Next Article in Journal
Sustainable Biorefineries Based on Catalytic Biomass Conversion: A Review
Next Article in Special Issue
Laser-Induced Nitrogen-Doped Graphene Composite Iron–Cobalt Hydroxide for Methylene Blue Degradation via Electrocatalytic Activation of Peroxymonosulfate
Previous Article in Journal
Novel Laser-Assisted Chemical Bath Synthesis of Pure and Silver-Doped Zinc Oxide Nanoparticles with Improved Antimicrobial and Photocatalytic Properties
Previous Article in Special Issue
Ofloxacin Degradation over Nanosized Fe3O4 Catalyst viaThermal Activation of Persulfate Ions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Y Doping on Catalytic Activity of CeO2, MnOx, and CeMnOx Catalysts for Selective Catalytic Reduction of NO by NH3

by
Eleonora La Greca
1,
Tamara S. Kharlamova
2,
Maria V. Grabchenko
2,
Valery A. Svetlichnyi
3,
Giuseppe Pantaleo
1,
Luca Consentino
1,
Olga A. Stonkus
4,
Olga V. Vodyankina
2,* and
Leonarda Francesca Liotta
1,*
1
Institute for the Study of Nanostructured Materials (ISMN), National Research Council (CNR), Via Ugo La Malfa 153, 90146 Palermo, Italy
2
Laboratory of Catalytic Research, Tomsk State University, Lenin Ave. 36, 634050 Tomsk, Russia
3
Laboratory of Advanced Materials and Technology, Tomsk State University, Lenin Ave. 36, 634050 Tomsk, Russia
4
Boreskov Institute of Catalysis SB RAS, Lavrentiev Ave. 5, 630090 Novosibirsk, Russia
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(5), 901; https://doi.org/10.3390/catal13050901
Submission received: 5 April 2023 / Revised: 11 May 2023 / Accepted: 14 May 2023 / Published: 17 May 2023

Abstract

:
Novel yttrium-doped CeO2, MnOx, and CeMnOx composites are investigated as catalysts for low-temperature NH3-SCR. The study involves the preparation of unmodified oxide supports using a citrate method followed by modification with Y (2 wt.%) using two approaches, including the one-pot citrate method and incipient wetness impregnation of undoped oxides. The NH3-SCR reaction is studied in a fixed-bed quartz reactor to test the ability of the prepared catalysts in NO reduction. The gas reaction mixture consists of 800 ppm NO, 800 ppm NH3, 10 vol.% O2, and He as a balance gas at a WHSV of 25,000 mL g−1 h−1. The results indicate that undoped CeMnOx mixed oxide exhibits significantly higher deNOx performance compared with undoped and Y-doped MnOx and CeO2 catalysts. Indeed, yttrium presence in CeMnOx promotes the competitive NH3-SCO reaction, reducing the amount of NH3 available for NO reduction and lowering the catalyst activity. Furthermore, the physical-chemical properties of the prepared catalysts are studied using nitrogen adsorption/desorption, XRD, Raman spectroscopy, temperature-programmed reduction with hydrogen, and temperature-programmed desorption of ammonia. This study presents a promising approach to enhancing the performance of NH3-SCR catalysts at low temperatures that can have significant implications for reducing NO emissions.

Graphical Abstract

1. Introduction

Nitrogen oxides (NOx), which include NO2 and NO, are highly detrimental to human health and the environment. These emissions, commonly released by road vehicles, maritime transport, and industrial activities, contribute to a range of problems such as acid rain, photochemical smog, and ozone depletion [1,2,3]. Additionally, they can cause respiratory diseases in humans. To address these issues, governments have implemented stringent regulatory acts aimed at reducing atmospheric emissions [1,2,3]. In pursuit of this goal, numerous technologies, including electron beam processes [4], low-temperature adsorption [5], wet oxidative scrubbing (WOS) [6], NOx storage-reduction (NSR) [1], and selective catalytic reduction (SCR) [1,7], are under investigation. Innovative approaches such as electron beam processes, low-temperature adsorption, and WOS offer promising solutions to efficiently reduce and eliminate harmful nitrogen oxide pollutants from industrial emissions. However, these emerging technologies still face some limitations, such as energy consumption and wastewater treatment costs. To further enhance their applicability, ongoing investigation and research are required. Conversely, SCR technology has emerged as one of the most advanced, economical, and effective options to reduce NOx emissions from diesel engines, both from mobile and stationary sources. The system that operates in the presence of excess oxygen employs a reducing agent, typically ammonia (NH3-SCR) or urea solution (urea-SCR), which undergoes hydrolysis to produce ammonia in situ [7]. In the presence of catalysts, this reaction enables the NOx conversion to nitrogen (N2) and water (H2O) through the following overall reaction:
4NH3 + 4NO + O2 → 4N2 + 6H2O.
Among the catalysts that are currently under investigation, the V2O5-WO3/TiO2 composite is the most widely used in the SCR reaction. This is due to its high activity at medium temperatures (in the range of 300–400 °C), excellent chemical stability of the titania support, and good resistance to SO2 poisoning [8]. Despite its widespread use, this catalyst possesses some intrinsic drawbacks that require further searching for alternative systems. These limitations include a narrow operating temperature range [9,10], low selectivity towards N2 at higher temperatures, high toxicity of V2O5, which sublimes above 500 °C, leading to health concerns [11], and the transformation of TiO2 support from anatase to rutile at high temperatures [12]. Thus, there is a strong need to develop new, non-vanadium-based catalysts that demonstrate high performance, good hydrothermal stability, and are not susceptible to poisoning by exhaust components.
Oxide-based materials, particularly mixed oxides, have recently attracted significant interest as potential catalysts for the reduction of NOx emissions from diesel engine exhaust gases [9,13]. Manganese oxide (MnOx), in particular, has been extensively investigated due to its ability to convert NOx at low temperatures, variable valence states, and excellent oxygen migration ability that facilitates NO oxidation to NO2. However, the practical application of manganese-based catalysts is limited by their low N2 selectivity, poor resistance to H2O and sulfur, and narrow operating temperature window, which requires further improvements [14,15].
Cerium oxide (CeO2), a material that has been successfully employed in various reactions, has also shown promise in eliminating NOx emissions due to its good redox properties and high oxygen storage/release capacity resulting from oxygen vacancies in the material and the Ce4+/Ce3+ redox couple. In addition, CeO2 has a low cost and low environmental impact. The NOx adsorption by catalysts, the NO oxidation to NO2, the strength of Brønsted acid sites, and the water and sulfur resistances can all be enhanced by using CeO2, all of which are crucial and beneficial for the NH3-SCR reaction [16,17].
Several studies [18,19,20,21,22,23] have reported that the combination of cerium and manganese oxides can improve the catalytic activity and sulfur tolerance in the NH3-SCR reaction of NOx. This improvement is attributed to the promotion of redox reactions between Ce and Mn species as well as the improvement of oxygen mobility and storage capacity [24].
In addition to the combination of Ce and Mn oxides, the use of dopants to further enhance the catalytic activity and stability of various materials has been suggested. Dopants with ionic radius and electronegativity similar to those of supports have been identified as suitable candidates to modify the physical-chemical properties. Several studies have reported that the redox capacity, oxygen vacancy and mobility, surface area, hydrothermal stability, and SO2 tolerance of various oxides can be significantly improved due to doping with rare-earth metals [25,26,27]. In the case of ceria, doping with trivalent rare-earth ions such as Y3+ results in the incorporation and enrichment of these ions within the CeO2 lattice. Y3+ is highly stable, and its valence state is lower than that of Ce4+. Thus, the CeO2 doping with Y can increase the concentration of oxygen vacancies to compensate for the negative charge, leading to the formation of a defective structure that improves the mobility of the oxygen lattice [25].
Based on the above studies, the aim of this work is to explore the potential to use manganese oxide (MnOx), cerium oxide (CeO2), and mixed oxide (CeMnOx) that have been previously studied as supports for Ag catalysts in the selective catalytic reduction of NO with hydrocarbons (HC-SCR) [28] in the NO-SCR by NH3, as well as to estimate the impact of yttrium doping on the materials’ performance.

