Next Article in Journal
Forming a Cu-Based Catalyst for Efficient Hydrogenation Conversion of Starch into Glucose
Next Article in Special Issue
Influence of Particle Size of CeO2 Nanospheres Encapsulated in SBA-15 Mesopores on SO2 Tolerance during NH3-SCR Reaction
Previous Article in Journal
Regulating Lattice Oxygen on the Surfaces of Porous Single-Crystalline NiO for Stabilized and Enhanced CO Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improvement of NH3-SCR Performance by Exposing Different Active Components in a VCeMn/Ti Catalytic System

1
A State Key Laboratory of Featured Metal Materials and Life-Cycle Safety for Composite Structures, Guangxi University, Nanning 530004, China
2
Guangxi Colleges and Universities Key Laboratory of Applied Chemistry Technology and Resource Development, School of Chemistry and Chemical Engineering, Guangxi University, Nanning 530004, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(2), 131; https://doi.org/10.3390/catal14020131
Submission received: 9 January 2024 / Revised: 31 January 2024 / Accepted: 5 February 2024 / Published: 7 February 2024
(This article belongs to the Special Issue Rare Earth Catalysis: From Synthesis to Sustainable Applications)

Abstract

:
The physicochemical properties of active components play a key role in enhancing catalytic performance. In multi-component catalysts, different components offer a wide range of structural possibilities and catalytic potential. However, determining the role of specific components in enhancing efficiency may be blurry. This study synthetized a range of catalysts with various metal compositions on their external surfaces to investigate their catalytic activity on NH3-SCR. The V/CeMn/Ti catalysts exhibited exceptional catalytic efficiency and strong tolerance to SO2 during the SCR process. In the system, Mn and Ce facilitated electron transfer during the catalytic removal of NOx. As an assisting agent, increased the number of active species and acidic sites, playing a crucial role in oxidizing NO to NO2 and facilitating the denitrogenation reaction process at low temperatures. Further studies showed that the three ingredients exhibited unique adsorbent behaviors on the reacting gases, which provided different catalytic possibilities. This work modeled the particular catalysis of V and Ce (Mn) species, respectively, and offers experimental instruction for improving the activity and excellent tolerance to SO2 by controlling active ingredients.

1. Introduction

NH3-SCR is the most prevalent technology for effectively removing NOx [1,2]. Currently, several types of catalysts are widely used for the removal of NOx, including ion-exchanged zeolite catalysts, V-based catalysts, and other oxide catalysts [3,4]. V-based catalysts are widely used for their superior SO2 resistance [5,6,7]. V2O5-WO3(MoO3)/TiO2 catalysts are commonly used for commercial applications. However, such catalysts are unsuited for low-temperature flue gases in industries such as those of glass and steel, as they have weak low-temperature NH3-SCR activity and a confined operating temperature window [1,5,8,9]. While high V-loads can effectively enhance catalytic activity at low temperatures, they are also linked to several issues, including enhanced oxidation of SO2 to SO3, a narrow operating temperature window, poor alkali resistance, low N2 selectivity, and toxicity of V [1,6]. Therefore, the development of new V-based catalysts with a low vanadium loading and outstanding activity at low temperatures is important.
Anatase TiO2 was extensively used in commercial V-based catalysts for many years because of its outstanding tolerance to SO2 and its ability to improve the dispersion of vanadium-based materials [1,6]. Modifying the supports can effectively enhance the low-temperature activity and SO2 durability of V-based catalysts [7]. A loaded V2O5/Ce1−xTixO2 catalyst exhibited better NOx conversion and SO2 resistance than those of V2O5/TiO2 or V2O5/CeO2 at low temperatures after CeO2 was doped into TiO2, which was attributed to the improved surface dispersion of vanadium-based substances and higher reducibility [8]. According to the above results, V-oxides loaded on TiO2-containing mixed oxides consistently exhibited higher reactivity and excellent SO2 resistance in comparison with those loaded on pure Ti oxides. Additionally, mixed oxide loading may impact the surface active components. Therefore, further exploration of catalysts for low-temperature NH3-SCR reactions is necessary. Manganese-doped or loaded mixed oxides appear to be among the most prominent material candidates for reducing the temperature required for the NH3-SCR reaction [10,11,12]. The redox cycle of superficial Mn3+/Mn4+ or Mn2+/Mn3+ in manganese oxides (MnOx) is known for its high catalytic activity. Mn3+/Mn4+ or Mn2+/Mn3+ in manganese oxides (MnOx) is known for its high catalytic activity. It has been shown that Mn-Ce/TiO2 [13], MnOx/CeO2/AC [14], Mn-Ce-V/AC [15], and MnOx/SAPO-34 [16] have outstanding low-temperature NH3-SCR activity. For example, Yang et al. [17] investigated a MnOx-CeO2 catalyst synthesized through co-precipitation, which achieved a NO conversion of 95% at 150 °C. Huang and colleagues discovered that 10 wt% MnOx loaded on multi-walled carbon nanotubes (MWCNTs) in the range of 180–240 °C exhibited the best NO conversion [18]. However, the above debate was mainly concerned with activating NH3 and NO on the surface catalysts, but the functions of the different active ingredients in complicated catalytic systems remain unclear. Inspired by the above studies, by using graded impregnation to load V, Ce, and Mn on a TiO2 support, it is possible to extend the operating temperature window of the catalysts, improve the activity of NH3-SCR at low temperatures, and increase the resistance to H2O/SO2. This multicomponent approach not only optimizes the chemico-physical properties but also influences the adsorption and conversion of reactive gases via different pathways [19,20]. However, the roles of the components in the NH3-SCR reaction within such a complex system are intricate. Hence, it is essential and worthwhile to elucidate the roles of various ingredients in complicated catalytic systems.
In this study, a stepwise co-impregnation method was used to synthesize a series of catalysts with different exposed active components. The graded catalysts were characterized to investigate the effects of exposure to different active components on the NH3-SCR performance. Finally, we propose a possible mechanism for the enhancement of the SCR performance of V/MnCe/Ti catalysts. This study provides a guide to the experimental investigation of the catalytic potential of multi-composite V-based catalysts by elucidating the functions of various active species in catalysis.

