Next Article in Journal
Quo Vadis Dry Reforming of Methane?—A Review on Its Chemical, Environmental, and Industrial Prospects
Previous Article in Journal
Impact of Black Body Material Enhanced Gas Movement on CO2 Photocatalytic Reduction Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low-Temperature NH3-SCR on Cex-Mn-Tiy Mixed Oxide Catalysts: Improved Performance by the Mutual Effect between Ce and Ti

1
Key Laboratory for Advanced Materials, Research Institute of Industrial Catalysis, School of Chemistry and Molecular Engineering, East China University of Science and Technology, Shanghai 200237, China
2
International Joint Laboratory of Catalytic Chemistry, Department of Chemistry, Research Center of Nano Science and Technology, College of Sciences, Shanghai University, Shanghai 200444, China
3
State Key Laboratory of Photocatalysis on Energy and Environment College of Chemistry, Fuzhou University, Fuzhou 350116, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(5), 471; https://doi.org/10.3390/catal12050471
Submission received: 10 March 2022 / Revised: 20 April 2022 / Accepted: 21 April 2022 / Published: 22 April 2022

Abstract

:
A series of Cex-Mn-Tiy catalysts were synthesized using the coprecipitation method, and sodium carbonate solution was used as a precipitant. The various catalysts were assessed by selective catalytic reduction of NOx with NH3, and characterized by X-ray diffraction, Raman spectroscopy, H2 temperature-programmed reduction, NH3 temperature-programmed desorption, and X-ray photoelectron spectroscopy to investigate the physicochemical properties, surface acidity, and redox abilities of the Cex-Mn-Tiy catalysts. The Ce0.1-Mn-Ti0.1 catalyst exhibited the best catalytic performance (more than 90% NOx conversion in the range of 75 to 225 °C), as a result of proper redox ability, abundant acid sites, high content of Mn4+ and Ce3+, and surface-adsorbed oxygen (OS). The results of in situ DRIFT spectroscopy showed that the NH3-SCR reaction followed both the E-R and L-H paths over the Ce0.1-Mn-Ti0.1 catalyst, and it occurred faster and more sharply when it mainly abided by the E-R mechanism.

Graphical Abstract

1. Introduction

The combustion of fossil fuels, especially at thermoelectric power plants as stationary sources and in diesel vehicle emissions as mobile sources [1], produces excessive nitrogen oxides (including NO, NO2, and N2O), causing the greenhouse effect, acid rain, and photochemical smog [2,3,4]. In order to deal with the serious harm caused by NOx emissions, it is of great importance to find an effective approach to eliminate NOx.
The selective catalytic reduction of NOx with ammonia (NH3-SCR) is considered to be one of the most efficient ways to eliminate emissions of NOx [5]. As a reduction reagent, NH3 is used to reduce NOx, forming nitrogen and water. Owing to variable valence states, vanadium-based catalysts are currently the most commonly used catalysts in the industrial application of NH3-SCR. However, vanadium-based catalysts have some inevitable disadvantages, such as their narrow active temperature window (300–400 °C) and the toxicity of V2O5 [6]. In addition, under real operating conditions, the existence of sulfur and water in the flue gas of power plants can cause the deactivation of the catalyst. Therefore, a catalyst with excellent activity and resistance to SO2 and H2O is urgently needed.
In addition to vanadium-based catalysts, transition metal oxide catalysts, especially Mn oxides, exhibit great SCR activity, even at temperatures below 200 °C. However, pure MnOx catalysts show a narrow operating temperature window and poor tolerance to SO2 and H2O, restraining practical applications. Much effort has been devoted to improving the performance of the MnOx catalysts, for example, doping with other metals (M = Fe [7,8], Co [9,10], Cu [11,12], Sm [13,14], Ti [15,16], etc.) to optimize the redox capacity and acidity of the catalyst by forming M–Mn composite oxides. Nevertheless, it is hard to simultaneously optimize the redox capacity and acidity through single-metal doping, since the redox capacity and acidity restrict one another. Liu et al. prepared a series of MnyTi1-yOx composite oxides to adjust the surface acidity while the reducibility declined with a decrease in Mn content [17]. Conversely, when two metals are introduced into pure MnOx, the modified catalyst may possess a greater possibility of optimizing acidity and redox ability. CuwMnyTi1-yOx catalysts were synthesized to improve the acidity and redox ability simultaneously, and showed superior SCR performance and SO2 resistance [18]. Concurrently, Ce is known for its excellent oxygen storage and release capacity, which can significantly improve the oxidation of MnOx [19,20]. Previous studies [21] have found that the incorporation of Ti can efficiently tune the MnOx phase and construct surface acidity to promote the adsorption of reactants.
In this study, Cex-Mn-Tiy mixed-oxide catalysts were prepared via the coprecipitation method to investigate their catalytic performance, sulfur dioxide/water resistance, and structure–activity relationships. Mn4+ forms the main active sites for the adsorption of NO and NH3. Due to their high oxygen storage capacity and excellent redox properties, Ce and Mn cations can form a redox cycle [22]. The existence of Ti can efficiently decrease the MnOx crystallinity and construct surface acid sites to promote the adsorption of reactants. Cex-Mn-Tiy composite oxides show the best low-temperature activity at 75–225 °C, and better SO2/H2O resistance under a mixed flow of 2% H2O and 50 ppm of SO2 for 8 h at 180 °C. The effects of adding Ce and Ti to the structure and physicochemical properties of MnOx were systematically investigated through various characterizations, and the reaction mechanism on the Cex-Mn-Tiy catalyst was clarified.

2. Results and Discussion

2.1. Effects of Ce and Ti on Catalytic Activity and SO2/H2O Resistance

The low-temperature activity of Ce0.1-Mn-Ti0.1, Ce0.1-Mn, Mn-Ti0.1, and Ce0.1-Ti0.1 for NH3-SCR at different GHSVs (50,000, 100,000, and 150,000 h−1) is shown in Figure 1. The results showed that the Ce0.1MnTi0.1 catalyst exhibited the best catalytic performance, delivering ~90% NOx conversion at 75–225 °C. Meanwhile, the Ce0.1-Ti0.1 catalyst was almost inactive in the temperature range 50–350 °C, and the Mn-Ti0.1 catalyst exhibited significantly worse NH3-SCR activity than the Ce0.1-Mn-Ti0.1 and Ce0.1-Mn catalysts. The Ce0.1-Mn-Ti0.1 and Ce0.1-Mn catalysts showed similar activity curves at 50,000 h−1. As the GHSV increased from 50,000 to 150,000 h−1, the catalytic performance of the four catalysts decreased to some extent. However, the activity of the Ce0.1-Mn-Ti0.1 catalyst only slightly declined (125–225 °C with NOx conversion > 90%); in contrast, the activities of the Ce0.1-Mn and Mn0.1-Ti catalysts dropped significantly, seldom achieving 90% NOx conversion. Clearly, the Ce0.1-Mn-Ti0.1 catalyst exhibited the best catalytic activity. Hence, we conjectured that the synergistic effects between Ce and Mn as well as Mn and Ti [23] existed for the Ce0.1-Mn-Ti0.1 catalyst, dramatically enhancing its catalytic performance. NOx conversion per unit of SBET for 150,000 h−1 of the Cex-Mn-Tiy catalysts is shown in Figure S1 from Supplementary Materials.
Due to the poor activity of the Ce0.1-Ti0.1 catalyst, it was of little significance in testing the N2 selectivity. Thus, e Ce0.1-Ti0.1 catalyst was not included in the subsequent experiments. Therefore, Figure 1D only shows the N2 selectivity of the Ce0.1-Mn-Ti0.1, Ce0.1-Mn, and Mn-Ti0.1 catalysts, with a GHSV of 50,000 h−1. The N2 selectivity of the Ce0.1-Mn-Ti0.1 catalyst was almost above 95%, and the curve was similar to that of the Ce0.1-Mn catalyst, and higher than that of the Mn-Ti0.1 catalyst, at moderate and high temperatures (200–330 °C). Due to the relatively lower NOx conversion, the smallest amount of byproduct (N2O) was produced with the Mn-Ti0.1 catalyst.
The influence of H2O and SO2 is exhibited in Figure 2 for the Ce0.1-Mn-Ti0.1, Ce0.1-Mn, and Mn0.1-Ti catalysts. Firstly, the catalysts were kept in ideal conditions for 4 h at 180 °C. The Ce0.1-Mn-Ti0.1 and Ce0.1-Mn catalysts maintained complete conversion, while the Mn0.1-Ti catalyst only achieved 86% NOx conversion. When introducing 2 vol.% H2O into the feed gas for 8 h, the NOx conversion of the Ce0.1-Mn-Ti0.1, Ce0.1-Mn, and Mn0.1-Ti catalysts declined quickly, to approximately 93%, 85%, and 32% in almost an hour, respectively, and then remained stable. After 4 h of H2O removal, the catalytic activity was recovered to 96%, 96%, and 78%, respectively, illustrating that H2O poisoned the Cex-Mn-Tiy catalysts and caused irreversible deactivation. This may be attributable to the partial dissociation of H2O at 180 °C, and the –OH formed on the catalysts being difficult to desorb or decompose. Then, 2 vol.% H2O and 50 ppm of SO2 were introduced simultaneously for 10 h, and the conversion gradually dropped to 75%, 67%, and 19%, respectively, implying that the formed sulfates covered part of the active sites, resulting in the decrease in activity. Ce can act as a sacrificial site on the catalyst, and CeO2 preferentially reacts with SO2, alleviating the sulfation of the active Mn oxide phase. The introduction of Ce may induce the deposition of bulk sulfate on CeO2, thereby inhibiting the SO2 poisoning of the active MnOx sites. Once H2O and SO2 were removed, the conversion rebounded to about 96%, 90%, and 22%, respectively. It is worth noting that, after 10 h of the injection of 2 vol.% H2O and 50 ppm of SO2, the NOx conversion of Ce0.1-Mn showed a downward trend, while that of Ce0.1-Mn-Ti0.1 tended to stabilize. The overall performance within 2 vol.% H2O and the H2O/SO2 resistance of the Cex-Mn-Tiy catalysts at 100 °C are shown in Figures S2 and S3 from Supplementary Materials, respectively. The Ce-Mn-Ti catalyst exhibited a better catalytic performance at 2 vol.% H2O in a wide temperature range of 50 to 350 °C and a better H2O/SO2 resistance at 100 °C. To summarize, the Ce0.1-Mn-Ti0.1 catalyst possessed a better H2O/SO2 resistance than the Ce0.1-Mn and Mn-Ti0.1 catalysts. Hence, the introduction of small amounts of Ce and Ti led to the improvement of catalytic performance and SO2/H2O resistance.