2. Results and Discussion

Table 1 lists the samples studied. The citrate sol-gel method was used to synthesize the undoped oxides (samples designated as CeO2, MnOx, and CeMnOx), while two synthesis methods were employed to prepare the Y-doped materials, i.e., an impregnation of the prepared undoped oxides with an aqueous solution of Y(NO3)3 (samples designated as Y/CeO2, Y/MnOx, Y/CeMnOx) and a one-pot sol-gel citrate method (samples designated as Y-CeO2, Y-MnOx, Y-CeMnOx). The catalysts were characterized by a complex of methods such as low-temperature nitrogen adsorption, X-ray diffraction (XRD), Raman spectroscopy, temperature-programmed reduction with H2 (H2-TPR), and temperature-programmed desorption of ammonia (NH3-TPD), to evaluate and compare physical properties and chemistries of the materials. The real composition of the samples was similar to the nominal one.

2.1. Textural Characteristics

Table 1 presents the values of specific surface area (SSA) and total pore volume of the samples determined from low-temperature nitrogen adsorption data. Manganese oxide-based samples are characterized by relatively low values of SSA (from 9 to 19 m2/g) and total pore volume (from 0.10 to 0.16 cm3/g), while the samples based on cerium oxide and mixed cerium and manganese oxides are characterized by rather high values of SSA (within 32–43 and 50–62 m2/g, respectively) and total pore volume (within 0.14–0.16 and 0.21–0.24 cm3/g, respectively). In general, a comparison of undoped and Y-doped samples indicates that the Y introduction using the one-pot citrate sol-gel method leads to an increase in both SSA and pore volume of the obtained materials, while the Y introduction via the impregnation of undoped oxides is accompanied by a significant decrease in the SSA with a slight change in the pore volume.

2.2. XRD Analysis

Figure 1 shows the XRD patterns for the samples, and Table 2 lists the phase composition of the samples and the characteristics of crystalline phases according to the XRD data.
For the unmodified MnOx sample, the formation of primarily tetragonal Mn3O4 (75 wt.%) with some cubic Mn2O3 (25 wt.%) was observed. The tetragonal Mn3O4 with lattice parameters a = 5.76 Å and c = 9.46 Å was rather dispersed (mean crystallite size was ~33 nm), while the Mn2O3 phase with the lattice parameter a = 9.41 Å was coarse (mean crystallite size was ~61 nm). In general, the amount of cubic Mn2O3 in the undoped sample depended on thermal treatment, with the increase in calcination temperature up to 650 °C resulting in the formation of only the Mn2O3 phase (Figure S1, in the Supplementary Materials). The MnOx modification via impregnation resulted in the formation of monoclinic Mn5O8 as the main crystalline phase (55 wt.%) with crystal lattice parameters a = 10.42 Å, b = 5.73 Å, c = 4.87 Å, and β = 109.9° and a crystallite size of 24 nm in the Y/MnOx sample. Along with the monoclinic phase, tetragonal Mn3O4 (34 wt.%) and some cubic Mn2O3 (11 wt.%) oxides were also formed. Finally, the MnOx modification during sol-gel synthesis resulted in the formation of a single-phase Mn3O4 oxide, which was stable even after calcination at 650 °C (Figure S1, in the Supplementary Materials). The lattice parameters of the tetragonal Mn3O4 phase in the Y-MnOx sample were practically unchanged as compared with the unmodified MnOx sample.
According to the XRD data, only a phase with a fluorite structure was revealed in the unmodified CeO2, CeMnOx, and Y-modified Y/CeO2, Y-CeO2, Y/CeMnOx, and Y-CeMnOx samples. In the case of the CeO2 sample, the fluorite phase was characterized by a lattice parameter a of 5.41 Å and an average crystallite size of 18 nm. In the case of the CeMnOx sample, the fluorite phase was characterized by a lower lattice parameter a = 5.40–5.38 Å and an average crystallite size of 5–7 nm as compared with the CeO2 sample. The observed decrease in lattice parameter can be caused by the substitution of Ce4+/Ce3+ ions with smaller Mn4+/Mn3+ ions in the fluorite lattice [29,30]. Modification of both CeO2 and CeMnOx samples with Y did not strongly affect the lattice parameter of the fluorite phase, which can be due to the low modifier content; however, the Y introduction during the sol-gel synthesis was accompanied by a decrease in the average crystallite size.

2.3. Raman Spectroscopy

The structural features of the samples were additionally studied by Raman spectroscopy. Figure 2 shows the Raman spectra for the samples.
Raman spectra for MnOx and Y-MnOx samples are similar and contain weak bands at 262, 309, and 360 cm−1, and an intense one at 644 cm−1, which are characteristic of Mn3O4 with the spinel structure [31,32,33,34]. The intense band at 644 cm–1 is assigned to the A1g mode, and the weak bands at 263, 308, and 360 cm–1 are referred to as the Eg, B2g, and Eg modes, respectively [32]. In the case of the Y/MnOx sample obtained by MnOx impregnation, the spectrum contains the bands at 170, 222, 262, 390, 427, 475, 495, 531, 577, 623sh, and 644 cm−1 characteristic of the Mn5O8 phase [35], which is consistent with the XRD data. The observed transition of Mn3O4 to Mn5O8 in the Y/MnOx sample and Mn3O4 stability in the Y-MnOx sample are associated with different Y distributions, with the Y introduction through the sol-gel method resulting in the stabilization of the Mn3O4 phase.
The spectrum for the CeO2 sample contains the bands at 263, 404sh, 464, 590, and 827 cm–1 typical for a fluorite-type ceria [36,37]. The intense band at 464 cm–1 is the F2g band associated with the Ce–O stretching in the [CeO8] cubic subcell, while the weak bands at 263, 404, and 590 cm–1 correspond to overtones, and the one at 827 cm–1 is due to the surface peroxide O22- structures. The Y introduction in ceria during the sol-gel synthesis leads to the appearance of well-defined bands at 547 and 600 cm−1 in the spectrum for the Y-CeO2 sample. These bands are associated with the Y3+ incorporation into CeO2 [38,39,40]. The band at 600 cm–1 (band D2) is associated with defect modes generated in ceria due to Ce4+ replacement by trivalent M3+ ions (M3+ = Ce3+, Gd3+, La3+, Y3+, etc.) and is assigned to defects with Oh symmetry [38,39,40]. The band at 547 cm–1 (band D1) is associated with defects, including the oxygen vacancy, with a symmetry different from that of the Oh point group [39,40]. In the case of Y introduction in ceria via impregnation, the spectrum for the Y/CeO2 sample does not contain the bands at 547 and 600 cm–1, which indicates the predominant formation of undoped ceria as well as Y distribution primarily in the surface and subsurface layers.
The Raman spectra for CeMnOx, Y/CeMnOx, and Y/CeMnOx samples contain two intense broad bands with maxima at 456–462 and 644 cm−1 due to the stretching of Ce-O in ceria and Mn-O in manganese oxides, respectively. On the shoulders of these peaks as well as in the region below 400 cm−1, weak bands due to overtones and defects in CeO2 and other modes in Mn3O4, Mn2O3, and Mn5O8 are distinguishable. A strong broadening of the bands indicates a high dispersion of ceria and manganese oxides present in the samples, which is additionally confirmed by the HR TEM data showing the nanodomain structure for such samples (Figure 3). Therefore, the Raman spectra indicate the formation of a mixture of highly dispersed ceria and manganese oxide in the CeMnOx, Y/CeMnOx, and Y/CeMnOx samples but not a solid solution based on the fluorite phase as indicated by the XRD data. However, the partial incorporation of Mn as well as Y into the ceria fluorite phase cannot be excluded. This is consistent with our recent study on Ag/CeMnOx catalysts, where a “patchwork” domain microstructure composed of nanocrystallites enriched by either Mn or Ce was revealed for the mixed oxide support [28].