2. Results and Discussion

2.1. Catalytic Performance

Figure 1a displays the results for the activity of various catalysts. The NO conversion decreased in the sequence of CeMn/V/Ti > V/CeMn/Ti > VCeMn/Ti at temperatures between 100 and 175 °C. This suggested that at low temperatures, the Ce (Mn) exposed on the outer surface of the catalysts was the dominant active species. Nevertheless, as the reaction temperature increased, the NO conversion was increased with respect to that of V/CeMn/Ti, and there was >90% NO conversion over a broad temperature range of 175–325 °C. With the further increase in the reaction temperature, the NO conversion slightly decreased because of the partial oxidation of NH3 [20]. Moreover, the NO conversion of the V/CeMn/Ti catalyst was higher than that of the CeMn/V/Ti and VCeMn/Ti catalysts at 175–325 °C, which indicated the synergistic effect between V and Ce (Mn). In summary, the V/CeMn/Ti catalysts had better NO conversion over the entire temperature range, which may have been associated with the good dispersion and reducibility of the active species of V exposed on the outer surface of the catalyst at medium and high temperatures.
The N2 selectivity results of the catalysts are displayed in Figure 1b. Compared with other previously reported high-activity catalysts, the V/CeMn/Ti, CeMn/V/Ti, and VCeMn/Ti catalysts exhibited high N2 selectivity over the entire range of testing temperatures, and for the CeMn/V/Ti and VCeMn/Ti samples, the N2 selectivity decreased from 100% to 80% from 150 to 325 °C, which may have been the result of the increase in the test temperature increases, while the NH3 adsorbed on the CeMn/V/Ti and VCeMn/Ti catalysts was more easily oxidized by O2, leading to the production of N2O. Interestingly, the N2 selectivity of the V/CeMn/Ti catalyst was maintained at 100% throughout the tested temperature range. N2O is one of the byproducts produced in the over-oxidation of NH3 at high temperatures. The above results indicate that the N2 selectivity was improved and the non-selective oxidation of NH3 was significantly suppressed when V was exposed to the outer surface of the catalysts.

H2O and SO2 Tolerance Tests

The catalysts’ resistance to H2O and SO2 was tested at 225 °C, as illustrated in Figure 2. The introduction of 5 vol.% H2O into the reaction system slightly reduced the activities of the V/CeMn/Ti and CeMn/V/Ti catalysts. The reduction could be explained by the competitive absorption of NH3 and H2O [21]. When the H2O was removed, the NOx conversion was restored, suggesting that the impact of H2O was reversible. However, the slight increase in NOx conversion for the VCeMn/Ti catalysts suggested that they were more effective in reducing NOx with NH3 in the presence of H2O. This is crucial for practical applications. The introduction of 50 ppm SO2 caused a decline in the NOx conversion of all catalysts, which may have been due to the reaction of SO2 with certain metal sites in the catalysts to form metal sulphates, which disrupted the redox SCR pathway of the catalysts, resulting in their irreversible inactivation [3,22]. If active sites are covered by surface metal sulfates or accumulated NH4HSO4 and become inactive, the catalyst lacks sufficient active sites [19]. The introduction of H2O + SO2 simultaneously for 2 h led to a significant decrease in NO conversion for all catalysts, which was attributed to the following: Firstly, there was a competitive adsorption of SO2 and excess H2O (5 vol%, 50,000 ppm) with NH3 and NO; secondly, SO2 + H2O could react directly with the reducing NH3 gas to produce ammonium bisulfate (NH4HSO4), which tended to inhibit the activity of the sites or be deposited on the surface of the catalysts. The NOx conversion of all catalysts did not return to the original values when SO2 and H2O were removed, indicating that the impact of SO2 + H2O on the catalysts was irreversible. The V/CeMn/Ti catalyst had superior tolerance for SO2 + H2O compared to that of the CeMn/V/Ti and VCeMn/Ti catalysts, which may have been because V covered the outside surface of the activated Ce (Mn) to prevent SO2 from interacting with the active compounds.

2.2. Structure and Morphology

The crystal structure of the catalysts was the subject of X-ray diffraction measurements, as shown in Figure 3. It was evident that the catalysts had similar 2θ diffraction locations but with various diffraction strengths. The diffraction peaks of 2θ at 25.2°, 37.6°, 48.2°, 52.9°, 55.3°, 62.6°, 68.7°, 70.4°, and 75.1° corresponded to the (101), (004), (200), (105), (211), (204), (116), (220), and (215) anatase TiO2 facets, respectively [21]; the diffraction peaks of 2θ at 27.5°corresponded to the rutile TiO2, and the diffraction peaks of the three catalysts at 2θ = 28.5° were the typical diffraction peaks of MnO2. The Scherrer equation was used to calculate the particle size of MnO2, and the particle sizes of MnO2 in the VCeMn/Ti, CeMn/V/Ti, and V/CeMn/Ti catalysts were 13.1472, 13.1480, and 13.1483 nm. The VCeMn/Ti catalyst had the smallest particle size, which indicated that the particles were reduced and the relative surface area was increased, which was in accordance with the conclusions from BET (Figure S1). No distinct diffraction peaks corresponding to the oxides of V and Ce were observed in the catalysts, which suggested that the doped metal oxides were present in an altitudinally dispersed state, which facilitated mutual contact with the activated ingredients and was conducive to strong reactions between the metal oxides [23]. SEM was used to observe the morphology of the catalysts, as shown in Figure S2. The morphologies of the three catalysts were similar, consisting of nanoparticles of varying sizes [23]. The results showed that the morphology of the catalysts was not affected by the order in which the active ingredients were loaded. The relevant BET surface areas and pore size distributions are summarized in Table 1; the catalysts exhibited “IV” adsorption–desorption isotherms with H3-type hysteresis loops, and their pore sizes were predominantly distributed between 2 and 50 nm, suggesting that the catalysts were formed from mesoporous structures [24,25]. The BET surface area declined in the order of V/CeMn/Ti ≈ CeMn/V/Ti > VCeMn/Ti, and the specific surface area of the catalysts that were synthetized through step-impregnation was larger than that of the catalysts that were synthetized through one-step impregnation. This could be attributed to the aggregation of MnO2 on the VCeMn/Ti catalyst during its preparation, leading to a reduction in its specific surface area (Figure S1).

2.3. Raman Analysis

The structural characteristics of V and Ce were studied with Raman spectroscopy. As shown in Figure 4, the weakly adsorbed peak at about 168 cm−1 was attributed to the B1g mode of anatase TiO2 [26,27]. No TiO2 peaks were detected for the VCeMn/Ti catalysts because of the presence of more of the MnO2 crystal phase on the surface, resulting in a weakening of its peak intensity. The results were consistent with those of XRD, and a prominent peak was centered at 450 cm−1, which suggested that CeO2 was present in the form of microcrystals [28,29], though this was undetectable in the XRD analysis. The peaks at 230 and 685 cm−1 were attributed to oxygen vacancies because of the presence of Ce3+ [29,30]. Oxygen vacancies enhance oxygen uptake and improve the efficiency of conversion between Ce3+ and Ce4+ [30]. Moreover, the strongest peak strengths were found for the V/CeMn/Ti catalysts, suggesting the presence of more oxygen vacancies that would promote NO oxidation and further facilitate the NH3-SCR reaction [31]. The peak at 1005 cm−1 was attributed to the monomeric V species. The peak at 960 cm−1 was attributed to the polymeric V species; according to the literature [32], the polymeric state of vanadium exhibits significantly higher activity than that of its monomeric state. Interestingly, the band at 960 cm–1 was the strongest for the V/CeMn/Ti catalyst, indicating that the V/CeMn/Ti catalyst had the best SCR performance, which was consistent with the results for the activity.