2.2. Physicochemical Properties of Cex-Mn-Tiy Catalysts

2.2.1. X-ray Diffraction

Figure 3A shows the XRD patterns of all catalysts. In the XRD pattern of Ce0.1-Mn-Ti0.1, the diffraction peaks of MnO2, Mn5O8, and Mn2O3 are the main ones that can be seen [24,25]. For Ce0.1-Mn, however, in addition to the diffraction peaks of MnO2, Mn5O8, and Mn2O3, the diffraction peaks of CeO2 can also be seen [3,23], which indicates the enrichment of the Ce0.1-Mn surface by CeO2. Compared with Ce0.1-Mn, the XRD pattern of the Ce0.1-Mn-Ti0.1 catalyst showed weaker peak intensities and a smaller FWHM, ascribed to the MnO2 and Mn5O8. Owing to the similar ionic radius of Mn4+ and Ti4+, Ti4+ could be doped into the MnOx lattice [21], thereby inhibiting the crystallization of MnOx. The XRD pattern of Mn-Ti0.1 shows the diffraction peaks of MnO2, Mn5O8, Mn2O3, and TiO2 [26]. Comparing the XRD patterns of the Ce-Mn-Ti and Mn-Ti catalysts, the peak ascribed to TiO2 was not observed on Ce-Mn-Ti, while it appeared on the Mn-Ti catalyst. Thus, it can be inferred that the introduction of Ce can promote the dispersion of TiO2 on the Ce-Mn-Ti catalyst. The XRD patterns of the Ce0.1-Ti0.1 catalyst are displayed in Figure S4 from Supplementary Materials, showing only the diffraction of CeO2. These XRD results illustrated that the incorporation of Ti, inhibiting the MnOx crystallization, resulted in a low crystallinity of MnOx species, and Ce doping promoted the dispersion of Ti on the Ce0.1-Mn-Ti0.1 catalyst.

2.2.2. Raman Spectroscopy

Raman spectroscopy was used to further explain the composition and crystal phase of Cex-Mn-Tiy. The bands at 639, 642, 653, and 677 cm−1 were assigned to MnOx, and the bands at 150 and 456 cm−1 belonged to TiO2 and CeO2 [17,23], respectively (Figure 3B). It can be said that the Raman data agreed with the XRD results perfectly. In more detail, the bands at 639 and 642 cm−1 were assigned to MnO2 for the Ce0.1-Mn-Ti0.1 and Ce0.1-Mn catalysts, with a shift to a higher wavenumber [27]. Meanwhile, the intensity of the band at 639 cm−1 was much higher than that at 642 cm−1, representing more high-valence Mn. In addition, the band at 653 cm−1 was derived from the overlapping of the MnO2 and Mn2O3 phases, leading to a strong intensity. The band at 653 cm−1 with a shoulder band at 677 cm−1 was well-matched with the standard spectrum of Mn2O3 on the Mn-Ti0.1 catalyst [27]. That is, the Mn-Ti0.1 catalyst was more likely to exist in the phase of Mn2O3. Thus, the Ce0.1-Mn-Ti0.1 catalyst may have the highest concentration of high-valence Mn, which was later confirmed by XPS.

2.2.3. Scanning Electron Microscopy and N2 Adsorption–Desorption

The surfaces of the Ce0.1-Mn-Ti0.1, Ce0.1-Mn, and Mn-Ti0.1 catalysts were characterized by SEM. As shown in Figure 4A, Ce0.1-Mn-Ti0.1 was composed of many spherical particles, with a diameter of 1.5–2 μm. The shape of Ce0.1-Mn was similar to that of Ce0.1-Mn-Ti0.1, except that the particles were larger (Figure 4B). For Mn-Ti0.1, no spherical particles could be observed due to sintering (Figure 4C). These results implied that the incorporation of Ce could promote the dispersion of metals, thereby preventing the catalyst from sintering, in accordance with the XRD results. The pore structures and BET surface areas of the catalysts are given in Table 1 and Figure S5 from Supplementary Materials. The surface area of Ce0.1-Mn-Ti0.1 (52 m2/g), as well as its pore volume (0.23 cm3/g), were much larger than those of Ce0.1-Mn and Mn-Ti0.1. The mole ratios of Ce/Mn and Ti/Mn of the corresponding catalysts from ICP-AES were both 0.12—slightly higher than the theoretical value (0.1).

2.2.4. X-ray Photoelectron Spectroscopy

XPS was used to further analyze the valence states of Mn, Ce, O, and the surface atom concentrations. Figure 5 shows the XPS spectra of Mn 2p, Ce 3d, and O 1s of the Cex-Mn-Tiy catalysts. The surface atom concentrations are listed in Table 2. The Mn 2p spectra could be decomposed into three spin–orbit doublets, which are located at 642.2 ± 0.1, 641.3 ± 0.1, and 644 ± 0.4 eV, representing Mn4+, Mn3+, and satellites [28], respectively (Figure 5A). According to Table 2, the surface atom ratio of Mn4+/(Mn4++Mn3+) on the Ce0.1-Mn-Ti0.1 catalyst was the highest (60.1%) among the three catalysts, which is consistent with the Raman results. Compared with Mn-Ti0.1, the concentration of Mn4+ atoms increased significantly after Ce doping of the catalyst. The order of Mn4+ /(Mn4++Mn3+) was Ce0.1-Mn-Ti0.1 > Ce0.1-Mn > Mn-Ti0.1, which was consistent with the performance of NH3-SCR.
As shown in Figure 5B, the Ce 3d spectra could be decomposed into five spin–orbit doublets, denoted as u0, u, u′, u′′, u′′′ and v0, v, v′, v′′, v′′′ [29,30], where u0/v0 and u′/v′ belonged to Ce3+ and the others were assigned to Ce4+. It can be observed from Table 2 that the Ce3+/(Ce4++Ce3+) atom ratios on the Ce0.1-Mn-Ti0.1 and Ce0.1-Mn catalysts were 14.7 and 6.9%, respectively. The O 1s peaks displayed in Figure 5C could be split into two peaks, assigned to lattice oxygen (OL) at 529.7 ± 0.2 eV and surface active oxygen (OS) at 531.4 ± 0.1 eV [31]. It has been reported [32] that chemically adsorbed oxygen (OS) with higher mobility is more active than lattice oxygen (OL), and is a key factor affecting the catalytic activity. OS can oxidize NO to NO2, thus promoting the “Fast SCR” process [33]. Table 2 shows that the OS/(OS+OL) atomic ratio on the Ce0.1-Mn-Ti0.1 catalyst (16.6%) was the highest among the catalysts. As shown in Figure 5D, the binding energy of Ti 2p3/2 for Ce0.1-Mn-Ti0.1 and Mn-Ti0.1 was 458.3 and 458.1 eV, respectively, which can be ascribed to Ti4+ [18].
The above results illustrate that Ce0.1-Mn-Ti0.1 possessed the highest Mn4+/(Mn4++Mn3+) and OS/(OS+OL) atomic ratios. As a consequence, the Ce0.1-Mn-Ti0.1 catalyst exhibited the best NH3-SCR performance. In addition, the XPS data analysis resulted in an atomic ratio of Ce/Mn and Ti/Mn. The atomic ratios of Ce/Mn on the Ce0.1-Mn-Ti0.1 and Ce0.1-Mn catalysts were 0.29 and 0.32, respectively, while those for Ti/Mn on the Ce0.1-Mn-Ti0.1 and Mn-Ti0.1 catalysts were 0.14 and 0.16, respectively—higher than the molar ratio tested with ICP, suggesting that Ce and Ti were abundant on the surface of the Cex-Mn-Tiy catalysts.