2.4. H2-TPR

Reducibility is one of the most important parameters influencing the SCR activity of the catalysts. For this reason, the reducibility of samples was investigated by H2-TPR. Figure 4 represents the TPR profiles. As shown in the figure, two peaks at 491 and 868 °C appear in the H2-TPR curve for CeO2, which can be assigned to the surface/subsurface and bulk reduction of CeO2, respectively [41]. Doping with yttrium leads to a shift of both reduction peaks to the low-temperature region independently of the method of ceria doping. Doping ceria with Y3+ (whose ionic radius is 1.02 Å) leads to the production of oxygen vacancies; in this way, the oxygen mobility and, consequently, the redox properties of ceria are changed [42]. Additionally, the profiles for Y-CeO2 and Y/CeO2 samples exhibit a well-defined shoulder at 313 °C that can be attributed to the reduction of Y-promoted ceria surface. A comparison of the TPR profiles shows that there are no substantial differences resulting from the method of yttrium introduction; this suggests that the distribution of yttrium in the bulk or in the surface and subsurface layers of the fluorite phase does not significantly influence the reduction process.
The TPR profile for the MnOx sample contains two low-intense peaks at 246 and 365 °C and an intense one at 518 °C. These results are consistent with the previously reported data and with the XRD and Raman spectroscopy data indicating Mn2O3 and Mn3O4 to be the primary active phases for NO reduction in MnOx catalysts. As reported in the literature, manganese oxide reduction can occur through a three-step process: MnO2→Mn2O3→Mn3O4→MnO [43,44]. Specifically, the peaks observed at lower temperatures (<450 °C) correspond to the two-step reduction of MnO2 (MnO2 to Mn2O3 and Mn2O3 to Mn3O4, respectively); the peak at higher temperatures is attributed to the complete reduction of Mn3O4 to MnO [45]. Subsequent reduction of MnO to Mn0 is generally not observed even up to reduction temperatures above 700 °C since this reduction is characterized by a higher negative reduction potential [46]. Thereby, the reduction peaks at 246 and 365 °C in the TPR profile for the MnOx sample are attributed to the reduction of coarse Mn2O3 particles to Mn3O4 on the surface and in the bulk, respectively, while the peak at 518 °C is assigned to Mn3O4 reduction to MnO.
The Y doping via one-pot sol-gel synthesis results in hydrogen consumption only at temperatures above 400 °C, with practically no reduction peaks being observed at lower temperatures. A wide peak at 536 °C with a shoulder at 612 °C is attributed to the reduction of the Y-doped Mn3O4 phase in accordance with the XRD data. In the case of Y doping via MnOx impregnation, the TPR profile contains a new peak at 438 °C along with a high temperature peak at 566 °C with a shoulder at 612 °C. In accordance with the phase composition of the Y/MnOx sample, this peak can be attributed to Mn5O8 reduction to Mn3O4.
Compared with the reducibility of undoped and Y-doped CeO2 and MnOx, the reduction temperatures of the mixed CeMnOx oxides are shifted towards lower temperatures. The decrease in the reduction temperature indicates that the reducibility of MnOx and CeO2 has been promoted, resulting in better mobility of the oxygen species, which improves the redox performance of the catalyst [47,48]. Based on the XRD results, introducing moderate amounts of Mn can generate more lattice defects and oxygen-free places, develop oxygen migration, transformation, and release capacities, and improve the redox capacity of the catalyst. The Y doping results in some changes within the same temperature interval in the TPR profiles for Y-CeMnOx and Y/CeMnOx samples, suggesting that some differences in the phase composition of the nanodomains of the oxide composites may occur. The yttrium introduction via the sol-gel method in general results in improving the reducibility of Y-CeMnOx, while its introduction via the impregnation technique results in some hindered reducibility. This finding also seems to be caused by a different yttrium distribution, with impregnation techniques resulting in its distribution primarily on the surface.

2.5. Catalytic Performance

The obtained catalysts were studied in the NO SCR with NH3 in the temperature range of 25 to 400 °C. Figure 5 shows the catalytic activity of the oxide catalysts. The NO conversion increases as the temperature rises, while the N2 selectivity exhibits the opposite trend. SCR activity generally decreases in the following order: CeMnOx > Y-CeMnOx > Y/CeMnOx > MnOx > Y-MnOx > Y/MnOx > Y-CeO2 > Y/CeO2 > CeO2. With the Y addition, the catalytic efficiency increases nonlinearly in various catalysts. Thus, CeO2 shows a noticeable catalytic activity above 200 °C, but the NO conversion observed within 200–400 °C does not exceed 28%. The NO conversion at 200–400 °C increases for the Y-doped samples prepared by the one-pot sol-gel method.
The MnOx-based samples feature a higher NO conversion than those based on CeO2. The MnOx sample shows a noticeable catalytic activity already at 50 °C that goes through its maximum at 200 °C. However, the NO conversion over the MnOx catalyst only reaches 91.8% at 200 °C. The Y doping results in a decrease in the catalytic activity of the Y-MnOx sample below 200 °C, but the NO conversion is promoted to 94.6% at 200 °C. For the Y/MnOx catalyst, the Y doping results in a remarkable decrease in the catalytic activity below 225 °C, with the highest NO conversion of 78% being observed at 225 °C. This is consistent with hindering the manganese oxide reducibility in the Y-doped samples, with the phase composition of the Y-doped samples also affecting their activity. In the Y/MnOx sample, the monoclinic Mn5O8 phase consisting of Mn2+ and Mn4+ [35] seems to be less active as compared with the tetragonal Mn3O4 phase consisting of Mn2+ and Mn3+ [32] in the Y-MnOx sample, which indicates the importance of the Mn3+ presence in the catalyst.
The catalytic activity is greatly improved for the CeMnOx-based mixed oxide catalysts. In fact, a significant NO conversion (26%) is observed for the CeMnOx sample at 50 °C, and it reaches 100% at 100–200 °C. However, in both cases, the sample doping with Y does not enhance the catalytic activity of the mixed oxide catalysts. For the Y-CeMnOx sample, the effect of Y doping on the NO conversion is negligible. The NO conversion remains almost constant between 100 and 200 °C, which is accompanied by the narrowing of the activity window for the Y/CeMnOx sample. The latter can be caused by the formation of the less active monoclinic Mn5O8 phase in the Y/CeMnOx sample, as in the case of Y/MnOx.
The so far reported differences in NO conversion between undoped oxides, Y-doped samples prepared by impregnation, and those synthesized using the one-pot sol-gel method can be attributed to several factors. It is likely that the one-pot method favors a more homogeneous Y distribution in the sample than the impregnation one. The Y-doped samples prepared by the one-pot method are characterized by higher SSA and lower crystal sizes than those prepared by impregnation, while the impregnation method results in the sample sintering. Both methods should generate a great number of structural defects in ceria due to the Y3+ introduction in the surface/subsurface layer or bulk of the fluorite phase, resulting in an improvement in the sample reducibility. However, the one-pot method results in stabilization of the tetragonal Mn3O4 (Mn2+Mn3+2O3) phase, while the impregnation method results in the formation of the monoclinic Mn5O8 (Mn2+2Mn4+3O8) phase, accompanied by a hindering of the sample reducibility in the temperature range of ~100–400 °C.
An ideal NH3-SCR catalyst should exhibit not only high conversion but also high selectivity towards N2 to avoid the emission of harmful nitrogen oxides such as NO2 and N2O. Among all the investigated oxides, MnOx and CeMnOx show higher NOx conversion than CeO2, but for all samples, the N2 selectivity drops drastically with the temperature increase, as confirmed also by the N2 yield of the samples (Figure 5 and Figure 6). The Y doping improves the N2 selectivity, which, in turn, results in enhanced catalyst efficiency at temperatures above ~125 °C. Such an effect is more evident for the Y-doped MnOx oxides, which exhibit higher N2 selectivity than the undoped sample. However, the maximum N2 yield shifts to higher temperatures. Regarding the Y-doped CeMnOx samples, the one-pot-prepared Y-oxide gives almost the same N2 selectivity curve, while a slightly enhanced selectivity and N2 yield were registered between ~125 and 175 °C for the Y/CeMnOx sample (Figure 5 and Figure 6). As for CeO2 and Y-doped ceria oxides, the N2 yield increases to a small extent for the Y-CeO2 sample in the range of 200–400 °C.
To sum up, the Y addition to CeO2, MnOx, and CeMnOx influences the N2 selectivity and yield in a different manner depending on the preparation method used and the nature of the manganese oxide formed. The higher selectivity of the Y-doped MnOx and CeMnOx samples is consistent with the lower sample reducibility.
The observed decrease in N2 selectivity and N2 yield registered for the CeO2, MnOx, and CeMnOx oxides and the corresponding Y-doped samples can be attributed to the competitive oxidation of ammonia (NH3-SCO) catalyzed by manganese [49] and cerium oxides, which leads to the formation of NO, NO2, and N2O and consequently decreases the selectivity. To confirm this hypothesis, Figure 7 presents the temperature dependencies of NH3, NO, NO2, and N2O concentrations for the most efficient CeMnOx, Y-CeMnOx, and Y/CeMnOx samples. The NH3 concentration rapidly decreased by increasing the temperature and was totally consumed at ~200 °C, resulting in an increase in the N2O concentration between ~150 and 300 °C, according to Refs. [50,51]. Then, above 250 °C, the presence of NO and NO2 was also observed.