2.4. Acidic Site Distribution (NH3-TPD)

A catalyst’s surface acidity can influence the adsorption of NH3 and further impact the catalytic reaction. Figure 5a displays the NH3-TPD results. All catalysts were fitted into four desorption peaks in the range of 100–700 °C. The four peaks were the physical adsorption peak (at 100–150 °C; corresponding peak I) and the peaks of NH3 desorption on weak acids (Aw: 150–350 °C; corresponding peak II), medium-strong acids (Am: 400–550 °C; corresponding peak III), and strong acids (As: 550–700 °C; corresponding peak IV) [33,34]. Previous studies have reported [35,36] that NH4+ ions adsorbed at weak acid sites have lower thermal stability than that of NH3 adsorbed at strong acid sites. The peaks of desorption at lower temperatures were caused by the desorption of NH3 from the weak acid sites, and the peaks at higher temperatures were due to the desorption of NH3 from strong acid sites. Figure 5b displays the number of distinct acid sites. The analysis of the peak areas showed that the VCeMn/Ti catalyst had the smallest amount of acid compared to the other catalysts. However, the V/CeMn/Ti catalysts had a greater amount of acid—especially weak acids—than the CeMn/V/Ti and VCeMn/Ti catalysts. Therefore, the acidity of the catalyst was increased when the active component V was exposed to the outer surface of the catalyst.

2.5. The Oxidation–Reduction Properties of the Catalysts

The O2-TPD of the catalysts is shown in Figure 6a. Three O2 desorption peaks were observed in the range of 100–400 °C, and they were attributed to physically adsorbed oxygen (Op, around 200 °C), chemically adsorbed oxygen (Oc, around 250 °C), and lattice oxygen (surface lattice oxygen Os), respectively. The V/CeMn/Ti catalyst had the smallest adsorption capacity for Os, indicating that it had more oxygen vacancies, which led to more chemisorbed oxygen and was consistent with it having more Oc, which helped to upgrade its catalytic properties. The desorption peaks of the VCeMn/Ti catalyst were weaker, suggesting that it had a poorer capacity for storing oxygen. The area of the Oc peaks of the V/CeMn/Ti catalyst was much larger than those of the CeMn/V/Ti and VCeMn/Ti catalysts, which indicated that the oxygen storage capacity of the V/CeMn/Ti catalysts was much larger than that of the CeMn/V/Ti and VCeMn/Ti catalysts, and the exposure of active V species to the outer surface of the catalysts significantly increased the oxygen storage capacity, which was consistent with the Raman results.
The reductive characteristics of the catalysts were determined through H2-TPR experiments, as shown in Figure 6b. The catalysts had two large reduction peaks; the initial broad peak at 200 °C–400 °C was attributed to MnO2→Mn2O3→Mn3O4, and the second broad peak at 450 °C–550 °C was attributed to Mn3O4→MnO [37,38,39]. The reduction peak onset temperature of the V/CeMn/Ti catalyst was lower than those of the CeMn/V/Ti and VCeMn/Ti catalysts. This suggested that the VO-V species in the polymeric state improved the interactions between Mn and Ce, which made the active substances easier to reduce and greatly improved the surface oxygen mobility, thus improving the catalyst’s low-temperature activity [40]. There have been reports that this synergistic effect can lead to changes in the electron structures of the catalyst surface, as well as the generation of many oxygen vacancies, which facilitate the diffusion of surface oxygen into the bulk phase. These features contributed to the performance of NH3-SCR [41,42].

2.6. XPS Analysis

X-ray photoelectron spectroscopy (XPS) was used to study the chemical state of the surface elements of the catalysts. Figure 7 reveals the XPS spectra of Mn 2p, Ce 3d, O 1s, and V 2p, and the percentages are shown in Table 2. Figure 7a exhibits two dominant peaks of Mn 2p3/2 and Mn 2p1/2 at 633 eV and 665 eV. The Mn 2p3/2 spectrum was segmented into three distinctive peaks, which were assigned to Mn2+ (~640.7 eV), Mn3+ (~641.6V), and Mn4+ (~643.6 eV) [43,44]. Table 2 shows that the ratio of Mn4+/Mn2+ was significantly higher in V/CeMn/Ti (42.7%) than in CeMn/V/Ti (25.3%) and VCeMn/Ti (29.6%). Manganese oxide species with a higher valence state exhibited greater redox activity on Mn-based catalysts [44]. Furthermore, Mn4+ was shown to expedite the conversion of NO into NO2 and enhance the SCR reaction via the “fast SCR” pathway, which was in agreement with the catalytic performance of the catalysts.
As shown in Figure 7b, the Ce 3d XPS spectra of the catalysts were fitted with eight peaks: V (882.3 eV), V′ (884.2 eV), V″ (889.0 eV), V‴ (899.2 eV), U (901.2 eV), U′ (903.4 eV), U″ (906.5 eV), and U‴ (916.6 eV), which corresponded to four couples of spin–orbit doubles [44,45]. The peaks marked V and U belonged to the spin–orbit fractions of Ce3 d5/2 and Ce 3d3/2, respectively. The peaks denoted by V, V″, V‴, U, U″, and U‴ were attributed to characteristic peaks of Ce4+, whereas the peaks labeled V′ and U′ are attributed to Ce3+ [46,47,48]. The Ce 3d XPS spectra indicated the coexistence of Ce3+ and Ce4+ species on the catalysts, which were already shown to produce charge imbalances, oxygen vacancies, and unsaturated chemical bonds on the catalysts’ exterior, which were helpful for the formation of chemically adsorbed oxygen on the catalyst surface [44,46]. As demonstrated in Table 2, a higher proportion of Ce4+ (79%) was observed when the active V species were exposed to the outermost layer of the catalyst, indicating that V could boost the Mn3+ + Ce4+↔Mn4+ + Ce3+ oxidation–reduction cycle in the catalysts that were prepared in a stepwise manner, which corresponded to the results of O2-TPD.
Figure 7c shows the XPS spectra of O 1s. The deconvolution of the O 1s spectra produced three distinct peaks: a peak with a bond energy of about 528.9–530.4 eV was attributed to lattice oxygen (O2−, supplied as Oβ); a peak with a binding energy of 530.9–532.9 eV was attributed to the oxygen that was chemically adsorbed on the surface (O22− or O, supplied as Oα); meanwhile, a peak with a bond energy of about 529.9–535.6 eV was assigned to species of oxygen in hydroxyl groups (marked as Oγ) [49]. It was confirmed that the catalytic activity of oxygen adsorbed on a surface was better than that of lattice oxygen due to its high migration rate, which plays an important role in oxidation reactions [13]. As shown in Table 2, the percentage of Oα on the V/CeMn/Ti catalyst was higher than that on the CeMn/V/Ti and VCeMn/Ti catalysts, which suggested that the V/CeMn/Ti catalysts produced more oxygen vacancies that could promote the properties of NH3-SCR. The V 2p XPS results are shown in Figure 7d. The active V species in the three catalysts were predominantly in the form of V5+.
On the basis of the analysis of the above characterization results, a possible redox cycle on the V/CeMn/Ti catalyst was proposed. As shown in Figure 8, in the primary impregnation products of CeMn/Ti, there was a redox cycle in the form of Mn4+ + Ce3+↔Mn3+ + Ce4+. The coexistence of Ce3+ and Ce4+ species was shown to generate charge imbalances, oxygen vacancies, and unsaturated chemical bonds at the catalyst surface, and it facilitated the formation of chemisorbed oxygen on the catalyst surface [44,46]. The exposure of V to the outer surface of CeMn/Ti significantly improved the redox cycle and electron transfer, and it promoted NO/NH3 adsorption/activation.