2.3. Redox Ability and Acidity of the Cex-Mn-Tiy Catalysts

Both redox and acidic sites play important roles in the NH3-SCR reaction. Generally, the redox properties determine the low-temperature activity, while the acidity governs the high-temperature activity.

2.3.1. H2 Temperature-Programmed Reduction and NO Oxidation Reaction

H2-TPR tests were performed to judge the redox ability of the Cex-Mn-Tiy catalysts. As shown in Figure 6A, the three catalysts all exhibited two reduction peaks, denoted as α and β, respectively. The curve of the Ce0.1-Mn-Ti0.1 catalyst displays two apparent peaks at 338 and 467 °C, which were attributed to MnO2/Mn2O3 → Mn3O4 (α), Mn5O8/Mn3O4 → MnO, and CeO2 → Ce2O3 (β) [34,35,36], respectively (Table 3). The curves of Ce0.1-Mn and Mn-Ti0.1 were fairly similar to that of Ce0.1-Mn-Ti0.1. However, the reduction peaks of Ce0.1-Mn shifted to lower temperatures (328 and 434 °C), while those of Mn-Ti0.1 moved to higher temperatures (351 and 469 °C). Furthermore, the similar amounts of H2 consumption for the Mn-Ti0.1 and Ce0.1-Mn catalysts (6.58 and 6.56 mmol/g, respectively) were larger than that of the Ce0.1-Mn-Ti0.1 catalyst (5.35 mmol/g). H2 consumption originated from the reduction of Mn4+ and Ce4+. However, the Ce0.1-Mn-Ti0.1 catalyst is a tri-metal mixed oxide, and its Mn content was the lowest among the bimetal oxide catalysts with the same quality, contributing to a lower H2 consumption. In more detail, the area ratios of peaks α and β were ranked as Ce0.1-Mn > Mn-Ti0.1 > Ce0.1-Mn-Ti0.1, which seemed to be inconsistent with the results of XPS. The reason for the discrepancy between H2-TPR and XPS was that XPS characterized the surface atom concentration, while H2-TPR represented the overall reduction of the metal oxides. Consequently, though possessing the highest surface Mn4+ concentration (60.1%), the α/β of the Ce0.1-Mn-Ti0.1 catalyst was the lowest, implying that its surface was rich in Mn4+. The results obtained based on Table 3 indicated that Ce0.1-Mn showed the strongest reducibility, with lower reduction temperatures and a relatively larger amount of H2 consumption.
Figure 6B displays the NO oxidation capability of the Cex-Mn-Tiy catalysts under the reaction conditions of 500 ppm of NO and 5 vol.% O2. The order of NO oxidation capability of the three catalysts was Ce0.1-Mn > Ce0.1-Mn-Ti0.1 > Mn-Ti0.1, completely consistent with the reduction temperatures of H2-TPR, as shown in the illustration. Combined with H2-TPR, the redox ability of the Ce0.1-Mn catalyst was higher than that of the Ce0.1-Mn-Ti0.1 and Mn-Ti0.1 catalysts. In general, excessively strong redox properties led to non-selective catalytic oxidation of NH3 and the generation of N2O as a byproduct [37]. Nevertheless, the order of redox ability was inconsistent with the catalytic performance, implying that redox ability was not the key factor determining the NH3-SCR performance of the catalysts. Moderate redox ability has a beneficial effect on the SCR reaction, while excessive redox ability would be detrimental [23,38].

2.3.2. NH3/NO Temperature-Programmed Desorption

The adsorption capacity of NO and NH3 is an essential factor for evaluating catalysts. In order to characterize the adsorption capacity of NO and NH3, NH3/NO-TPD was carried out to analyze the properties of the Cex-Mn-Tiy catalysts.
Since the capacity for NH3 adsorption is one of the determinants for NH3-SCR catalysts, the surface acidity is of great importance for an excellent NH3-SCR catalyst. The area and position of the desorption peaks reflect the number and strength of acid sites, respectively. Before NH3-TPD was performed, the catalysts were pretreated at 400 °C in a flow of Ar for 1 h. The NH3-TPD profiles of all catalysts are shown in Figure 7A. In the NH3-TPD profiles of the Mn-Ti0.1 and Ce0.1-Mn catalysts, there was one weak desorption peak at 106 and 127 °C, respectively, and for the Ce0.1-Mn-Ti0.1 catalyst there was an obvious desorption peak that appeared at 89 °C. Compared with the Mn-Ti0.1 and Ce0.1-Mn catalysts, the desorption peak of the Ce0.1-Mn-Ti0.1 catalyst shifted slightly to a lower temperature, and the peak area per unit of SBET was high (183), as listed in Table 4, suggesting that the number of acid sites on the Ce0.1-Mn-Ti0.1 catalyst was larger than that of the Mn-Ti0.1 and Ce0.1-Mn catalysts. The number of acid sites clearly decreased in the order Ce0.1-Mn-Ti0.1 > Ce0.1-Mn > Mn-Ti0.1, which was consistent with the order of catalytic performance. It was speculated that the main factors affecting the number of acid sites may be the content of Mn4+ and the introduction of TiO2. Mn with a high valence state exhibits strong electron-adsorbing ability, and TiO2 is known to have abundant acidic sites [39]. Thus, combined with the high surface concentration of Mn4+ (60.1%) and the introduction of TiO2, the number of acidic sites on the Ce0.1-Mn-Ti0.1 catalyst increased significantly, improving its ability to adsorb NH3 and, finally, promoting the SCR activity of Ce0.1-Mn-Ti0.1.
NO-TPD was conducted on an activity evaluation device, which was used to quantitatively analyze the NO adsorption capacity of the Cex-Mn-Tiy catalysts. When the SCR reaction mainly follows the Langmuir–Hinshelwood (L-H) mechanism, the NO adsorption capacity plays a vital role in NH3-SCR. In the NO-TPD curves of the Cex-Mn-Tiy catalysts, there were two overlapped NO desorption peaks in two temperature regions of 150–250 and 250–500 °C (Figure 7B). Generally, the low-temperature desorption peak was attributed to desorption of physically adsorbed NO and decomposition of nitrite species, and the high-temperature desorption peak was ascribed to the decomposition of bridged nitrate species and bidentate nitrate species with higher thermal stability [27]. Furthermore, in comparison with the Mn-Ti0.1 and Ce0.1-Mn catalysts, the two desorption peaks of the Ce0.1-Mn-Ti0.1 catalyst shifted to a lower temperature, illustrating that the ability of NO to bind to adsorbed sites for the Ce0.1-Mn-Ti0.1 catalyst was weaker. To summarize, more gaseous NO on the Ce0.1-Mn-Ti0.1 catalyst can desorb and then be involved in the SCR reaction at lower temperatures, thereby improving the catalytic performance. Nevertheless, as listed in Table 4, the order of the adsorption capacity was Ce0.1-Mn (2.00 × 107) > Ce0.1-Mn-Ti0.1 (1.82 × 107) > Mn-Ti0.1 (1.57 × 107), indicating that the adsorption of NO occurs mainly over both Ce and Mn active sites. NO adsorbed on basic surfaces of Mn4+/Ce4+, forming NO intermediate species and Mn3+/Ce3+, while O2 adsorbed on Ce3+/Mn3+, generating active O and Ce4+/Mn4+. Then, NO intermediate species reacted with active O, forming adsorbed nitrate/nitrite species. The integrated areas of desorption peaks per unit of SBET were ranked as follows: Mn-Ti0.1 (5.3 × 105) > Ce0.1-Mn (4.3 × 105) > Ce0.1-Mn-Ti0.1 (3.5 × 105). These results indicate that the adsorption capacity per unit of specific surface area is not the decisive factor affecting the activity. In this work, acidity played a key role in catalytic activity.