2.6. Role of Redox Properties and Surface Acidity

According to the literature [23,50,51,52,53], the low-temperature NH3-SCR of NO can proceed over the oxide catalysts through the Eley–Rideal or Langmuir–Hinshelwood mechanisms. Both mechanisms include NH3 adsorption on Lewis acid sites followed by its transformation to active species (NH2, NH, etc.) and/or interaction with the adsorbed NO or the one from the gas phase to form N2. The NO adsorption can be accompanied by its transformation into an active NO2 species. The oxidative transformation of both adsorbed NH3 and NO molecules was shown to occur via the interaction with the lattice oxygen [50,51] and with gaseous oxygen taking part in the catalyst reoxidation. However, an abundance of the labile lattice oxygen species on the catalyst surface led to the N2O overoxidation [50]. Thereby, both the redox properties and surface acidity of the oxide catalysts play a crucial role in the NH3-SCR of NO.
NH3-TPD was used to study the surface acidity of the MnOx- and CeMnOx-based samples as more efficient ones. Figure 8 shows the NH3-TPD profiles for the samples, while in Table 3 the maximal NH3 desorption temperatures and the amounts of desorbed NH3 (µmol g−1) are listed for each sample. All supports exhibit a large NH3 desorption peak with a maximum in the range of 200–300 °C. According to the literature, the low-temperature peaks can be assigned to the desorption of ammonia chemisorbed on the weak Lewis acidic sites [54,55]. The amount of NH3 desorbed from the CeMnOx support (243.0 µmol g−1) is higher than that for MnOx (52.0 µmol g−1), indicating that this sample has a higher acidic capacity and, thus, is able to activate NH3 in the range of temperature typical for ammonia SCR in accordance with the catalytic results. For the Y-doped samples, the total acidity and strength of the acid sites increase notably as compared with MnOx and CeMnOx samples, with the samples prepared by the one-pot method being more acidic than those obtained by impregnation. The enhanced acidity can be attributed to the presence of Y3+ cations on the surface, which act as weak Lewis acid sites and can affect the surface acidity of the samples [56].
Despite the high surface acidity as compared with the undoped samples, the observed decrease of the catalytic activity for Y-MnOx, Y/MnOx, and Y/CeMnOx samples was attributed to their hindered reducibility. This resulted in the hindering of the oxidation of adsorbed NH3 with labile lattice oxygen species to active species (NH2, NH, etc.) at low temperatures and prevented its overoxidation to more NO at high temperatures. Thereby, to improve the catalytic performance of the oxide catalyst towards the NH3-SCR of NO, optimization of their redox properties and surface acidity is required. In this respect, the CeMnOx composite materials are good candidates for developing an effective catalyst for SCR of NO, with their doping being a promising way for further catalyst enhancement.

3. Materials and Methods

3.1. Sample Preparation

Unmodified and Y-modified (2 wt.%) oxide samples were prepared. The unmodified oxide supports were prepared by a citrate method, while two approaches were used to modify oxide supports with Y, including one-pot citrate method (series 1) and incipient wetness impregnation of undoped oxides with water solution of Y(NO3)3 (series 2).

3.1.1. Sample Preparation Using Sol-Gel Citrate Method

The unmodified and Y-modified (series 1) CeO2, MnOx and CeO2–MnOx supports with a molar ratio of Ce/Mn = 1 were synthesized by the citrate method. Analytical-grade Ce(NO3)3·6H2O, Mn(NO3)2·6H2O, and Y(NO3)3·6H2O salts were used as Ce, Mn, and Y precursors, and citric acid (C6H8O7·H2O) was used as a complexing agent. All reagents were used directly without any further purification. The concentrated aqueous solutions of Ce(NO3)3 (0.84 mol/L), Mn(NO3)2 (2.6 mol/L), and Y(NO3)3 (1.84 mol/L) were prepared from metal nitrates for further use.
The individual or mixed solution of the corresponding metal nitrates was added to an aqueous solution of citric acid with vigorous stirring, followed by heating to 55–60 °C and constant stirring to form sol. The citric acid-to-metal (Ce, Mn, Y) molar ratio was 1.2 (pH of solution was ~1–2). The obtained sol was held at 60 °C with constant stirring for 2 h to form a gel, followed by the gel’s aging in a drying furnace overnight at 80 °C for undoped samples or for 10–18 h at 60 °C for the Y-doped samples.
Then the resulting gel was heated (10 deg/min) up to 120 °C and additionally dried at 120 °C for 5 h in a muffle furnace, followed by heating (5 deg/min) up to 500 °C and calcination of the dried gel in air at 500 °C for 3 h.
The samples synthesized by this approach were designated as CeO2, MnOx, CeMnOx, Y-CeO2, Y-MnOx, and Y-CeMnOx.

3.1.2. Sample Modification Using Impregnation Technique

The Y-modified oxide supports (series 2) were additionally prepared by incipient wetness impregnation of undoped oxide supports Ce-3, Mn-2, CeMn-2 with the Y(NO3)3 aqueous solutions. The required volume of the Y(NO3)3 solution was added to the powder of oxide support, followed by a thorough mixing of the paste formed. The impregnated sample was dried at room temperature for 5 h and then at 80 °C for 6 h in a drying furnace. Then it was heated (10 deg/min) up to 500 °C and calcined at 500 °C for 2 h in the muffle furnace.
The samples synthesized by this approach were designated as Y/CeO2, Y/MnOx, and Y/CeMnOx.

3.2. Characterization

The prepared samples were studied by several characterization methods, including nitrogen adsorption/desorption at −196 °C, X-ray analysis (XRD), Raman spectroscopy, hydrogen temperature-programmed reduction (TPR), and NH3-TPD.

3.2.1. Low-Temperature Nitrogen Adsorption/Desorption

The specific surface area, total pore volume, and average pore diameter were determined from the low-temperature nitrogen adsorption/desorption (at −196 °C) using the TriStar II 3020 specific analyzer (Micromeritics, Norcross, GA, USA). Prior to experiments, all samples were degassed at 200 °C in a vacuum (10–2 Torr) for 2 h using the laboratory degassing station VacPrep Degasser (Micromeritics, Norcross, GA, USA). The specific surface area was determined by the Brunauer-Emmett-Teller (BET) method, and the pore volume and pore size distributions were determined by the Barrett-Joyner-Halenda (BJH) method using the desorption branch of the adsorption-desorption isotherm.