2.7. In Situ DRIFTS

In this study, in situ DRIFTS was performed to illustrate the interaction of the surface active sites of NH3 and NO intermediates on the catalysts.

2.7.1. NH3 Adsorption–Desorption

In situ DRIFTS was conducted on three catalysts to elucidate their excellent performance. As shown in Figure 9, the peak located at ~1456 cm−1 was attributed to NH4+ ions adsorbed at the B-acid site. The peaks that emerged at ~1153 cm−1 and ~1293 cm−1 were assigned to NH3 species that were absorbed at the L-acid site [50]. The peaks observed at about ~1078 cm−1 and ~1530 cm−1 were assigned to NH2 species that were absorbed on the L-acid site and NH2 coordinate to the L-acid [51]. According to these results, L-acid and B-acid coexisted in the catalysts. Upon contrasting the peak intensities of all catalysts, it was discovered that the strongest peak intensities were those of NH4+ and NH3 species absorbed on the B-acid sites and L-acid sites in the V/CeMn/Ti catalyst, which implied that the V/CeMn/Ti catalyst had the strongest acidic sites when V was exposed to the outer surface of the catalyst, and this was consistent with the results for the activity.

2.7.2. NO + O2 Co-Adsorption

Figure 10 displays the in situ DRIFTS spectra of the adsorbed species for the co-adsorption of NO + O2 on the catalysts, and the presence of adsorbed NO2 species (~1602 cm−1), bidentate nitrate (1570–1578 cm−1), and NO2/NO3 (1217–1370 cm−1, ~1078 cm−1) was shown. The peaks for chelated nitrite and bridged nitrate overlapped [52]. The intensity of bidentate nitrate and monodentate nitrate initially increased and then decreased as the temperature increased. In contrast, the intensity of the nitrate peak increased with the increase in the temperature, which showed that nitrite was steady at 350 °C. The V/CeMn/Ti catalyst produced more bridged nitrate and bidentate nitrate species than the CeMn/V/Ti and VCeMn/Ti catalysts did. Thus, by combining the benefits of Ce (Mn) species, the V/CeMn/Ti catalyst’s adsorption capacity for NO and O2 was significantly improved, which led to a substantial quantity of nitrogen oxide being adsorbed onto the catalyst surface.

2.7.3. NH3 + NO + O2 Co-Adsorption

The catalysts’ reaction mechanism was investigated through in situ DRIFTS of NH3 + NO + O2 co-adsorption. As shown in Figure 11, adsorbed NO2 species (~1602 cm−1), bidentate nitrate (~1548 cm−1), -NH2 species formed through the coordination of NH3 to L-acid (1540 cm−1), NH4+ ions, and monodentate nitrite (1435 cm−1–1442 cm−1) adsorbed on B-acid sites were present. The NH3 species were found to be adsorbed on L-acid sites at ~1212 cm−1 and ~1289 cm−1. Overlapping peaks of NH2 and NO2/NO3 were also detected at 1352 cm−1, along with NH4NO2 at 1078 cm−1 [23,50,51,53]. The strength of NH4+ adsorbed on the B-acid sites in the V/CeMn/Ti catalyst increased and then decreased with the increase in the temperature, indicating that the B-acid sites were primarily involved in the catalytic reaction of NH3-SCR at low temperatures. The intensity of NH3 adsorbed at the L-acid site increased as the temperature increased, which was because the NH3 from the strong L-acid was chiefly desorbed at high temperatures and participated in the mega-temperature SCR reaction. The NO2 peak decreased as the temperature increased and disappeared when 200 °C was exceeded, indicating that NO2 primarily participated in the catalytic reaction at low temperatures. The peak intensities of bridging nitrate (NO3), NH2−, and NO2 species remained constant throughout the whole process, suggesting their involvement in the catalytic reaction. The above results suggested that the active V species were exposed to the external surface of the catalyst, inducing more acid sites, particularly B-acid sites. The strength of NH3 adsorption on the L-acid and B-acid sites of the V/CeMn/Ti catalyst was strongest at 150–300 °C. Similarly, the CeMn/V/Ti catalyst exhibited the strongest band strength at 50–150 °C. Surface acidity was shown to be a crucial factor in improving the low-temperature activity of the NH3-SCR reaction, which was in agreement with the NO conversion rate (refer to Figure 1). The results showed that the Langmuir–Hinshelwood (L−H) mechanism dominated the NH3-SCR reaction in the V/CeMn/Ti catalyst system.
Based on the in situ DRIFTS analyses, a feasible L−H reaction pathway is proposed in Figure 12 and Figure 13:

3. Experimental Section

3.1. Materials and Methods

V/CeMn/Ti, CeMn/V/Ti, and VCeMn/Ti catalysts were synthesized through stepwise co-impregnation. Taking the V/CeMn/Ti catalyst as an example, firstly, TiO2 (P25) was immersed in deionized water with completely dissolved Mn(NO3)2 (10 wt% MnO2) and Ce(NO3), as well as 6H2O (5.0 wt% CeO2), and stirred at room temperature for 1 h. Then, the turbid liquid was dried in an oil bath at 80 °C to remove the solvent and placed in an oven at 110 °C for 12 h. Finally, the solid powder was calcined at 400 °C for 4 h to obtain CeMn/Ti. The same method was used to load the outermost V species; 0.02 g of NH4VO3 (1.0 wt% V2O5) and 0.045 g of H2C2O4·2H2O were dissolved in 2 mL of water with constant stirring, and oxalic acid was added to promote the dissolution of NH4VO3. The solution was immersed in CeMn/Ti and stirred at room temperature for 1 h. Finally, the solid powder was calcined at 400 °C for 4 h to obtain a V/CeMn/Ti catalyst with exposed V species on the outer surface. Similarly, the graded CeMn/V/Ti and VCeMn/Ti catalysts were obtained by changing the order of the impregnants and exposing Ce Mn species and VCeMn on the outer surfaces of the catalysts (wt%:Mn:Ce = 2:1).

