. 2008, 254, 2441–2449. [Google Scholar] [CrossRef]
  37. Merte, L.R.; Gustafson, J.; Shipilin, M.; Zhang, C.; Lundgren, E. Redox behavior of iron at the surface of an Fe0.01Mg0.99O(100) single crystal studied by ambient-pressure photoelectron spectroscopy. Catal. Struct. React. 2017, 3, 95–103. [Google Scholar]
  38. Sayed, F.N.; Polshettiwar, V. Facile and sustainable synthesis of shaped iron oxide nanoparticles: Effect of iron precursor salts on the shapes of iron oxides. Sci. Rep. 2015, 5, 09733. [Google Scholar] [CrossRef]
  39. Kapteijn, F.; Rodriguez-Mirasol, J.; Moulijn, J. Heterogeneous catalytic decomposition of nitrous oxide. Appl. Catal. B Environ. 1996, 9, 25–64. [Google Scholar] [CrossRef]
  40. Ivanov, D.V.; Pinaeva, L.G.; Sadovskaya, E.M.; Isupova, L.A. Isotopic transient kinetic study of N2O decomposition on LaMnO3+δ. J. Mol. Catal. A Chem. 2016, 412, 34–38. [Google Scholar] [CrossRef]
  41. Malakhov, V.; Boldyreva, N.; Vlasov, A.; Dovlitova, L. Methodology and procedure of the stoichiographic analysis of solid inorganic substances and materials. J. Analyt. Chem. 2011, 66, 458–464. [Google Scholar] [CrossRef]
Figure 1. Dependence of N2O conversion (A) and (NO + NO2) and N2O yield at NH3 oxidation (B) on Fe content at 800 °C for Ce/Al and (Ce + Al) based samples.
Figure 1. Dependence of N2O conversion (A) and (NO + NO2) and N2O yield at NH3 oxidation (B) on Fe content at 800 °C for Ce/Al and (Ce + Al) based samples.
Catalysts 12 00938 g001
Figure 2. Structures formed in 2.5 Fe/(Ce + Al) (A) and 3.8 Fe/Ce/Al (B) samples.
Figure 2. Structures formed in 2.5 Fe/(Ce + Al) (A) and 3.8 Fe/Ce/Al (B) samples.
Catalysts 12 00938 g002
Figure 3. FT-Raman (A) and Raman (B) spectra of Ce/Al-based samples.
Figure 3. FT-Raman (A) and Raman (B) spectra of Ce/Al-based samples.
Catalysts 12 00938 g003
Figure 4. FT-Raman (A) and Raman (B) spectra of (Ce + Al) based samples.
Figure 4. FT-Raman (A) and Raman (B) spectra of (Ce + Al) based samples.
Catalysts 12 00938 g004
Figure 5. Ce 3d (A) Fe 2p (B) spectra of Fe-Ce-Al-O samples including contributions of Fe2O3 (710.4 eV) and FeOOH (711.7 eV)-like structures to overall Fe2p spectrum. 6.6 Fe/Al2O3 and 0.86 Fe/CeO2 samples were studied by Pinaeva et al. [4].
Figure 5. Ce 3d (A) Fe 2p (B) spectra of Fe-Ce-Al-O samples including contributions of Fe2O3 (710.4 eV) and FeOOH (711.7 eV)-like structures to overall Fe2p spectrum. 6.6 Fe/Al2O3 and 0.86 Fe/CeO2 samples were studied by Pinaeva et al. [4].
Catalysts 12 00938 g005
Figure 6. Responses of O2 and inert label (Ar) during the switch from He to He + 0.5 vol.% O2 flow over different samples. T = 800 °C.
Figure 6. Responses of O2 and inert label (Ar) during the switch from He to He + 0.5 vol.% O2 flow over different samples. T = 800 °C.
Catalysts 12 00938 g006
Figure 7. Time dependencies of the quantity of exchanged oxygen (No) as related to weight unit (A) or one surface Fe site (B) in 18O SSITKA experiments for different samples. T = 800 °C.
Figure 7. Time dependencies of the quantity of exchanged oxygen (No) as related to weight unit (A) or one surface Fe site (B) in 18O SSITKA experiments for different samples. T = 800 °C.
Catalysts 12 00938 g007
Table 1. Structural characteristics of Ce/Al and (Ce + Al) based samples.