2.4. In Situ DRIFT Spectroscopy

2.4.1. Adsorption of NO + O2 Followed by NH3

The Cex-Mn-Tiy catalysts were pretreated with Ar at 400 °C for 1 h to eliminate the surface-adsorbed H2O and CO2, and then NO + O2 co-adsorption in situ DRIFT spectroscopy analysis was carried out on the Cex-Mn-Tiy catalysts at 100 °C using 500 ppm of NO and 5 vol.% O2 to further investigate the adsorbed NOx species. In Figure 8A, it is clear that 1 min after the introduction of mixed gas (500 ppm of NO, 5% O2 and Ar for balance), there were two prominent vibration bands on Ce0.1-Mn-Ti0.1 at 1259 and 1217 cm−1, which were both assigned to bridged nitrates [29,40]. The bands at 1608 and 1394 cm−1, corresponding to bridged nitrates and trans-N2O22− species [9,41], appeared after 5 min. Moreover, with the increase in the adsorption time, the band intensity enlarged gradually after 60 min. Figure 8B displays several bands at 1568, 1456, 1396, 1338, and 1207 cm−1, showing that NOx species adsorbed on the surface of the Ce0.1-Mn catalyst. The band at 1207 cm−1 appeared after 1 min and disappeared after 10 min, assigned to bridged nitrates. The bands at 1396 and 1568 cm−1 could be assigned to trans-N2O22− and bidentate nitrates, while the bands at 1456 and 1338 cm−1 belonged to monodentate nitrates [27,42,43]. Referring to the Mn-Ti0.1 catalyst, it can be clearly seen from Figure 8C that several bands appeared. The bands at 1599, 1267, and 1225 cm−1 were recognized as bridged nitrates, and the bands at 1475, 1348, and 1055 cm−1 were related to monodentate nitrates [27]. The band at 1383 cm−1 was assigned to trans-N2O22− species. The band intensity increased rapidly on Ce0.1-Mn-Ti0.1, implying that adding Ti and Ce could enhance the adsorption ability of NO + O2 at 100 °C.
Prior to the introduction of 500 ppm of NH3, it is necessary to purge with Ar (50 mL/min) for 60 min to completely remove the residual mixed gas. After purging with Ar for 60 min, most strongly adsorbed bands remained. Figure 9A shows the in situ DRIFT spectra of the Ce0.1-Mn-Ti0.1 catalyst under NO + O2 adsorption conditions, followed by the introduction of NH3. The bands at 1259 and 1217 cm−1 assigned to bridged nitrates declined rapidly when continuously exposed to NH3, implying that bridged nitrates could take part in the NH3-SCR reaction. Surprisingly, a few new bands appeared; the band at 1660 cm−1 was assigned to NH3 coordinated to B acid sites, and the bands at 1595 and 1576 cm−1 were ascribed to NH3 coordinated on L acid sites [44,45]. The band at 1386 cm−1 was stronger than before, which may have been caused by the overlap of the bands. As mentioned in the literature [9], the band at 1380–1400 cm−1 could be NH3 adsorbed on B acid sites. The bands ascribed to trans-N2O22− species and NH3 adsorbed on B acid sites overlapped at 1386 cm−1. The in situ DRIFT spectra of Ce0.1-Mn are shown in Figure 9B; the intensity of adsorbed bands at 1568, 1456, and 1338 cm−1 all became weaker, illustrating that the ads-NOx species were slowly consumed by NH3. The band at 1575 was ascribed to NH3 coordinated on L acid sites, and the bands at 1647 and 1471 cm−1 were related to NH3 bonded to B acid sites [46,47]. Similarly, as for the Mn-Ti0.1 catalyst, the bands at 1599, 1571, 1377, and 1352 cm−1 after NH3 adsorption for 60 min were still observed (Figure 9C), implying that the ads-NOx species did not completely react with NH3. In addition, the band at 1599 cm−1 became stronger due to the NH3 adsorbed on L acid sites. As can be seen in Figure 8C and Figure 9C, the bands at 1267 and 1225 cm−1 disappeared after exposure to Ar for 60 min, demonstrating the weak adsorption of bridged nitrates. When NH3 and NO both adsorb on the surface and then react to generate N2 and H2O, the reaction path is called the L-H mechanism. Therefore, it can be inferred from the consumption of ads-NOx by NH3 that the Ce0.1-Mn-Ti0.1 catalyst exhibited the best catalytic performance at 100 °C, following the L-H path.

2.4.2. Adsorption of NH3 Followed by NO + O2

Similar to Section 2.4.1, after Ar pretreatment at 400 °C for 60 min to remove the interference of H2O and CO2, 500 ppm of NH3 was introduced to the Cex-Mn-Tiy catalysts to analyze the reactively adsorbed NH3 species in more detail. In Figure 10A, the band at 1572 cm−1 was assigned to NH4+ on L acid sites, while the bands at 1658 and 1469 cm−1 were attributed to NH3 corresponding to B acid sites of Ce0.1-Mn-Ti0.1. Additionally, the bands at 933 and 966 cm−1 were correlated with gaseous NH3 or weakly adsorbed NH3 [27]. The bands at 1377 and 1352 cm−1 were assigned to the oxidized species of the adsorbed ammonia species and amide (NH2) species bonded to L acid sites [48,49]. The peak intensities increased with the extension of adsorption time. In Figure 10B, for the Ce0.1-Mn catalyst, a series of peaks were observed at 1660, 1577, 1373, 964, and 924 cm−1. Similar to the Ce0.1-Mn-Ti0.1 catalyst, the peaks at 964 and 924 cm−1 were related to weakly adsorbed NH3 and gaseous NH3, respectively. The peaks at 1660 and 1577 cm−1 were related to NH4+ on B acid sites and NH3 coordinated on L acid sites, respectively, while the peak at 1373 cm−1 was the same as the peak at 1377 cm−1. As seen in Figure 10C, similar peaks appeared at 1655, 1599, 1577, 1468, 1377, 1352, 1176, 966, and 930 cm−1. The peaks at 1655 and 1468 cm−1 were assigned to NH4+ species on B acid sites, while the peaks at 1599, 1577, and 1176 cm−1 were attributed to NH3 on L acid sites [50]. Considering the three spectra comprehensively, the intensities of the NH3 adsorption peaks on Ce0.1-Mn-Ti0.1 were the largest—much higher than those of the Ce0.1-Mn and Mn-Ti0.1 catalysts—implying the strongest ability to adsorb NH3. With the increase in adsorption time, the intensities of the adsorption peaks reached the maximum value when injecting NH3 for 60 min.
Figure 11A–C show the in situ DRIFT spectroscopy spectra of Ce0.1-Mn-Ti0.1, Ce0.1-Mn, and Mn-Ti0.1 catalysts with pre-adsorbed NH3 followed by NO + O2, respectively. As shown in Figure 11A, the intensities of the peaks at 1658, 1572, 1469, 1377, and 1352 cm−1 declined gradually when exposed to NO + O2 for 60 min, suggesting that the above adsorbed NH3 species were active in the NH3-SCR reaction at 100 °C. A new peak appeared at 1273 cm−1 at 5 min, which was assigned to bridged nitrates, and increased with the exposure time, implying that NO could be adsorbed on the extra sites of the Ce0.1-Mn-Ti0.1 surface. When it came to the Ce0.1-Mn catalyst, the three peaks at 1660, 1574, and 1373 cm−1 were weakened, as shown in Figure 11B, meaning that the ads-NH3 reacted with gaseous NO. For the Mn-Ti0.1 catalyst, the intensities of the peaks at 1655, 1599, and 1572 cm−1 were slightly decreased, while the peaks at 1379 and 1352 cm−1 remained the same, as shown in Figure 11C. Additionally, there was a new weak peak at 1273 cm−1, attributed to bridged nitrates at 5 min. As adsorbed NH3 species reacting with gaseous NO, the reaction path is the E-R mechanism. Therefore, it can be concluded that the Ce0.1-Mn-Ti0.1, Ce0.1-Mn, and Mn-Ti0.1 catalysts all follow the E-R mechanism.
Moreover, in situ DRIFT spectroscopy was performed on Cex-Mn-Tiy catalysts at 50 °C, as displayed in Figures S6–S9, where a similar phenomenon was found. Combined with the results of both pre-adsorbed NH3 followed by NO + O2 and pre-adsorbed NO + O2 followed by NH3 at 100 °C, the NH3-SCR reaction is more likely to take place on the Ce0.1-Mn-Ti0.1 catalyst, thanks to its greater adsorption capacity for NH3 and NO + O2. On the one hand, in the process of pre-adsorbed NO and O2 followed by NH3, the disappearance of adsorbed NOx species (1259 and 1217 cm−1) demonstrated that the reaction between the active NOx species and the adsorbed NH3 followed the L-H mechanism. On the other hand, during the procedure of pre-adsorbed NH3 followed by injecting NO + O2, the peaks related to ads-NH3 (1658, 1572, 1469, 1377, and 1352 cm−1) were all obviously reduced as soon as NO + O2 was introduced into the feed, implying that the reaction can follow the E-R mechanism as well.