3.2. Catalyst Characterization

X-ray diffraction (XRD) was performed using a DX-2700A diffractometer (Dandong Kemait NDT Co.,Ltd , Dandong, China) under Cu–Kα radiation conditions with a voltage of 40 kV and a current of 30 mA. The sweep range was fixed at 10° to 80° with a sweep rate of 10°/min. We obtained N2 adsorption–desorption isotherms and N2 pore size distribution curves. The surface area (BET) values were calculated using the Brunauer–Emmet–Teller formula, and the pore size distributions were determined using the Barrett–Joyner–Halenda formula. A scanning electron microscope used to capture the SEM images was a Hitachi S-3400N (Hitachi, Tokyo, Japan). XPS was performed using a Thermo ESCALAB 250Xi+ electron spectrometer (Thermo Fisher Scientific, Waltham, MA USA) equipped with an –Kα X-ray source. C1s (284.8 eV) was used to correct the obtained results.
H2-TPR was conducted using a Fine-sorb-3010 chemisorption unit under the following conditions: pretreatment at 110 °C for 30 min in N2 atmosphere, pre-adsorption of 7 vol% H2 in Ar equilibrium gas for 30 min at room temperature, and a programmed temperature increase to 500 °C at a ramp-up rate of 10 °C/min. Raman spectrometry was carried out at room temperature using a Renishaw in-via reflection spectrometer fitted with an opto-microscope.
NH3 Program Thermal Desorption (NH3-TPD) was conducted using a Fine-sorb-3010 chemisorption unit, and the experimental parameters were the following: The pretreatment involved exposing the sample to a N2 atmosphere at a temperature of 300 °C for a duration of 60 min. Pre-adsorption was achieved by introducing an NH3 atmosphere (5 vol% NH3, N2 as equilibrium gas) at room temperature for 60 min.
In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) was performed using a Nicolet IS50 spectrometer (Thermo Fisher Scientific, Waltham, MA USA) with a resolution of 4 cm−1 and 32 scans. The catalysts were first pre-treated in an N2 atmosphere at 110 °C for 30 min and then cooled to room temperature to collect the background spectra. Reaction gases were introduced, and a start-up program was automatically programmed for the unit to heat up and record spectra under the following reaction conditions: NH3 (500 ppm), NO + O2 (NO = 500 ppm, O2 = 5 vol%), NH3 + NO + O2 (NH3 = 500 ppm, NO = 500 ppm, O2 = 5 vol%).
NH3-SCR activity tests were performed in a fixed-bed reactor, as were H2O and SO2 resistance tests, and the exhaust gas concentrations were recorded using an FGA10 flue gas analyzer. Gas conditions: NO = NH3 = 500 ppm, O2 = 5 vol%, H2O = 5 vol%, SO2 = 50 ppm. Reaction conditions: 40–60 mesh sample particles, and the N2-pretreated catalyst at 110 °C adsorbed the reaction gas for 60 min at room temperature. NOx conversion and N2 selectivity were determined with the following equations:
NO   conversion   ( % ) = [ N O ] i n [ N O ] o u t [ N O ] i n   × 100 %
N 2   selectivity   ( % ) = ( 1 2 [ N 2 O ] o u t ( [ N O x ] i n [ N O x ] o u t ) + ( [ N H 3 ] i n N H 3 o u t ) )   × 100 %
The subscripts “in” and “out” indicate the inlet (before the NH3-SCR reactor) and outlet (after the NH3-SCR reactor) gas concentrations of NO, respectively.

4. Conclusions

This study explored the effects of catalysts on NH3-SCR performance by exposing different active ingredients. The NH3-SCR properties of the V/CeMn/Ti catalyst were markedly better than those of the CeMn/V/Ti and VCeMn/Ti catalysts, and this catalyst also had a wider operating temperature range. In the V/CeMn/Ti catalyst, Mn and Ce facilitated electron transfer during the catalytic removal of NOx, while V, as an ancillary species, oxidated NO to NO2, increasing the number of active species and acidic sites. Furthermore, V was able to protect the active components of Mn and Ce by covering their outer surfaces and preventing interactions with SO2. This resulted in a stronger tolerance for H2O + SO2 while also increasing the number of surface acidic sites, adsorbed nitrites, and nitrate-attached ammonia species, thus facilitating the SCR reaction process. In situ DRIFTS studies showed that the V/CeMn/Ti catalyst followed the L-H mechanism.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14020131/s1, Figure S1. (a) N2 adsorption–desorption isotherms; (b) BJH pore size distribution curves; Figure S2. SEM images of catalysts: (a) V/CeMn/Ti; (b) CeMn/V/Ti; (c) VCeMn/Ti.

Author Contributions

S.G.: Conceptualization, Investigation, Methodology, Formal analysis, Writing—original draft, Data curation. C.H.: Resources, Formal analysis, Validation, Visualization. X.H.: Validation, Formal analysis, Software, Visualization. Q.Q.: Software, Visualization, Resources. D.M.: Formal analysis, Software. C.L.: Validation, Formal analysis. Y.Y.: Validation, Visualization. L.D.: Funding acquisition, Supervision. B.L.: Funding acquisition, Project administration, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 22265002 and 22062002).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