3.2.2. XRD

The XRD patterns for the samples were recorded by the X-ray diffractometer XRD-7000 (Shimadzu, Kyoto, Japan) with monochromatic CuKα radiation (1.54 Å) in the angle range of 10–70° 2θ and a scanning rate of 0.02°/s. The data were obtained using Bragg–Brentano geometry. Crystalline Si (a = 5.4309 Å, λ = 1.540562 Å) was used as an external standard for diffractometer calibration. The phase composition was analyzed using the PDF-4 database (Release 2021 RDB). To refine the lattice parameters and determine the crystalline size, the POWDER CELL 2.4 full profile analysis program was used.

3.2.3. Raman Spectroscopy

Raman spectra were obtained on the InVia spectrometer (Renishaw, Wotton-under-Edge, UK) equipped with the DM 2500M microscope (Leica, Wetzlar, Germany) with a 50× objective. For excitation, lasers with wavelengths of 532 nm and a power of 100 mW were used; the spectral resolution was 2 cm−1. To prevent changes in the samples, only 5% of the full laser power and 50% beam defocus were applied.

3.2.4. Transmission Electron Microscopy

Transmission electron microscopy (TEM) data were obtained using the double aberration-corrected electron microscope Themis Z (Thermo Fisher Scientific, Eindhoven, The Netherlands) operated at 200 kV. The samples for the TEM study were dispersed ultrasonically and deposited on copper grids covered with a holey carbon film.

3.2.5. H2-TPR

Hydrogen temperature-programmed reduction (TPR) measurements were carried out with the Automated Catalyst Characterization System AutoChem 2950HP (Micromeritics, Norcross, GA, USA) equipped with a thermal conductivity detector (TCD). About 0.1 g of a sample was used for each measurement. The samples were pre-treated with a mixture of 5 vol% O2 in He at 50 mL/min, heated up (10 °C/min) to 400 °C, and held at this temperature for 30 min. After cooling down to room temperature, the gas mixture of 5 vol.% H2 in Ar was introduced at 30 mL/min into the sample tube and was also used as a reference gas. During the analysis, the temperature was increased up to 1000 °C at a rate of 10 °C/min. The effluent gas was analyzed with a TCD.

3.2.6. NH3-TPD

Temperature-programmed desorption of ammonia (NH3-TPD) was performed using the Micromeritics Autochem 2910 equipped with an ultraviolet gas analyzer (ABB, Limas 11, ABB S.p.a. Milano, Italy). Prior to the ammonia adsorption experiment, the sample was pretreated in a He flow at 200 °C for 30 min. After cooling down to room temperature, a stream of 5% NH3/He (30 mL/min) was flowed over the sample for 1 h. To remove the physically adsorbed ammonia, the sample was purged in flowing He (100 mL/min) at 100 °C for 1 h. Then, after cooling down to room temperature, ammonia desorption was monitored with the UV gas analyzer. After flowing He (30 mL/min) and heating up to 600 °C (rate of 10 °C/min), holding time at 600 °C was 30 min. A cold trap was used before the gas detection system to condense any water desorbed from the sample. The ammonia concentration in terms of ppm NH3 desorbed/gcat under He flow (30 mL/min) was plotted versus time and temperature.

3.3. Activity Tests in SCR of NO by NH3

The NH3-SCR tests were performed in a fixed-bed, continuous-flow U quartz reactor with an inner diameter of 12 mm. The feed gas, consisting of 800 ppm NO + 800 ppm NH3 + 10 vol% O2 in He, was flowed over the catalyst (120 mg) at a rate of 50 mL·min−1, equivalent to a weight hourly space velocity (WHSV) of 25,000 mL g−1 h−1. The conversion values were measured as a function of temperature from 25 °C to 500 °C with a heating rate of 5 °C/min, holding for 40 min at each temperature that was increased by steps of 50 °C. The inlet and outlet gas compositions were analyzed by the ABB detectors: infrared (Limas 11) for NO, N2O, NO2, NH3, paramagnetic (Magnos 206) for O2. Each test was repeated three times to estimate its reproducibility. The error obtained was less than 2% for the reported values.
The NO conversion, N2 selectivity, and N2 yield were calculated according to the following equations [57]:
NO   conversion   ( % ) :   NO in NO out NO in × 100 ;
N 2   selectivity   ( % ) :   1 NO 2 out   + 2 N 2 O out   NO in   + NH 3 in   NH 3 out   NO out   × 100   ;
N 2   yield   ( % ) :   NO   conversion   ×   N 2   selectivity 100 .

4. Conclusions

The CeO2, MnOx, CeMnOx, and yttrium-doped supports were successfully synthesized through a combination of the citrate sol-gel method for support synthesis and incipient wetness impregnation with Y(NO3)3 aqueous solution for the Y deposition, or via a one-pot sol-gel synthesis approach. Activity tests confirmed that the CeMnOx support exhibited superior low-temperature performance compared with MnOx or CeO2 oxides. Specifically, full NO conversion was achieved between 75 and 200 °C, with the selectivity to N2 being close to 90% at low temperatures; however, above 150 °C, the selectivity rapidly decreased due to the ammonia oxidation. The results suggested that the CeMnOx mixed oxide may have considerable potential for the development of NH3-SCR technologies for the reduction of NO emissions, although further improvements in N2 selectivity are needed. While Y doping did not enhance the NO conversion activity of the mixed oxide, it widened the selectivity range for MnOx oxides. The remarkable NO SCR activity of the CeMnOx sample in such a large temperature window can be attributed to its excellent reduction capability between ~200 and 350 °C, likely resulting from highly active oxygen species derived from dispersed nanocrystallites enriched by either Mn or Ce oxides in intimate contact, thus enhancing the low-temperature catalytic properties.
Both Y-doped CeO2 and MnOx samples, prepared by the one-pot method, achieved slightly higher NO conversion values than the corresponding undoped oxides, although the curve was shifted to higher temperatures in the case of Y-MnOx. Conversely, no improvement of the low-temperature performance for Y/CeO2 by incipient wetness impregnation or even a negative effect was registered for the Y/MnOx sample.
Further studies are required to gain a complete understanding of the reaction mechanism governing the enhanced NO SCR performance of the CeMnOx mixed oxide, optimize the catalyst composition in terms of the Ce:Mn molar ratio, and further improve the physical-chemical properties. Nevertheless, our study provides promising results for the future development of effective and eco-friendly catalysts for NO emission control.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13050901/s1, Figure S1: XRD patterns of samples calcined at 650 °C.

Author Contributions

Conceptualization, L.F.L. and O.V.V.; investigation, E.L.G., T.S.K., M.V.G., G.P., L.C., O.A.S. and V.A.S.; formal analysis, E.L.G., T.S.K., G.P., L.C., O.A.S. and V.A.S.; writing—original draft preparation, E.L.G., T.S.K. and L.F.L.; editing, L.F.L. and O.V.V.; project administration, O.V.V. and L.F.L. All the authors contributed to the data discussion and manuscript preparation. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Ministry for Science and Education of the Russian Federation (Project No. 075-15-2021-1388), by the Italian Ministry of Foreign Affairs and International Cooperation (“Progettazione di Catalizzatori Attivi a base di Ag-Pt depositati su Ce e Mn modificati con Y per il Post-Trattamento dei Gas di Scarico emessi dai Motori Diesel” Prot. MAE01538512021-10-26), and by the Project NAUSICA (PON “R&S 2014–2020”, grant n. ARS01_00334).

Data Availability Statement

Not applicable.