3. Materials and Methods

3.1. Catalyst Preparation

The Ce-Mn-Ti mixed-oxide catalysts were prepared using the coprecipitation method, as previously reported [27]. All the reagents used in the experiment were purchased from Sinopharm Corporation in Shanghai, China. MnSO4, Ce(NO3)3·6H2O, and Ti(SO4)2 were dissolved in 50 mL of deionized water at room temperature. After stirring for 30 min, the mixed solution and 0.2 M Na2CO3 were dropped simultaneously into a 500 mL beaker kept at pH 9 for precipitation. After this solution was stirred for 6 h, the precipitate was obtained by filtration and washing with H2O. The filter cake was dried overnight at 80 °C and, finally, calcined at 550 °C for 4 h. The prepared catalysts were named Cex-Mn-Tiy, where x represents the theoretical molar ratio of Ce/Mn (x = 0, 0.1), and y represents the theoretical molar ratio of Ti/Mn (y = 0, 0.1). Ce0.1-Mn, Mn-Ti0.1, and Ce0.1-Ti0.1 were all synthesized by the same method.

3.2. Catalyst Characterization

The XRD measurement was conducted using a Brook/D8 diffractometer (Brooks Automation, Chelmsford, MA, USA) employing Cu Kα radiation. The diffraction patterns were obtained in the 2θ range of 10 to 80°, with a step size of 0.02°. The crystalline phase was confirmed through comparison with the reference data from ICDD files. Raman spectroscopy was performed using a Via Reflex spectrophotometer (Renishaw, Gloucestershire, UK). The N2 adsorption–desorption isotherms were measured with a TriStar instrument (Micromeritics Instrument Corp.; Norcross, GA, USA). Before the measurements, all samples were degassed at 180 °C until a stable vacuum of ca. 5 mTorr was reached. The specific surface area was calculated from desorption data via the Brunauer–Emmett–Teller (BET) method. The X-ray photoelectron spectroscopy (XPS) spectra were obtained using a Thermo ESCALAB 250Xi spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) with Al Kα radiation. The binding energy of C 1s was 284.8 eV, which was defined as the reference.
Temperature-programmed desorption of NH3 (NH3-TPD) was conducted on a PX200 apparatus (Tianjin Pengxiang Technology Corporation, Tianjin, China) with a thermal conductivity detector (TCD). The catalyst (50 mg) was added to the quartz reactor and then pretreated at 400 °C in a flow of Ar (50 mL/min) for 1 h. When being cooled to room temperature, the sample was exposed to a flow of 10% NH3/Ar (50 mL/min) for 30 min. After the sample was purged with Ar (50 mL/min) for 1 h, NH3-TPD was carried out by heating the catalysts in Ar (50 mL/min) from room temperature to 400 °C at a rate of 10 °C/min.
Temperature-programmed desorption of NO (NO-TPD) was performed via the activity evaluation technique with a Thermo Fisher 42i-HL-NO-NOx analyzer (Thermo Fisher Scientific, Waltham, MA, USA) as the detector. The sample was pretreated in Ar (300 mL/min) at 400 °C for 1 h, and then cooled to room temperature. Then, the sample was exposed to a flow of 500 ppm of NO/Ar (300 mL/min) for 1 h (when the NOx on the samples reached a saturated state), followed by pure Ar (300 mL/min) purging for 1 h. Finally, NO-TPD was performed by heating the sample in Ar (300 mL/min) from room temperature to 500 °C at 10 °C/min.
Temperature-programmed reduction of H2 (H2-TPR) experiments were also conducted, using a VDSorb-92i-TPR apparatus (Tianjin Pengxiang Technology Corporation, Tianjin, China) with a TCD. The 50 mg sample was directly heated from room temperature to 800 °C at a rate of 10 °C/min, in a flow of 10 vol.% H2/N2 (40 mL/min). The hydrogen consumption was quantitatively evaluated by the TCD signal.
In situ DRIFT measurements were performed using a Nicolet 6700 FTIR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) with an MCT detector. The DRIFT cell contained ZnSe windows and a connected gas flow system. The sample was pretreated at 400 °C in Ar for 1 h, and then cooled to 50 °C in Ar. The background spectra were collected after the temperature became stable, and the background spectra were subtracted from the sample spectra accordingly.

3.3. Catalytic Activity Testing

The catalytic activities of the Cex-Mn-Tiy catalysts for NH3-SCR in excess oxygen were investigated at atmospheric pressure in a fixed-bed continuous-flow quartz reactor (inner diameter 6 mm). The amount of catalyst (40–60 mesh) used was 300 mg. The reactant gas was composed of 500 ppm of NO, 500 ppm of NH3, 5% O2, 2% H2O (when used), 50 ppm of SO2 (when used), and Ar as the balance gas. The gas hourly space velocity (GHSV) was about 50,000 h−1. The concentrations of NO and NO2 remaining in the reaction gases were analyzed using the NOx analyzer (Thermo Fisher 42i-HL-NO-NOx analyzer). Since the oxidation of ammonia in the converter of the NO/NOx analyzer may cause unexpected generation of NOx, an ammonia trap containing phosphoric acid solution was installed prior to the analyzer. N2O and NH3 were monitored using a Nicolet 6700 FT-IR spectrometer with an MCT detector. NOx conversion (x(NOx)) and N2 selectivity (S(N2)) were calculated as follows:
x ( NOx ) = C ( NO x ) in C ( NO x ) out C ( NO x ) in × 100 %
S ( N 2 ) = C ( NO x ) in + C ( NH 3 ) in C ( NO x ) out C ( NH 3 ) out 2 C ( N 2 O ) out C ( NO x ) in + C ( NH 3 ) in C ( NO x ) out C ( NH 3 ) out × 100 %
where C(NOx)in and C(NOx)out are the concentrations of NOx in the inlet and outlet, respectively, C(NH3)in and C(NH3)out are the concentrations of NH3 in the inlet and outlet, respectively, and C(N2O) is the concentration of N2O generated during the reaction process.

4. Conclusions

The Ce0.1-Mn-Ti0.1 catalyst with abundant acidic sites and proper redox ability reveals an outstanding ability to adsorb NH3 and NO, which gives rise to significant low-temperature activity (more than 90% NOx conversion in the range of 75 to 225 °C) and improved SO2/H2O resistance. Compared to the Ce0.1-Mn and Mn-Ti0.1 catalysts, the incorporation of Ti can restrain the MnOx crystallization, while Ce doping can promote the dispersion of Ti and further increase the specific surface area of the Ce0.1-Mn-Ti0.1 catalyst. On the basis of the results of XPS, the surface of the Ce0.1-Mn-Ti0.1 catalyst possessed the greatest contents of high-valence Mn ions (Mn4+), active adsorbed oxygen species (OS), and Ce3+. In addition, charge transfer can occur between Mn and Ce cations to promote the redox cycle. The results of in situ DRIFT spectroscopy reflected that the NH3-SCR reaction can take place simultaneously according to both the L-H and E-R paths. Hence, the Ce0.1-Mn-Ti0.1 catalyst can be regarded as a potential low-temperature catalyst for the industrial application of NH3-SCR.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12050471/s1: Figure S1: NOx conversion per unit of SBET under 150,000 h−1 over Cex-Mn-Tiy catalysts. Figure S2: H2O resistance under 50,000 h−1 over Cex-Mn-Tiy catalysts. Figure S3: H2O/SO2 resistance of Cex-Mn-Tiy catalysts at 100 °C. Figure S4: XRD patterns of the Ce0.1Ti0.1 catalyst. Figure S5: Pore size distribution of the Cex-Mn-Tiy catalysts. Figure S6: In situ DRIFTs of 500 ppm NO + 5 vol.% O2/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts at 50 °C. Figure S7: In situ DRIFTs of 500 ppm NH3/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts after adsorption of 500 ppm NO + 5 vol.% O2/Ar at 50 °C. Figure S8: In situ DRIFTs of 500 ppm NH3 adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts at 50 °C. Figure S9: In situ DRIFTs of 500 ppm NO + 5% O2/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts after adsorption of 500 NH3/Ar at 50 °C.