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

References

  1. 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] [PubMed]
  2. Chu, B.; Ma, Q.; Liu, J.; Ma, J.; Zhang, P.; Chen, T.; Feng, Q.; Wang, C.; Yang, N.; Ma, H.; et al. Air Pollutant Correlations in China: Secondary Air Pollutant Responses to NOx and SO2 Control. Environ. Sci. Technol. Lett. 2020, 7, 695–700. [Google Scholar] [CrossRef]
  3. Wu, X.; Yu, X.; He, X.; Jing, G. Insight into Low-Temperature Catalytic NO Reduction with NH3 on Ce-Doped Manganese Oxide Octahedral Molecular Sieves. J. Phys. Chem. C 2019, 123, 10981–10990. [Google Scholar] [CrossRef]
  4. Forzatti, P. Present status and perspectives in de-NOx SCR catalysis. Appl. Catal. A Gen. 2001, 222, 221–236. [Google Scholar] [CrossRef]
  5. Xu, J.; Chen, G.; Guo, F.; Xie, J. Development of wide-temperature vanadium-based catalysts for selective catalytic reducing of NOx with ammonia: Review. Chem. Eng. J. 2018, 353, 507–518. [Google Scholar] [CrossRef]
  6. Andana, T.; Rappé, K.G.; Gao, F.; Szanyi, J.; Pereira-Hernandez, X.; Wang, Y. Recent advances in hybrid metal oxide–zeolite catalysts for low-temperature selective catalytic reduction of NOx by ammonia. Appl. Catal. B Environ. 2021, 291, 120054. [Google Scholar] [CrossRef]
  7. Resitoglu, I.A.; Keskin, A. Hydrogen applications in selective catalytic reduction of NOx emissions from diesel engines. Int. J. Hydrogen Energy 2017, 42, 23389–23394. [Google Scholar] [CrossRef]
  8. Vuong, T.H.; Radnik, J.; Rabeah, J.; Bentrup, U.; Schneider, M.; Atia, H.; Armbruster, U.; Grünert, W.; Brückner, A. Efficient VOx/Ce1–xTixO2 Catalysts for Low-Temperature NH3-SCR: Reaction Mechanism and Active Sites Assessed by in Situ/Operando Spectroscopy. ACS Catal. 2017, 7, 1693–1705. [Google Scholar] [CrossRef]
  9. Chen, Z.; Wu, X.; Ni, K.; Shen, H.; Huang, Z.; Zhou, Z.; Jing, G. Molybdenum-decorated V2O5–WO3/TiO2: Surface engineering toward boosting the acid cycle and redox cycle of NH3-SCR. Catal. Sci. Technol. 2021, 11, 1746–1757. [Google Scholar] [CrossRef]
  10. Beck, B.; Harth, M.; Hamilton, N.G.; Carrero, C.; Uhlrich, J.J.; Trunschke, A.; Shaikhutdinov, S.; Schubert, H.; Freund, H.-J.; Schlögl, R.; et al. Partial oxidation of ethanol on vanadia catalysts on supporting oxides with different redox properties compared to propane. J. Catal. 2012, 296, 120–131. [Google Scholar] [CrossRef]
  11. Boningari, T.; Smirniotis, P.G. Impact of nitrogen oxides on the environment and human health: Mn-based materials for the NO x abatement. Curr. Opin. Chem. Eng. 2016, 13, 133–141. [Google Scholar] [CrossRef]
  12. Deng, S.; Meng, T.; Xu, B.; Gao, F.; Ding, Y.; Yu, L.; Fan, Y. Advanced MnOx/TiO2 Catalyst with Preferentially Exposed Anatase {001} Facet for Low-Temperature SCR of NO. ACS Catal. 2016, 6, 5807–5815. [Google Scholar] [CrossRef]
  13. Boningari, T.; Ettireddy, P.R.; Somogyvari, A.; Liu, Y.; Vorontsov, A.; McDonald, C.A.; Smirniotis, P.G. Influence of elevated surface texture hydrated titania on Ce-doped Mn/TiO2 catalysts for the low-temperature SCR of NO under oxygen-rich conditions. J. Catal. 2015, 325, 145–155. [Google Scholar] [CrossRef]
  14. Zhang, D.; Zhang, L.; Shi, L.; Fang, C.; Li, H.; Gao, R.; Huang, L.; Zhang, J. In situ supported MnOx–CeOx on carbon nanotubes for the low-temperature selective catalytic reduction of NO with NH3. Nanoscale 2013, 5, 1127. [Google Scholar] [CrossRef] [PubMed]
  15. Jiang, L.; Liu, Q.; Ran, G.; Kong, M.; Ren, S.; Yang, J.; Li, J. V2O5-modified Mn-Ce/AC catalyst with high SO2 tolerance for low-temperature NH3-SCR of NO. Chem. Eng. J. 2019, 370, 810–821. [Google Scholar] [CrossRef]
  16. Yu, C.; Huang, B.; Dong, L.; Chen, F.; Yang, Y.; Fan, Y.; Yang, Y.; Liu, X.; Wang, X. Effect of Pr/Ce addition on the catalytic performance and SO2 resistance of highly dispersed MnOx /SAPO-34 catalyst for NH3-SCR at low temperature. Chem. Eng. J. 2017, 316, 1059–1068. [Google Scholar] [CrossRef]
  17. 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]
  18. Wang, L.; Huang, B.; Su, Y.; Zhou, G.; Wang, K.; Luo, H.; Ye, D. Manganese oxides supported on multi-walled carbon nanotubes for selective catalytic reduction of NO with NH3: Catalytic activity and characterization. Chem. Eng. J. 2012, 192, 232–241. [Google Scholar] [CrossRef]
  19. Li, J.; Liu, H.; Qin, Q.; Zhu, J.; Chu, B.; Sun, Y.; Mo, D.; Dong, L.; Li, B. Functional distinction of different components in SmCeMn/Ti catalytic system for NH3-SCR. Appl. Surf. Sci. 2023, 636, 157775. [Google Scholar] [CrossRef]
  20. Liu, J.; Li, X.; Zhao, Q.; Ke, J.; Xiao, H.; Lv, X.; Liu, S.; Tadé, M.; Wang, S. Mechanistic investigation of the enhanced NH3-SCR on cobalt-decorated Ce-Ti mixed oxide: In situ FTIR analysis for structure-activity correlation. Appl. Catal. B Environ. 2017, 200, 297–308. [Google Scholar] [CrossRef]
  21. Chen, L.; Li, R.; Li, Z.; Yuan, F.; Niu, X.; Zhu, Y. Effect of Ni doping in NixMn1−xTi10 (x = 0.1–0.5) on activity and SO2 resistance for NH3-SCR of NO studied with in situ DRIFTS. Catal. Sci. Technol. 2017, 7, 3243–3257. [Google Scholar] [CrossRef]
  22. Ye, D.; Qu, R.; Zheng, C.; Cen, K.; Gao, X. Mechanistic investigation of enhanced reactivity of NH4HSO4 and NO on Nb- and Sb-doped VW/Ti SCR catalysts. Appl. Catal. A Gen. 2018, 549, 310–319. [Google Scholar] [CrossRef]
  23. Xie, S.; Li, L.; Jin, L.; Wu, Y.; Liu, H.; Qin, Q.; Wei, X.; Liu, J.; Dong, L.; Li, B. Low temperature high activity of M (M = Ce, Fe, Co, Ni) doped M-Mn/TiO2 catalysts for NH3-SCR and in situ DRIFTS for investigating the reaction mechanism. Appl. Surf. Sci. 2020, 515, 146014. [Google Scholar] [CrossRef]
  24. Zhang, L.; Qu, H.; Du, T.; Ma, W.; Zhong, Q. H2O and SO2 tolerance, activity and reaction mechanism of sulfated Ni–Ce–La composite oxide nanocrystals in NH3-SCR. Chem. Eng. J. 2016, 296, 122–131. [Google Scholar] [CrossRef]