Acknowledgments

The TEM studies were carried out using the facilities of the shared research center “National center of investigation of catalysts” at the Boreskov Institute of Catalysis. The authors acknowledge M.A. Salaev (Tomsk State University) for language review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gómez-García, M.A.; Pitchon, V.; Kiennemann, A. Pollution by Nitrogen Oxides: An Approach to NOx Abatement by Using Sorbing Catalytic Materials. Environ. Int. 2005, 31, 445–467. [Google Scholar] [CrossRef]
  2. Reşitoʇlu, I.A.; Altinişik, K.; Keskin, A. The Pollutant Emissions from Diesel-Engine Vehicles and Exhaust Aftertreatment Systems. Clean Technol. Environ. Policy 2015, 17, 15–27. [Google Scholar] [CrossRef]
  3. Flagiello, D.; Esposito, M.; Di Natale, F.; Salo, K. A Novel Approach to Reduce the Environmental Footprint of Maritime Shipping. J. Mar. Sci. Appl. 2021, 20, 229–247. [Google Scholar] [CrossRef]
  4. Chmielewski, A.G.; Sun, Y.; Zimek, Z.; Bułka, S.; Licki, J. Mechanism of NOx removal by electron beam process in the presence of scavengers. Radiat. Phys. Chem. 2002, 65, 397–403. [Google Scholar] [CrossRef]
  5. Wang, S.; Xu, S.; Gao, S.; Xiao, P.; Jiang, M.; Zhao, H.; Huang, B.; Liu, L.; Niu, H.; Wang, J.; et al. Simultaneous removal of SO2 and NOx from flue gas by low-temperature adsorption over activated carbon. Sci. Rep. 2021, 11, 11003. [Google Scholar] [CrossRef] [PubMed]
  6. Flagiello, D.; Di Natale, F.; Erto, A.; Lancia, A. Oxidative Scrubber for NOX Emission Control Using NaClO2 Aqueous Solutions. Chem. Eng. Trans. 2021, 86, 397–402. [Google Scholar] [CrossRef]
  7. Han, L.; Cai, S.; Gao, M.; Hasegawa, J.Y.; Wang, P.; Zhang, J.; Shi, L.; Zhang, D. Selective Catalytic Reduction of NOx with NH3 by Using Novel Catalysts: State of the Art and Future Prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef]
  8. Alemany, L.J.; Berti, F.; Busca, G.; Ramis, G.; Robba, D.; Toledo, G.P.; Trombetta, M. Characterization and Composition of Commercial V2O5-WO3-TiO2 SCR Catalysts. Appl. Catal. B Environ. 1996, 10, 299–311. [Google Scholar] [CrossRef]
  9. Wu, H.; He, M.; Liu, W.; Jiang, L.; Cao, J.; Yang, C.; Yang, J.; Peng, J.; Liu, Y.; Liu, Q. Application of Manganese-Containing Soil as Novel Catalyst for Low-Temperature NH3-SCR of NO. J. Environ. Chem. Eng. 2021, 9, 105426. [Google Scholar] [CrossRef]
  10. Li, Y.; Zhang, Z.; Zhao, X.; Liu, Z.; Zhang, T.; Niu, X.; Zhu, Y. Effects of Nb-Modified CeVO4 to Form Surface Ce-O-Nb Bonds on Improving Low-Temperature NH3-SCR DeNOx Activity and Resistance to SO2 & H2O. Fuel 2023, 331, 125799. [Google Scholar] [CrossRef]
  11. Chen, L.; Liu, J.; Hu, W.; Gao, J.; Yang, J. Vanadium in Soil-Plant System: Source, Fate, Toxicity, and Bioremediation. J. Hazard. Mater. 2021, 405, 124200. [Google Scholar] [CrossRef] [PubMed]
  12. Ding, L.; Wang, Y.; Qian, L.; Qi, P.; Xie, M.; Long, H. Flue Gas DeNOxing Spent V2O5-WO3/TiO2 Catalyst: A Review of Deactivation Mechanisms and Current Disposal Status. Fuel 2023, 338, 127268. [Google Scholar] [CrossRef]
  13. Casapu, M.; Kröcher, O.; Elsener, M. Screening of Doped MnOx-CeO2 Catalysts for Low-Temperature NO-SCR. Appl. Catal. B Environ. 2009, 88, 413–419. [Google Scholar] [CrossRef]
  14. Shi, J.W.; Gao, C.; Liu, C.; Fan, Z.; Gao, G.; Niu, C. Porous MnOx for Low-Temperature NH3-SCR of NOx: The Intrinsic Relationship between Surface Physicochemical Property and Catalytic Activity. J. Nanoparticle Res. 2017, 19, 194. [Google Scholar] [CrossRef]
  15. Kapteijn, F.; Singoredjo, L.; Andreini, A.; Moulijn, J.A. Activity and Selectivity of Pure Manganese Oxides in the Selective Catalytic Reduction of Nitric Oxide with Ammonia. Appl. Catal. B Environ. 1994, 3, 173–189. [Google Scholar] [CrossRef]
  16. Kašpar, J.; Fornasiero, P.; Graziani, M. Use of CeO2-Based Oxides in the Three-Way Catalysis. Catal. Today 1999, 50, 285–298. [Google Scholar] [CrossRef]
  17. Sun, H.; Park, S.J. Recent Advances in MnOx/CeO2-Based Ternary Composites for Selective Catalytic Reduction of NOx by NH3: A Review. Catalysts 2021, 11, 1519. [Google Scholar] [CrossRef]
  18. Wu, X.; Yu, H.; Weng, D.; Liu, S.; Fan, J. Synergistic Effect between MnO and CeO2 in the Physical Mixture: Electronic Interaction and NO Oxidation Activity. J. Rare Earths 2013, 31, 1141–1147. [Google Scholar] [CrossRef]
  19. Wang, C.; Yu, F.; Zhu, M.; Tang, C.; Zhang, K.; Zhao, D.; Dong, L.; Dai, B. Highly Selective Catalytic Reduction of NOx by MnOx–CeO2–Al2O3 Catalysts Prepared by Self-Propagating High-Temperature Synthesis. J. Environ. Sci. 2019, 75, 124–135. [Google Scholar] [CrossRef]
  20. Qi, G.; Yang, R.T.; Chang, R. MnOx-CeO2 Mixed Oxides Prepared by Co-Precipitation for Selective Catalytic Reduction of NO with NH3 at Low Temperatures. Appl. Catal. B Environ. 2004, 51, 93–106. [Google Scholar] [CrossRef]
  21. Xu, L.; Li, X.S.; Crocker, M.; Zhang, Z.S.; Zhu, A.M.; Shi, C. A Study of the Mechanism of Low-Temperature SCR of NO with NH3 on MnOx/CeO2. J. Mol. Catal. A Chem. 2013, 378, 82–90. [Google Scholar] [CrossRef]
  22. Chen, L.; Ren, S.; Xing, X.; Yang, J.; Li, J.; Yang, J.; Liu, Q. Effect of MnO2 crystal types on CeO2@MnO2 oxides catalysts for low-temperature NH3-SCR. J. Environ. Chem. Eng. 2022, 10, 108239. [Google Scholar] [CrossRef]
  23. Chen, Z.; Ren, S.; Zhou, Y.; Li, X.; Wang, M.; Chen, L. Comparison of Mn doped CeO2 with different exposed facets for NH3-SCR at low temperature. J. Energy Inst. 2022, 105, 114–120. [Google Scholar] [CrossRef]
  24. Wu, Z.; Jin, R.; Liu, Y.; Wang, H. Ceria Modified MnOx/TiO2 as a Superior Catalyst for NO Reduction with NH3 at Low-Temperature. Catal. Commun. 2008, 9, 2217–2220. [Google Scholar] [CrossRef]
  25. Zhou, B.; Xi, K.; Fan, L.J.; Zhou, Y.; Wang, Y.; Zhu, Q.L.; Lu, H.F. A Comparative Study on Ce–Pr and Ce–Mn Mixed Oxide Catalysts toward Soot Catalytic Combustion. Appl. Catal. A Gen. 2018, 562, 1–10. [Google Scholar] [CrossRef]
  26. Jiang, Z.; Chen, C.; Ma, M.; Guo, Z.; Yu, Y.; He, C. Rare-Earth Element Doping-Promoted Toluene Low-Temperature Combustion over Mesostructured CuMCeO: X (M = Y, Eu, Ho, and Sm) Catalysts: The Indispensable Role of in Situ Generated Oxygen Vacancies. Catal. Sci. Technol. 2018, 8, 5933–5942. [Google Scholar] [CrossRef]
  27. Zhang, S.; Liu, X.; Zhong, Q.; Yao, Y. Effect of y Doping on Oxygen Vacancies of TiO2 Supported MnOX for Selective Catalytic Reduction of NO with NH3 at Low Temperature. Catal. Commun. 2012, 25, 7–11. [Google Scholar] [CrossRef]
  28. La Greca, E.; Kharlamova, T.S.; Grabchenko, M.V.; Consentino, L.; Savenko, D.Y.; Pantaleo, G.; Kibis, L.S.; Stonkus, O.A.; Vodyankina, O.V.; Liotta, L.F. Ag Catalysts Supported on CeO2, MnO2 and CeMnOx Mixed Oxides for Selective Catalytic Reduction of NO by C3H6. Nanomaterials 2023, 13, 873. [Google Scholar] [CrossRef] [PubMed]
  29. Huang, H.; Liu, J.; Sun, P.; Ye, S.; Liu, B. Effects of Mn-Doped Ceria Oxygen-Storage Material on Oxidation Activity of Diesel Soot. RSC Adv. 2017, 7, 7406–7412. [Google Scholar] [CrossRef]
  30. Murugan, B.; Ramaswamy, A.V.; Srinivas, D.; Gopinath, C.S.; Ramaswamy, V. Nature of Manganese Species in Ce1-xMnxO2-δ Solid Solutions Synthesized by the Solution Combustion Route. Chem. Mater. 2005, 17, 3983–3993. [Google Scholar] [CrossRef]
  31. Yang, X.; Wang, X.; Zhang, G.; Zheng, J.; Wang, T.; Liu, X.; Shu, C.; Jiang, L.; Wang, C. Enhanced Electrocatalytic Performance for Methanol Oxidation of Pt Nanoparticles on Mn3O4-Modified Multi-Walled Carbon Nanotubes. Int. J. Hydrogen Energy 2012, 37, 11167–11175. [Google Scholar] [CrossRef]
  32. Larbi, T.; Doll, K.; Manoubi, T. Density Functional Theory Study of Ferromagnetically and Ferrimagnetically Ordered Spinel Oxide Mn3O4. A Quantum Mechanical Simulation of Their IR and Raman Spectra. J. Alloys Compd. 2016, 688, 692–698. [Google Scholar] [CrossRef]
  33. Gao, T.; Fjellvåg, H.; Norby, P. A Comparison Study on Raman Scattering Properties of α- and β-MnO2. Anal. Chim. Acta 2009, 648, 235–239. [Google Scholar] [CrossRef] [PubMed]
  34. Julien, C.M.; Massot, M.; Poinsignon, C. Lattice Vibrations of Manganese Oxides: Part I. Periodic Structures. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2004, 60, 689–700. [Google Scholar] [CrossRef] [PubMed]