Author Contributions

Conceptualization, Q.Z.; methodology, Q.Z. and W.Z.; validation, A.W. and W.Z.; formal analysis, Q.Z.; investigation, Q.Z., W.Z. and A.W.; resources, Q.Z.; data curation, Q.Z.; writing—original draft preparation, Q.Z.; writing—review and editing, Q.Z., W.Z. and A.W.; visualization, Q.Z. and A.W.; project administration, W.Z.; funding acquisition, W.Z., J.Z., Y.G. (Yanglong Guo), Y.G. (Yun Guo) and L.W. 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 (21922602, U1905214, 22106101), the Shanghai Science and Technology Innovation Action Plan (20dz1204200), and the Fundamental Research Funds for the Central Universities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, H.L.; Ding, L.; Long, H.M.; Li, J.X.; Tan, W.; Ji, J.W.; Sun, J.F.; Tang, C.J.; Dong, L. Influence of CeO2 loading on structure and catalytic activity for NH3-SCR over TiO2-supported CeO2. J. Rare Earths 2020, 38, 883–890. [Google Scholar] [CrossRef]
  2. Han, M.J.; Jiao, Y.L.; Zhou, C.H.; Guo, Y.L.; Guo, Y.; Lu, G.Z.; Wang, L.; Zhan, W.C. Catalytic activity of Cu-SSZ-13 prepared with different methods for NH3-SCR reaction. Rare Met. 2018, 38, 210–220. [Google Scholar] [CrossRef]
  3. Zhang, N.Q.; Li, L.C.; Guo, Y.Z.; He, J.D.; Wu, R.; Song, L.Y.; Zhang, G.Z.; Zhao, J.S.; Wang, D.S.; He, H. A MnO2-based catalyst with H2O resistance for NH3-SCR: Study of catalytic activity and reactants-H2O competitive adsorption. Appl. Catal. B 2020, 270, 118860. [Google Scholar] [CrossRef]
  4. Sun, P.; Guo, R.T.; Liu, S.M.; Wang, S.X.; Pan, W.G.; Li, M.Y. The enhanced performance of MnOx catalyst for NH3-SCR reaction by the modification with Eu. Appl. Catal. A 2017, 531, 129–138. [Google Scholar] [CrossRef]
  5. Shao, X.Z.; Wang, H.Y.; Yuan, M.L.; Yang, J.; Zhan, W.C.; Wang, L.; Guo, Y.; Lu, G.Z. Thermal stability of Si-doped V2O5/WO3-TiO2 for selective catalytic reduction of NOx by NH3. Rare Met. 2018, 38, 292–298. [Google Scholar] [CrossRef]
  6. Han, L.P.; Cai, S.X.; Gao, M.; Hasegawa, J.Y.; Wang, P.L.; Zhang, J.P.; Shi, L.Y.; Zhang, D.S. 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]
  7. Ma, Y.; Li, W.; Wang, H.; Chen, J.; Wen, J.; Xu, S.; Tian, X.; Gao, L.; Hou, Z.; Zhang, Q.; et al. Enhanced performance of iron-cerium NO reduction catalysts by sulfuric acid treatment: The synergistic effect of surface acidity and redox capacity. Appl. Catal. A 2021, 621, 118200. [Google Scholar] [CrossRef]
  8. Chen, Y.X.; Li, C.; Chen, J.X.; Tang, X.F. Self-Prevention of Well-Defined-Facet Fe2O3/MoO3 against Deposition of Ammonium Bisulfate in Low-Temperature NH3-SCR. Environ. Sci. Technol. 2018, 52, 11796–11802. [Google Scholar] [CrossRef]
  9. Huang, L.; Hu, X.N.; Yuan, S.; Li, H.R.; Yan, T.T.; Shi, L.Y.; Zhang, D.S. Photocatalytic preparation of nanostructured MnO2-(Co3O4)/TiO2 hybrids: The formation mechanism and catalytic application in SCR deNOx reaction. Appl. Catal. B 2017, 203, 778–788. [Google Scholar] [CrossRef]
  10. Meng, D.M.; Xu, Q.; Jiao, Y.L.; Guo, Y.; Guo, Y.L.; Wang, L.; Lu, G.Z.; Zhan, W.C. Spinel structured CoaMnbOx mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2018, 221, 652–663. [Google Scholar] [CrossRef]
  11. Chen, Z.; Fan, C.; Pang, L.; Ming, S.J.; Liu, P.; Zhu, D.J.; Wang, J.Y.; Cai, X.; Chen, H.P.; Lai, Y.H.; et al. Direct synthesis of submicron Cu-SAPO-34 as highly efficient and robust catalyst for selective catalytic reduction of NO by NH3. Appl. Surf. Sci. 2018, 448, 671–680. [Google Scholar] [CrossRef]
  12. Huang, L.M.; Wang, X.M.; Yao, S.L.; Jiang, B.Q.; Chen, X.Y.; Wang, X. Cu-Mn bimetal ion-exchanged SAPO-34 as an active SCR catalyst for removal of NOx from diesel engine exhausts. Catal. Commun. 2016, 81, 54–57. [Google Scholar] [CrossRef]
  13. Meng, D.M.; Zhan, W.C.; Guo, Y.; Guo, Y.L.; Wang, Y.S.; Wang, L.; Lu, G.Z. A highly effective catalyst of Sm-Mn mixed oxide for the selective catalytic reduction of NOx with ammonia: Effect of the calcination temperature. J. Mol. Catal. A 2016, 420, 272–281. [Google Scholar] [CrossRef]
  14. Tong, Y.M.; Li, Y.S.; Li, Z.B.; Wang, P.Q.; Zhang, Z.P.; Zhao, X.Y.; Yuan, F.L.; Zhu, Y.J. Influence of Sm on the low temperature NH3-SCR of NO activity and H2O/SO2 resistance over the SmaMnNi2Ti7Ox (a = 0.1, 0.2, 0.3, 0.4) catalysts. Appl. Catal. A 2020, 590, 117333. [Google Scholar] [CrossRef]
  15. Hu, H.; Cai, S.X.; Li, H.R.; Huang, L.; Shi, L.Y.; Zhang, D.S. Mechanistic Aspects of deNOx Processing over TiO2 Supported Co-Mn Oxide Catalysts: Structure-Activity Relationships and In Situ DRIFTs Analysis. ACS Catal. 2015, 5, 6069–6077. [Google Scholar] [CrossRef]
  16. Anand, A.; Manjuladevi, M.; Veena, R.; Veena, V.; Koshkid’ko, Y.; Sagar, S. The influence of Ti doping at the Mn site on structural, magnetic, and magnetocaloric properties of Sm0.6Sr0.4MnO3. J. Solid State Chem. 2022, 305, 122712. [Google Scholar] [CrossRef]
  17. Liu, H.; Li, X.; Dai, Q.G.; Zhao, H.L.; Chai, G.T.; Guo, Y.L.; Guo, Y.; Wang, L.; Zhan, W.C. Catalytic oxidation of chlorinated volatile organic compounds over Mn-Ti composite oxides catalysts: Elucidating the influence of surface acidity. Appl. Catal. B 2021, 282, 119577. [Google Scholar] [CrossRef]
  18. Yan, Q.H.; Chen, S.N.; Zhang, C.; Wang, Q.; Louis, B. Synthesis and catalytic performance of Cu1Mn0.5Ti0.5Ox mixed oxide as low-temperature NH3-SCR catalyst with enhanced SO2 resistance. Appl. Catal. B 2018, 238, 236–247. [Google Scholar] [CrossRef]
  19. Li, L.; Li, P.; Tan, W.; Ma, K.; Zou, W.; Tang, C.; Dong, L. Enhanced low-temperature NH3-SCR performance of CeTiOx catalyst via surface Mo modification. Chin. J. Catal. 2020, 41, 364–373. [Google Scholar] [CrossRef]
  20. Yao, X.; Cao, J.; Chen, L.; Kang, K.; Chen, Y.; Tian, M.; Yang, F. Doping effect of cations (Zr4+, Al3+, and Si4+) on MnOx/CeO2 nano-rod catalyst for NH3-SCR reaction at low temperature. Chin. Catal. 2019, 40, 733–743. [Google Scholar] [CrossRef]