  25. Goodman, A.L.; Li, P.; Usher, C.R.; Grassian, V.H. Heterogeneous Uptake of Sulfur Dioxide On Aluminum and Magnesium Oxide Particles. J. Phys. Chem. A 2001, 105, 6109–6120. [Google Scholar] [CrossRef]
  26. Zong, L.; Zhang, G.; Zhao, H.; Zhang, J.; Tang, Z. One pot synthesized CeO2-WO3-TiO2 catalysts with enriched TiO2 (0 0 1) facets for selective catalytic reduction of NO with NH3 by evaporation-induced self-assembly method. Chem. Eng. J. 2018, 354, 295–303. [Google Scholar] [CrossRef]
  27. Wang, X.; Shi, Y.; Li, S.; Li, W. Promotional synergistic effect of Cu and Nb doping on a novel Cu/Ti-Nb ternary oxide catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2018, 220, 234–250. [Google Scholar] [CrossRef]
  28. Li, L.; Zhang, L.; Ma, K.; Zou, W.; Cao, Y.; Xiong, Y.; Tang, C.; Dong, L. Ultra-low loading of copper modified TiO2/CeO2 catalysts for low-temperature selective catalytic reduction of NO by NH3. Appl. Catal. B Environ. 2017, 207, 366–375. [Google Scholar] [CrossRef]
  29. Cheng, X.; Zhang, X.; Su, D.; Wang, Z.; Chang, J.; Ma, C. NO reduction by CO over copper catalyst supported on mixed CeO2 and Fe2O3: Catalyst design and activity test. Appl. Catal. B Environ. 2018, 239, 485–501. [Google Scholar] [CrossRef]
  30. Zhang, G.; Han, W.; Zhao, H.; Zong, L.; Tang, Z. Solvothermal synthesis of well-designed ceria-tin-titanium catalysts with enhanced catalytic performance for wide temperature NH3-SCR reaction. Appl. Catal. B Environ. 2018, 226, 117–126. [Google Scholar] [CrossRef]
  31. Zhao, X.; Huang, L.; Li, H.; Hu, H.; Hu, X.; Shi, L.; Zhang, D. Promotional effects of zirconium doped CeVO 4 for the low-temperature selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2016, 183, 269–281. [Google Scholar] [CrossRef]
  32. He, G.; Lian, Z.; Yu, Y.; Yang, Y.; Liu, K.; Shi, X.; Yan, Z.; Shan, W.; He, H. Polymeric vanadyl species determine the low-temperature activity of V-based catalystsfor the SCR of NOx with NH3. Sci. Adv. 2018, 4, eaau4637. [Google Scholar] [CrossRef] [PubMed]
  33. Jia, X.; Liu, H.; Zhang, Y.; Chen, W.; Tong, Q.; Piao, G.; Sun, C.; Dong, L. Understanding the high performance of an iron-antimony binary metal oxide catalyst in selective catalytic reduction of nitric oxide with ammonia and its tolerance of water/sulfur dioxide. J. Colloid Interface Sci. 2021, 581, 427–441. [Google Scholar] [CrossRef] [PubMed]
  34. Fan, Z.; Wang, Z.; Shi, J.-W.; Gao, C.; Gao, G.; Wang, B.; Wang, Y.; Chen, X.; He, C.; Niu, C. Charge-redistribution-induced new active sites on (0 0 1) facets of α-Mn2O3 for significantly enhanced selective catalytic reduction of NO by NH3. J. Catal. 2019, 370, 30–37. [Google Scholar] [CrossRef]
  35. Zhou, X.; Huang, X.; Xie, A.; Luo, S.; Yao, C.; Li, X.; Zuo, S. V2O5-decorated Mn-Fe/attapulgite catalyst with high SO2 tolerance for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 326, 1074–1085. [Google Scholar] [CrossRef]
  36. Jia, B.; Guo, J.; Luo, H.; Shu, S.; Fang, N.; Li, J. Study of NO removal and resistance to SO2 and H2O of MnO/TiO2, MnO/ZrO2 and MnO/ZrO2–TiO2. Appl. Catal. A Gen. 2018, 553, 82–90. [Google Scholar] [CrossRef]
  37. Fang, D.; Xie, J.; Hu, H.; Yang, H.; He, F.; Fu, Z. Identification of MnO species and Mn valence states in MnO/TiO2 catalysts for low temperature SCR. Chem. Eng. J. 2015, 271, 23–30. [Google Scholar] [CrossRef]
  38. Zhang, P.; Sun, Y.; Su, W.; Wei, Y.; Liu, J. Low-temperature selective catalytic reduction of NO with NH3 over Ni–Mn–Ox catalysts. RSC Adv. 2016, 6, 107270–107277. [Google Scholar] [CrossRef]
  39. 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]
  40. Liu, S.-W.; Guo, R.-T.; Sun, X.; Liu, J.; Pan, W.-G.; Shi, X.; Wang, Z.-Y.; Liu, X.-Y.; Qin, H. Selective catalytic reduction of NOx over Ce/TiZrOx catalyst: The promoted K resistance by TiZrOx support. Mol. Catal. 2019, 462, 19–27. [Google Scholar] [CrossRef]
  41. Cai, S.; Zhang, D.; Zhang, L.; Huang, L.; Li, H.; Gao, R.; Shi, L.; Zhang, J. Comparative study of 3D ordered macroporous Ce0.75Zr0.2M0.05O2−δ(M = Fe, Cu, Mn, Co) for selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 2014, 4, 93–101. [Google Scholar] [CrossRef]
  42. Yu, J.; Si, Z.; Chen, L.; Wu, X.; Weng, D. Selective catalytic reduction of NO by ammonia over phosphate-containing Ce0.75Zr0.25O2 solids. Appl. Catal. B Environ. 2015, 163, 223–232. [Google Scholar] [CrossRef]
  43. Pappas, D.K.; Boningari, T.; Boolchand, P.; Smirniotis, P.G. Novel manganese oxide confined interweaved titania nanotubes for the low-temperature Selective Catalytic Reduction (SCR) of NOx by NH3. J. Catal. 2016, 334, 1–13. [Google Scholar] [CrossRef]
  44. Zhang, L.; Zhang, D.; Zhang, J.; Cai, S.; Fang, C.; Huang, L.; Li, H.; Gao, R.; Shi, L. Design of meso-TiO2@MnOx–CeOx/CNTs with a core–shell structure as DeNOx catalysts: Promotion of activity, stability and SO2-tolerance. Nanoscale 2013, 5, 9821. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, L.; Li, L.; Cao, Y.; Yao, X.; Ge, C.; Gao, F.; Deng, Y.; Tang, C.; Dong, L. Getting insight into the influence of SO2 on TiO2/CeO2 for the selective catalytic reduction of NO by NH3. Appl. Catal. B Environ. 2015, 165, 589–598. [Google Scholar] [CrossRef]
  46. Chen, H.; Xia, Y.; Huang, H.; Gan, Y.; Tao, X.; Liang, C.; Luo, J.; Fang, R.; Zhang, J.; Zhang, W.; et al. Highly dispersed surface active species of Mn/Ce/TiW catalysts for high performance at low temperature NH3-SCR. Chem. Eng. J. 2017, 330, 1195–1202. [Google Scholar] [CrossRef]
  47. Devaiah, D.; Smirniotis, P.G. Effects of the Ce and Cr Contents in Fe–Ce–Cr Ferrite Spinels on the High-Temperature Water–Gas Shift Reaction. Ind. Eng. Chem. Res. 2017, 56, 1772–1781. [Google Scholar] [CrossRef]