  35. Gao, T.; Norby, P.; Krumeich, F.; Okamoto, H.; Nesper, R.; Fjellvåg, H. Synthesis and Properties of Layered-Structured Mn5O8 Nanorods. J. Phys. Chem. C 2010, 114, 922–928. [Google Scholar] [CrossRef]
  36. 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]
  37. Schilling, C.; Ganduglia-Pirovano, M.V.; Hess, C. Experimental and Theoretical Study on the Nature of Adsorbed Oxygen Species on Shaped Ceria Nanoparticles. J. Phys. Chem. Lett. 2018, 9, 6593–6598. [Google Scholar] [CrossRef]
  38. Taniguchi, T.; Watanabe, T.; Sugiyama, N.; Subramani, A.K.; Wagata, H.; Matsushita, N.; Yoshimura, M. Identifying Defects in Ceria-Based Nanocrystals by UV Resonance Raman Spectroscopy. J. Phys. Chem. C 2009, 113, 19789–19793. [Google Scholar] [CrossRef]
  39. Nakajima, A.; Yoshihara, A.; Ishigame, M. Defect-Induced Raman Spectra in Doped CeO2. Phys. Rev. B 1994, 50, 13297–13307. [Google Scholar] [CrossRef]
  40. Derevyannikova, E.A.; Kardash, T.Y.; Kibis, L.S.; Slavinskaya, E.M.; Svetlichnyi, V.A.; Stonkus, O.A.; Ivanova, A.S.; Boronin, A.I. The Structure and Catalytic Properties of Rh-Doped CeO2 Catalysts. Phys. Chem. Chem. Phys. 2017, 19, 31883–31897. [Google Scholar] [CrossRef]
  41. Grabchenko, M.V.; Mamontov, G.V.; Zaikovskii, V.I.; La Parola, V.; Liotta, L.F.; Vodyankina, O.V. Design of Ag-CeO2/SiO2 Catalyst for Oxidative Dehydrogenation of Ethanol: Control of Ag–CeO2 Interfacial Interaction. Catal. Today 2019, 333, 2–9. [Google Scholar] [CrossRef]
  42. Ilieva, L.; Venezia, A.M.; Petrova, P.; Pantaleo, G.; Liotta, L.F.; Zanella, R.; Kaszkur, Z.; Tabakova, T. Effect of Y Modified Ceria Support in Mono and Bimetallic Pd-Au Catalysts for Complete Benzene Oxidation. Catalysts 2018, 8, 283. [Google Scholar] [CrossRef]
  43. Trawczyński, J.; Bielak, B.; Miśta, W. Oxidation of Ethanol over Supported Manganese Catalysts—Effect of the Carrier. Appl. Catal. B Environ. 2005, 55, 277–285. [Google Scholar] [CrossRef]
  44. Stobbe, E.R.; De Boer, B.A.; Geus, J.W. The Reduction and Oxidation Behaviour of Manganese Oxides. Catal. Today 1999, 47, 161–167. [Google Scholar] [CrossRef]
  45. Zhan, S.; Zhu, D.; Qiu, M.; Li, Y. Highly efficient Removal of NO with Ordered Mesoporous Manganese Oxide at Low Temperature. RSC Adv. 2015, 5, 29353–29361. [Google Scholar] [CrossRef]
  46. Ye, Q.; Zhao, J.; Huo, F.; Wang, J.; Cheng, S.; Kang, T.; Dai, H. Nanosized Ag/α-MnO2 Catalysts Highly Active for the Low-Temperature Oxidation of Carbon Monoxide and Benzene. Catal. Today 2011, 175, 603–609. [Google Scholar] [CrossRef]
  47. Tang, X.; Chen, J.; Li, Y.; Li, Y.; Xu, Y.; Shen, W. Complete Oxidation of Formaldehyde over Ag/MnOx–CeO2 Catalysts. Chem. Eng. J. 2006, 118, 119–125. [Google Scholar] [CrossRef]
  48. Wang, K.; Liu, X.; Tu, S.; Zhang, L.; Li, W.; Jiang, C.; Ye, D. Low Temperature Catalytic Performance of Manganese and Cerium Complex Oxide Catalyst towards Toluene. IOP Conf. Ser. Mater. Sci. Eng. 2020, 729, 012069. [Google Scholar] [CrossRef]
  49. Consentino, L.; Pantaleo, G.; Parola, V.L.; Migliore, C.; Greca, E.L.; Liotta, L.F. NH3-NO SCR Catalysts for Engine Exhaust Gases Abatement: Replacement of Toxic V2O5 with MnOx to Improve the Environmental Sustainability. Top. Catal. 2022, 1–10. [Google Scholar] [CrossRef]
  50. Gao, F.; Liu, Y.; Sani, Z.; Tang, X.; Yi, H.; Zhao, S.; Yu, Q.; Zhou, Y. Advances in Selective Catalytic Oxidation of Ammonia (NH3-SCO) to Dinitrogen in Excess Oxygen: A Review on Typical Catalysts, Catalytic Performances and Reaction Mechanisms. J. Environ. Chem. Eng. 2021, 9, 104575. [Google Scholar] [CrossRef]
  51. Yang, S.; Liao, Y.; Xiong, S.; Qi, F.; Dang, H.; Xiao, X.; Li, J. N2 Selectivity of NO Reduction by NH3 over MnOx-CeO2: Mechanism and Key Factors. J. Phys. Chem. C 2014, 118, 21500–21508. [Google Scholar] [CrossRef]
  52. Liao, Y.; Liu, Z.; Li, Z.; Gao, G.; Ji, L.; Xu, H.; Huang, W.; Qu, Z.; Yan, N. The Unique CO Activation Effects for Boosting NH3 Selective Catalytic Oxidation over CuOx–CeO2. Environ. Sci. Technol. 2022, 56, 10402–10411. [Google Scholar] [CrossRef]
  53. Jiang, Y.; Han, D.; Yang, L.; Yang, Z.; Ge, H.; Lin, R.; Wang, X. Improving the K resistance effectively of CeO2-TiO2 catalyst by Nb doping for NH3-SCR reaction. Process Saf. Environ. Prot. 2022, 160, 876–886. [Google Scholar] [CrossRef]
  54. Lee, S.M.; Park, K.H.; Hong, S.C. MnOx/CeO2-TiO2 Mixed Oxide Catalysts for the Selective Catalytic Reduction of NO with NH3 at Low Temperature. Chem. Eng. J. 2012, 195–196, 323–331. [Google Scholar] [CrossRef]
  55. Kharlamova, T.S.; Timofeev, K.L.; Salaev, M.A.; Svetlichnyi, V.A.; Vodyankina, O.V. Monolayer MgVOx/Al2O3 catalysts for propane oxidative dehydrogenation: Insights into a role of structural, redox, and acid-base properties in catalytic performance. Appl. Catal. A 2020, 598, 117574. [Google Scholar] [CrossRef]
  56. Zhang, Y.; Zhou, Y.; Shi, J.; Zhou, S.; Zhang, Z.; Zhang, S.; Guo, M. Propane Dehydrogenation over PtSnNa/La-Doped Al2O3 Catalyst: Effect of La Content. Fuel Process. Technol. 2013, 111, 94–104. [Google Scholar] [CrossRef]
  57. Inomata, Y.; Mino, M.; Hata, S.; Kiyonaga, E.; Morita, K.; Hikino, K.; Yoshida, K.; Haruta, M.; Murayama, T. Low-temperature NH3-SCR Activity of Nanoparticulate Gold Supported on a Metal Oxide. J. Japan Pet. Inst. 2019, 62, 234–243. [Google Scholar] [CrossRef]
Figure 1. XRD patterns for samples.
Figure 1. XRD patterns for samples.
Catalysts 13 00901 g001
Figure 2. Raman spectra for samples.
Figure 2. Raman spectra for samples.
Catalysts 13 00901 g002
Figure 3. Typical HR TEM image for nanodomain structure of the CeMnOx sample.
Figure 3. Typical HR TEM image for nanodomain structure of the CeMnOx sample.
Catalysts 13 00901 g003
Figure 4. Temperature-programmed reduction profiles for samples.
Figure 4. Temperature-programmed reduction profiles for samples.
Catalysts 13 00901 g004
Figure 5. NO conversion (a,c,e) and N2 selectivity (b,d,f) over undoped and Y-doped samples.
Figure 5. NO conversion (a,c,e) and N2 selectivity (b,d,f) over undoped and Y-doped samples.
Catalysts 13 00901 g005
Figure 6. N2 yields over undoped and Y-doped samples based on MnOx (a), CeO2 (b) and CeMnOx (c).
Figure 6. N2 yields over undoped and Y-doped samples based on MnOx (a), CeO2 (b) and CeMnOx (c).
Catalysts 13 00901 g006
Figure 7. Comparison of NH3, NO, NO2, and N2O concentrations for CeMnOx (a), Y-CeMnOx (b), and Y/CeMnOx (c) samples.
Figure 7. Comparison of NH3, NO, NO2, and N2O concentrations for CeMnOx (a), Y-CeMnOx (b), and Y/CeMnOx (c) samples.
Catalysts 13 00901 g007
Figure 8. The NH3-TPD profiles for samples.
Figure 8. The NH3-TPD profiles for samples.
Catalysts 13 00901 g008
Table 1. Chemical composition, specific surface area (SSA), and total pore volume (V) of materials studied.
Table 1. Chemical composition, specific surface area (SSA), and total pore volume (V) of materials studied.
SampleContent (wt.%)Ce/MnSSA (m2/g)V (cm3/g)
YCeMn
CeO2-80.6--400.14
Y/CeO22.778.6--320.13
Y-CeO21.879.5--430.16
MnOx--72.7-140.12
Y/MnOx3.3-66.8-90.10
Y-MnOx3.0-69.3-190.16
CeMnOx-58.922.61.2590.24
Y/CeMnOx2.856.821.81.2500.21
Y-CeMnOx2.157.819.01.2620.24
Table 2. Phase composition of samples and structural parameters of crystalline phases.
Table 2. Phase composition of samples and structural parameters of crystalline phases.
SamplePhase CompositionStructural ParametersDXRD, nm
Phasewt.%S.G.Symmetrya, Åb, Åc, Åβ, °
CeO2fluorite100Fm-3mcubic5.415.415.419018
Y/CeO2fluorite100Fm-3mcubic5.405.405.409018
Y-CeO2fluorite100Fm-3mcubic5.415.415.419015
MnOxMn3O475I41/amdtetragonal5.765.769.469033
Mn2O325Ia-3cubic9.419.419.419061
Y/MnOxMn3O434I41/amdtetragonal5.765.769.479048
Mn2O311Ia-3cubic9.429.429.429064
Mn5O855C12/m1monoclinic10.425.734.87109.924
Y-MnOxMn3O4100I41/amdtetragonal5.765.769.459028
CeMnOxfluorite100Fm-3mcubic5.385.385.38905
Y/CeMnOxfluorite100Fm-3mcubic5.385.385.38905
Y-CeMnOxfluorite100Fm-3mcubic5.385.385.38904
Table 3. Results of NH3-TPD for samples.
Table 3. Results of NH3-TPD for samples.
SampleT (°C)NH3 Desorbed (µmol g−1)
MnOx22552
Y-MnOx245105
Y/MnOx24089
CeMnOx235243
Y-CeMnOx225370
Y/CeMnOx280318
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