  21. Xu, Q.; Fang, Z.L.; Chen, Y.Y.; Guo, Y.L.; Guo, Y.; Wang, L.; Wang, Y.S.; Zhang, J.S.; Zhan, W.C. Titania-Samarium-Manganese composite oxide for the low-temperature selective catalytic reduction of NO with NH3. Environ. Sci. Technol. 2020, 54, 2530–2538. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, Z.; Zhu, J.; Li, J.; Ma, L.; Woo, S.I. Novel Mn-Ce-Ti mixed-oxide catalyst for the selective catalytic reduction of NOx with NH3. ACS Appl. Mater. Interfaces 2014, 6, 14500–14508. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, L.; Yao, X.J.; Cao, J.; Yang, F.M.; Tang, C.J.; Dong, L. Effect of Ti4+ and Sn4+ co-incorporation on the catalytic performance of CeO2-MnOx catalyst for low temperature NH3-SCR. Appl. Surf. Sci. 2019, 476, 283–292. [Google Scholar] [CrossRef]
  24. Jiang, H.X.; Wang, J.; Zhou, J.L.; Chen, Y.F.; Zhang, M.H. Effect of Promoters on the Catalytic Performance and SO2/H2O Resistance of α-MnO2 Catalysts for Low Temperature NH3-SCR. Ind. Eng. Chem. Res. 2019, 58, 1760–1768. [Google Scholar] [CrossRef]
  25. Shi, Y.R.; Yi, H.H.; Gao, F.Y.; Zhao, S.Z.; Xie, Z.L.; Tang, X.L. Facile synthesis of hollow nanotube MnCoOx catalyst with superior resistance to SO2 and alkali metal poisons for NH3-SCR removal of NOx. Sep. Purif. Technol. 2021, 265, 118517. [Google Scholar] [CrossRef]
  26. Fang, D.; He, F.; Liu, X.Q.; Qi, K.; Xie, J.L.; Li, F.X.; Yu, C.Q. Low temperature NH3-SCR of NO over an unexpected Mn-based catalyst: Promotional effect of Mg doping. Appl. Surf. Sci. 2018, 427, 45–55. [Google Scholar] [CrossRef]
  27. Ding, J.B.; Yang, X.W.; Wang, A.Y.; Yang, C.; Guo, Y.L.; Guo, Y.; Wang, L.; Zhan, W.C. Sm-MnOx catalysts for low-temperature selective catalytic reduction of NO with NH3: Effect of precipitation agent. J. Rare Earths 2021. [Google Scholar] [CrossRef]
  28. Wei, L.S.; Li, J.H.; Tang, X.F. NOx Storage at Low Temperature over MnOx-SnO2 Binary Metal Oxide Prepared Through Different Hydrothermal Process. Catal. Lett. 2008, 127, 107–112. [Google Scholar] [CrossRef]
  29. France, L.J.; Yang, Q.; Li, W.; Chen, Z.H.; Guang, J.Y.; Guo, D.W.; Wang, L.F.; Li, X.H. Ceria modified FeMnOx-Enhanced performance and sulphur resistance for low-temperature SCR of NOx. Appl. Catal. B Environ. 2017, 206, 203–215. [Google Scholar] [CrossRef]
  30. Skala, T.; Sutara, F.; Skoda, M.; Prince, K.C.; Matolin, V. Palladium interaction with CeO2, Sn-Ce-O and Ga-Ce-O layers. J. Phys. Condens Matter. 2009, 21, 055005. [Google Scholar] [CrossRef]
  31. Sun, P.; Huang, S.X.; Guo, R.T.; Li, M.Y.; Liu, S.M.; Pan, W.G.; Fu, Z.G.; Liu, S.W.; Sun, X.; Liu, J. The enhanced SCR performance and SO2 resistance of Mn/TiO2 catalyst by the modification with Nb: A mechanistic study. Appl. Surf. Sci. 2018, 447, 479–488. [Google Scholar] [CrossRef]
  32. Li, G.; Mao, D.S.; Chao, M.X.; Li, G.H.; Yu, J.; Guo, X.M. Low-temperature NH3-SCR of NO over MnCeOx/TiO2 catalyst: Enhanced activity and SO2 tolerance by modifying TiO2 with Al2O3. J. Rare Earths 2021, 39, 805–816. [Google Scholar] [CrossRef]
  33. Liu, Z.; Chen, C.; Zhao, J.C.; Yang, L.Z.; Sun, K.A.; Zeng, L.Y.; Pan, Y.; Liu, Y.Q.; Liu, C.G. Study on the NO2 production pathways and the role of NO2 in fast selective catalytic reduction DeNOx at low-temperature over MnOx/TiO2 catalyst. Chem. Eng. J. 2020, 379, 122288. [Google Scholar] [CrossRef]
  34. Thirupathi, B.; Smirniotis, P.G. Nickel-doped Mn/TiO2 as an efficient catalyst for the low-temperature SCR of NO with NH3: Catalytic evaluation and characterizations. J. Catal. 2012, 288, 74–83. [Google Scholar] [CrossRef]
  35. Gao, F.Y.; Yang, C.; Tang, X.L.; Yi, H.H.; Wang, C.Z. Co- or Ni-modified Sn-MnOx low-dimensional multi-oxides for high-efficient NH3-SCR De-NOx: Performance optimization and reaction mechanism. J. Environ. Sci. 2022, 113, 204–218. [Google Scholar] [CrossRef] [PubMed]
  36. Jiang, L.J.; Liang, Y.; Liu, W.Z.; Wu, H.L.; Aldahri, T.; Carrero, D.S.; Liu, Q.C. Synergistic effect and mechanism of FeOx and CeOx co-doping on the superior catalytic performance and SO2 tolerance of Mn-Fe-Ce/ACN catalyst in low-temperature NH3-SCR of NOx. J. Environ. Chem. Eng. 2021, 9, 106360. [Google Scholar] [CrossRef]
  37. Yang, S.; Qi, F.; Xiong, S.; Dang, H.; Liao, Y.; Wong, P.K.; Li, J. MnOx supported on Fe–Ti spinel: A novel Mn based low temperature SCR catalyst with a high N2 selectivity. Appl. Catal. B Environ. 2016, 181, 570–580. [Google Scholar] [CrossRef]
  38. Sun, C.; Liu, H.; Chen, W.; Chen, D.; Yu, S.; Liu, A.; Dong, L.; Feng, S. Insights into the Sm/Zr co-doping effects on N2 selectivity and SO2 resistance of a MnOx-TiO2 catalyst for the NH3-SCR reaction. Chem. Eng. J. 2018, 347, 27–40. [Google Scholar] [CrossRef]
  39. Sounak, R.; Viswanath, B.; Hegde, M.S.; Girigdar, M. Low-Temperature Selective Catalytic Reduction of NO with NH3 over Ti0.9M0.1O2−δ (M= Cr, Mn, Fe, Co, Cu). J. Phys. Chem. C 2008, 112, 6002–6012. [Google Scholar]
  40. Wang, X.B.; Duan, R.B.; Liu, W.; Wang, D.W.; Wang, B.R.; Xu, Y.R.; Niu, C.H.; Shi, J.W. The insight into the role of CeO2 in improving low-temperature catalytic performance and SO2 tolerance of MnCoCeOx microflowers for the NH3-SCR of NOx. Appl. Surf. Sci. 2020, 510, 145517. [Google Scholar] [CrossRef]
  41. Li, Y.L.; Han, X.J.; Hou, Y.Q.; Guo, Y.P.; Liu, Y.J.; Cui, Y.; Huang, Z.G. Role of CTAB in the improved H2O resistance for selective catalytic reduction of NO with NH3 over iron titanium catalyst. Chem. Eng. J. 2018, 347, 313–321. [Google Scholar] [CrossRef]
  42. Xiong, S.C.; Peng, Y.; Wang, D.; Huang, N.; Zhang, Q.F.; Yang, S.J.; Chen, J.J.; Li, J.H. The role of the Cu dopant on a Mn3O4 spinel SCR catalyst: Improvement of low-temperature activity and sulfur resistance. Chem. Eng. J. 2020, 387, 124090. [Google Scholar] [CrossRef]
  43. Guo, R.T.; Li, M.Y.; Sun, P.; Pan, W.G.; Liu, S.M.; Liu, J.; Sun, X.; Liu, S.W. Mechanistic Investigation of the Promotion Effect of Bi Modification on the NH3-SCR Performance of Ce/TiO2 Catalyst. J. Phys. Chem. C 2017, 121, 27535–27545. [Google Scholar] [CrossRef]
  44. Wang, B.; Wang, M.X.; Han, L.N.; Hou, Y.Q.; Bao, W.R.; Zhang, C.M.; Feng, G.; Chang, L.P.; Huang, Z.G.; Wang, J.C. 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]