Table 1. Structural characteristics of Ce/Al and (Ce + Al) based samples.
SampleSBET
(m2g−1)
Phase Composition
(Lattice Parameters)
DXRD
(nm)
Ce/Al63CeO2 (a = b = c = 5.411 Å)
(θ + γ)-Al2O3
40
-
3.8 Fe/Ce/Al50CeO2 (a = b = c = 5.411 Å)
(θ + γ)-Al2O3
41
-
9.9 Fe/Ce/Al35CeO2 (a = b = c = 5.411 Å)
(θ + γ)-Al2O3
α-Fe2O3 (a = b = 5.035 Å, c = 13.741 Å)
46
-
32
(Ce + Al)94CeO2 (a = b = c = 5.411 Å)
(θ + γ)-Al2O3
11
tr.
2.5 Fe/(Ce + Al)70CeO2 (a = b = c = 5.411 Å)
(θ + γ)-Al2O3
12
tr.
6.6 Fe/(Ce + Al)58CeO2 (a = b = c = 5.411 Å)
(θ + γ)-Al2O3
α-Fe2O3 (a = b = 5.035Å, c = 13.741 Å)
14
tr.
39
Table 2. Stoichiometry/quantity (mmol/g, without account of for O) of compounds dissolved in HCl and HF flows in different samples. Quantities of insoluble Ce compounds were calculated from balance equations.
Table 2. Stoichiometry/quantity (mmol/g, without account of for O) of compounds dissolved in HCl and HF flows in different samples. Quantities of insoluble Ce compounds were calculated from balance equations.
SampleCe + Al2.5 Fe/(Ce + Al)3.8 Fe/Ce/Al
HClAl/1.15Fe/0.06
Fe0.05Al/1.30
Fe0.05Ce0.26Al/0.65
Fe/0.07
Fe0.29Al/0.37
Fe0.19Ce/0.12
Ce0.05Al/0.67
HFCe0.05Al/7.92
Al/0.45
Ce/0.069
Fe0.05Ce0.26Al/2.62
Fe0.03Ce0.09Al/8.14
Fe0.02Ce/0.80
Fe/0.03
Fe0.08Ce0.2Al/2.93
Fe0.04Al/7.47
Insoluble Ce2.40.72.0
Table 3. Position, width (FWHM) and integral intensity of F2g vibrational mode of CeO2 in the FTRaman and Raman spectra.
Table 3. Position, width (FWHM) and integral intensity of F2g vibrational mode of CeO2 in the FTRaman and Raman spectra.
Sample Position, cm−1FWHM, cm−1Integral Intensity, arb.unit
FT-RamanRamanFT-RamanRamanFT-RamanRaman
Ce/Al462.9461.524.5 21.20.101 30399
3.8 Fe/Ce/Al465.1 455.810.6 21.80.087 27805
9.9 Fe/Ce/Al464.9433.010.8 66.30.079 26155
Ce + Al464.1 462.718.3 17.00.219 266717
2.5 Fe/(Ce + Al)464.0 455.817.3 23.40.0895 34159
6.6 Fe/(Ce + Al)464.0 459.616.3 20.00.0327 19715
Table 4. Surface (XPS) composition of different Fe-Ce-Al-O samples.
Table 4. Surface (XPS) composition of different Fe-Ce-Al-O samples.
SampleCe3+,
% of Total
Ces, a
at. %
Ceexp, rel.units bCe/AlFes a
at. % a
Feexp,
rel.units b
3.8 Fe/Ce/Al253.71840.121.680
2.5 Fe/(Ce + Al)31 2.41710.0751.7117
6.6 Fe/(Ce + Al)332.71590.0962.7166
a: calculated supposing C-containing compounds are absent under reaction conditions; accounting of C does not change the order of values. b: calculated as Fe(Ce)exp= Fe(Ce)s·SSA.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pinaeva, L.; Prosvirin, I.; Chesalov, Y.; Atuchin, V. High-Temperature Abatement of N2O over FeOx/CeO2-Al2O3 Catalysts: The Effects of Oxygen Mobility. Catalysts 2022, 12, 938. https://doi.org/10.3390/catal12090938

AMA Style

Pinaeva L, Prosvirin I, Chesalov Y, Atuchin V. High-Temperature Abatement of N2O over FeOx/CeO2-Al2O3 Catalysts: The Effects of Oxygen Mobility. Catalysts. 2022; 12(9):938. https://doi.org/10.3390/catal12090938

Chicago/Turabian Style

Pinaeva, Larisa, Igor Prosvirin, Yuriy Chesalov, and Victor Atuchin. 2022. "High-Temperature Abatement of N2O over FeOx/CeO2-Al2O3 Catalysts: The Effects of Oxygen Mobility" Catalysts 12, no. 9: 938. https://doi.org/10.3390/catal12090938

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