La Greca, E.; Kharlamova, T.S.; Grabchenko, M.V.; Svetlichnyi, V.A.; Pantaleo, G.; Consentino, L.; Stonkus, O.A.; Vodyankina, O.V.; Liotta, L.F. Influence of Y Doping on Catalytic Activity of CeO2, MnOx, and CeMnOx Catalysts for Selective Catalytic Reduction of NO by NH3. Catalysts 2023, 13, 901. https://doi.org/10.3390/catal13050901

AMA Style

La Greca E, Kharlamova TS, Grabchenko MV, Svetlichnyi VA, Pantaleo G, Consentino L, Stonkus OA, Vodyankina OV, Liotta LF. Influence of Y Doping on Catalytic Activity of CeO2, MnOx, and CeMnOx Catalysts for Selective Catalytic Reduction of NO by NH3. Catalysts. 2023; 13(5):901. https://doi.org/10.3390/catal13050901

Chicago/Turabian Style

La Greca, Eleonora, Tamara S. Kharlamova, Maria V. Grabchenko, Valery A. Svetlichnyi, Giuseppe Pantaleo, Luca Consentino, Olga A. Stonkus, Olga V. Vodyankina, and Leonarda Francesca Liotta. 2023. "Influence of Y Doping on Catalytic Activity of CeO2, MnOx, and CeMnOx Catalysts for Selective Catalytic Reduction of NO by NH3" Catalysts 13, no. 5: 901. https://doi.org/10.3390/catal13050901

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