  48. Damma, D.; Boningari, T.; Smirniotis, P.G. High-temperature water-gas shift over Fe/Ce/Co spinel catalysts: Study of the promotional effect of Ce and Co. Mol. Catal. 2018, 451, 20–32. [Google Scholar] [CrossRef]
  49. Chen, Y.; Zhang, Z.; Liu, L.; Mi, L.; Wang, X. In situ DRIFTS studies on MnO nanowires supported by activated semi-coke for low temperature selective catalytic reduction of NO with NH3. Appl. Surf. Sci. 2016, 366, 139–147. [Google Scholar] [CrossRef]
  50. Xie, S.; Qin, Q.; Liu, H.; Jin, L.; Wei, X.; Liu, J.; Liu, X.; Yao, Y.; Dong, L.; Li, B. MOF-74-M (M = Mn, Co, Ni, Zn, MnCo, MnNi, and MnZn) for Low-Temperature NH3-SCR and In Situ DRIFTS Study Reaction Mechanism. ACS Appl. Mater. Interfaces 2020, 12, 48476–48485. [Google Scholar] [CrossRef]
  51. Liu, L.; Xu, K.; Su, S.; He, L.; Qing, M.; Chi, H.; Liu, T.; Hu, S.; Wang, Y.; Xiang, J. Efficient Sm modified Mn/TiO2 catalysts for selective catalytic reduction of NO with NH3 at low temperature. Appl. Catal. A Gen. 2020, 592, 117413. [Google Scholar] [CrossRef]
  52. Liu, S.; Wang, H.; Wei, Y.; Zhang, R. Core-shell structure effect on CeO2 and TiO2 supported WO3 for the NH3-SCR process. Mol. Catal. 2020, 485, 110822. [Google Scholar] [CrossRef]
  53. Wang, B.; Wang, M.; Han, L.; Hou, Y.; Bao, W.; Zhang, C.; Feng, G.; Chang, L.; Huang, Z.; Wang, J. Improved Activity and SO2 Resistance by Sm-Modulated Redox of MnCeSmTiOx Mesoporous Amorphous Oxides for Low-Temperature NH3-SCR of NO. ACS Catal. 2020, 10, 9034–9045. [Google Scholar] [CrossRef]
Figure 1. Catalytic performance of the catalysts. (a) NO conversion; (b) N2 selectivity (condition: NO = NH3 = 500 ppm, O2 = 5 vol%).
Figure 1. Catalytic performance of the catalysts. (a) NO conversion; (b) N2 selectivity (condition: NO = NH3 = 500 ppm, O2 = 5 vol%).
Catalysts 14 00131 g001
Figure 2. Comparison of H2O and SO2 resistance (reaction conditions: 220 °C, NO = NH3 = 500 ppm, O2 = 5 vol%, H2O = 5 vol%, SO2 = 50 ppm).
Figure 2. Comparison of H2O and SO2 resistance (reaction conditions: 220 °C, NO = NH3 = 500 ppm, O2 = 5 vol%, H2O = 5 vol%, SO2 = 50 ppm).
Catalysts 14 00131 g002
Figure 3. Catalysts’ XRD patterns.
Figure 3. Catalysts’ XRD patterns.
Catalysts 14 00131 g003
Figure 4. Raman spectroscopy of the catalysts.
Figure 4. Raman spectroscopy of the catalysts.
Catalysts 14 00131 g004
Figure 5. (a) NH3-TPD; (b) corresponding NH3 consumption.
Figure 5. (a) NH3-TPD; (b) corresponding NH3 consumption.
Catalysts 14 00131 g005
Figure 6. (a) O2-TPD and (b) H2-TPR for CeMn/V/Ti, V/CeMn/Ti, and VCeMn/Ti.
Figure 6. (a) O2-TPD and (b) H2-TPR for CeMn/V/Ti, V/CeMn/Ti, and VCeMn/Ti.
Catalysts 14 00131 g006
Figure 7. XPS for the (a) Mn 2p, (b) Ce 3d, (c) O 1s, and (d) V 2P orbitals of the catalysts.
Figure 7. XPS for the (a) Mn 2p, (b) Ce 3d, (c) O 1s, and (d) V 2P orbitals of the catalysts.
Catalysts 14 00131 g007
Figure 8. Schematic diagram of redox cycling for the V/CeMn/Ti catalysts.
Figure 8. Schematic diagram of redox cycling for the V/CeMn/Ti catalysts.
Catalysts 14 00131 g008
Figure 9. In situ DRIFTS spectra of NH3 adsorption–desorption: (a) V/CeMn/Ti, (b) CeMn/V/Ti, and (c) VCeMn/Ti.
Figure 9. In situ DRIFTS spectra of NH3 adsorption–desorption: (a) V/CeMn/Ti, (b) CeMn/V/Ti, and (c) VCeMn/Ti.
Catalysts 14 00131 g009
Figure 10. In situ DRIFTS spectra for the catalysts used in the co-adsorption of NO + O2: (a) V/CeMn/Ti, (b) CeMn/V/Ti, and (c) VCeMn/Ti.
Figure 10. In situ DRIFTS spectra for the catalysts used in the co-adsorption of NO + O2: (a) V/CeMn/Ti, (b) CeMn/V/Ti, and (c) VCeMn/Ti.
Catalysts 14 00131 g010
Figure 11. In situ DRIFTS spectra of the catalysts for NH3 + NO + O2 co-adsorption: (a) V/CeMn/Ti, (b) CeMn/V/Ti, and (c) VCeMn/Ti.
Figure 11. In situ DRIFTS spectra of the catalysts for NH3 + NO + O2 co-adsorption: (a) V/CeMn/Ti, (b) CeMn/V/Ti, and (c) VCeMn/Ti.
Catalysts 14 00131 g011
Figure 12. Adsorption and conversion processes of reactive gases in the V/CeMn/Ti catalytic system.
Figure 12. Adsorption and conversion processes of reactive gases in the V/CeMn/Ti catalytic system.
Catalysts 14 00131 g012
Figure 13. Model of the reaction mechanism (L-H) for different components of the V/CeMn/Ti catalytic system.
Figure 13. Model of the reaction mechanism (L-H) for different components of the V/CeMn/Ti catalytic system.
Catalysts 14 00131 g013
Table 1. Textural properties of all catalysts.
Table 1. Textural properties of all catalysts.
SamplesBET
(m2 · g−1)
Pore Volume
(cm3 · g−1)
Pore Size
(nm)
V/CeMn/Ti520.1711.29
CeMn/V/Ti510.1311.00
VCeMn/Ti380.1510.12
Table 2. The XPS results for O 1s, Mn 2p, and Ce 3d were compared for the different catalysts.
Table 2. The XPS results for O 1s, Mn 2p, and Ce 3d were compared for the different catalysts.
SamplesMn4+/Mnn+ (%)Ce4+/(Ce4+ + Ce3+) (%) Oα (%)
V/CeMn/Ti42.77949
CeMn/V/Ti25.37235
VCeMn/Ti29.66033
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

Gu, S.; Huang, C.; Han, X.; Qin, Q.; Mo, D.; Li, C.; You, Y.; Dong, L.; Li, B. Improvement of NH3-SCR Performance by Exposing Different Active Components in a VCeMn/Ti Catalytic System. Catalysts 2024, 14, 131. https://doi.org/10.3390/catal14020131

AMA Style

Gu S, Huang C, Han X, Qin Q, Mo D, Li C, You Y, Dong L, Li B. Improvement of NH3-SCR Performance by Exposing Different Active Components in a VCeMn/Ti Catalytic System. Catalysts. 2024; 14(2):131. https://doi.org/10.3390/catal14020131

Chicago/Turabian Style

Gu, Shifei, Chengheng Huang, Xiaorong Han, Qiuju Qin, Donghai Mo, Chen Li, Yuhua You, Lihui Dong, and Bin Li. 2024. "Improvement of NH3-SCR Performance by Exposing Different Active Components in a VCeMn/Ti Catalytic System" Catalysts 14, no. 2: 131. https://doi.org/10.3390/catal14020131

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