  45. Wang, F.; Xie, Z.B.; Liang, J.S.; Fang, B.Z.; Piao, Y.A.; Hao, M.; Wang, Z.S. Tourmaline-Modified FeMnTiOx Catalysts for Improved Low-Temperature NH3-SCR Performance. Environ. Sci. Technol. 2019, 53, 6989–6996. [Google Scholar] [CrossRef]
  46. Sun, X.; Guo, R.T.; Liu, J.; Fu, Z.G.; Liu, S.W.; Pan, W.G.; Shi, X.; Qin, H.; Wang, Z.Y.; Liu, X.Y. The enhanced SCR performance of Mn/TiO2 catalyst by Mo modification: Identification of the promotion mechanism. Int. J. Hydrogen Energy 2018, 43, 16038–16048. [Google Scholar] [CrossRef]
  47. Gao, C.; Shi, J.W.; Fan, Z.Y.; Wang, B.R.; Wang, Y.; He, C.; Wang, X.B.; Li, J.; Niu, C.M. “Fast SCR” reaction over Sm-modified MnOx-TiO2 for promoting reduction of NOx with NH3. Appl. Catal. A Gen. 2018, 564, 102–112. [Google Scholar] [CrossRef]
  48. Chen, L.; Yang, J.; Ren, S.; Chen, Z.C.; Zhou, Y.H.; Liu, W.Z. Effects of Sm modification on biochar supported Mn oxide catalysts for low-temperature NH3-SCR of NO. J. Energy Inst. 2021, 98, 234–243. [Google Scholar] [CrossRef]
  49. Qiu, M.Y.; Zhan, S.H.; Yu, H.B.; Zhu, D.D.; Wang, S.Q. Facile preparation of ordered mesoporous MnCo2O4 for low-temperature selective catalytic reduction of NO with NH3. Nanoscale 2015, 7, 2568–2577. [Google Scholar] [CrossRef]
  50. Guo, M.Y.; Zhao, P.P.; Liu, Q.L.; Liu, C.X.; Han, J.F.; Ji, N.; Song, C.F.; Ma, D.G.; Lu, X.B.; Liang, X.Y.; et al. Improved Low-Temperature Activity and H2O Resistance of Fe-Doped Mn-Eu Catalysts for NO Removal by NH3-SCR. ChemCatChem 2019, 11, 4954–4965. [Google Scholar] [CrossRef]
Figure 1. NOx conversions with different GHSVs = 50,000 h−1 (A), 100,000 h−1 (B), and 150,000 h−1 (C) and N2 selectivity (D) for Cex-Mn-Tiy catalysts. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, with Ar as the balanced gas.
Figure 1. NOx conversions with different GHSVs = 50,000 h−1 (A), 100,000 h−1 (B), and 150,000 h−1 (C) and N2 selectivity (D) for Cex-Mn-Tiy catalysts. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, with Ar as the balanced gas.
Catalysts 12 00471 g001
Figure 2. H2O/SO2 resistance of Cex-Mn-Tiy catalysts at 180 °C. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, 50 ppm SO2 (when needed), 2 vol.% H2O (when needed), with Ar as the balanced gas and GHSV = 50,000 h−1.
Figure 2. H2O/SO2 resistance of Cex-Mn-Tiy catalysts at 180 °C. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, 50 ppm SO2 (when needed), 2 vol.% H2O (when needed), with Ar as the balanced gas and GHSV = 50,000 h−1.
Catalysts 12 00471 g002
Figure 3. XRD patterns (A) and Raman spectra (B) of Cex-Mn-Tiy catalysts.
Figure 3. XRD patterns (A) and Raman spectra (B) of Cex-Mn-Tiy catalysts.
Catalysts 12 00471 g003
Figure 4. SEM images of the Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts.
Figure 4. SEM images of the Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts.
Catalysts 12 00471 g004
Figure 5. XPS Mn 2p (A), Ce 3d (B), O 1s (C), and Ti 2p (D) spectra of the Cex-Mn-Tiy catalysts.
Figure 5. XPS Mn 2p (A), Ce 3d (B), O 1s (C), and Ti 2p (D) spectra of the Cex-Mn-Tiy catalysts.
Catalysts 12 00471 g005
Figure 6. H2-TPR profiles (A) and the activity of NO oxidation (B) for the Cex-Mn-Tiy catalysts (the reaction conditions of NO oxidation: [NO] = 500 ppm, [O2] = 5 vol.%, with Ar as balanced, total flow rate 300 mL/min and GHSV = 50,000 h−1).
Figure 6. H2-TPR profiles (A) and the activity of NO oxidation (B) for the Cex-Mn-Tiy catalysts (the reaction conditions of NO oxidation: [NO] = 500 ppm, [O2] = 5 vol.%, with Ar as balanced, total flow rate 300 mL/min and GHSV = 50,000 h−1).
Catalysts 12 00471 g006
Figure 7. NH3-TPD (A) and NO-TPD (B) profiles of the Cex-Mn-Tiy catalysts (the reaction conditions of NH3-TPD: [NH3] = 10 vol.%, Ar as balanced, total flow rate 50 mL/min. The reaction conditions of NO-TPD: [NO] = 500 ppm, Ar as balanced, total flow rate 300 mL/min and GHSV = 50,000 h−1).
Figure 7. NH3-TPD (A) and NO-TPD (B) profiles of the Cex-Mn-Tiy catalysts (the reaction conditions of NH3-TPD: [NH3] = 10 vol.%, Ar as balanced, total flow rate 50 mL/min. The reaction conditions of NO-TPD: [NO] = 500 ppm, Ar as balanced, total flow rate 300 mL/min and GHSV = 50,000 h−1).
Catalysts 12 00471 g007
Figure 8. In situ DRIFT spectroscopy of 500 ppm NO + 5 vol% O2/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts at 100 °C.
Figure 8. In situ DRIFT spectroscopy of 500 ppm NO + 5 vol% O2/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts at 100 °C.
Catalysts 12 00471 g008
Figure 9. In situ DRIFT spectroscopy of 500 ppm NH3/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts after adsorption of 500 ppm of NO + 5 vol% O2/Ar at 100 °C.
Figure 9. In situ DRIFT spectroscopy of 500 ppm NH3/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts after adsorption of 500 ppm of NO + 5 vol% O2/Ar at 100 °C.
Catalysts 12 00471 g009
Figure 10. In situ DRIFT spectroscopy of 500 ppm NH3 adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts at 100 °C.
Figure 10. In situ DRIFT spectroscopy of 500 ppm NH3 adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts at 100 °C.
Catalysts 12 00471 g010
Figure 11. In situ DRIFT spectroscopy of 500 ppm NO + 5% O2/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts after adsorption of 500 NH3/Ar at 100 °C.
Figure 11. In situ DRIFT spectroscopy of 500 ppm NO + 5% O2/Ar adsorption on Ce0.1-Mn-Ti0.1 (A), Ce0.1-Mn (B), and Mn-Ti0.1 (C) catalysts after adsorption of 500 NH3/Ar at 100 °C.
Catalysts 12 00471 g011
Table 1. BET surface area and pore structure of the Cex-Mn-Tiy catalysts.
Table 1. BET surface area and pore structure of the Cex-Mn-Tiy catalysts.
SampleSBET (m2/g)Pore Size (nm)Pore Volume (cm3/g)Mole Ratio a
Ce/MnTi/Mn
Ce0.1-Mn-Ti0.152.4617.70.230.120.12
Ce0.1-Mn46.6116.90.200.12-
Mn-Ti0.129.4525.10.18-0.12
a Molar ratio of Ce/Mn and Ti/Mn according to ICP-AES.
Table 2. The XPS results and atomic molar ratios of the Cex-Mn-Tiy catalysts.
Table 2. The XPS results and atomic molar ratios of the Cex-Mn-Tiy catalysts.
SampleSurface Atom Concentration b
Ce/MnTi/MnMn4+/(Mn4+ + Mn3+)OS/(OS + OL)Ce3+/(Ce4+ + Ce3+)
Ce0.1-Mn-Ti0.10.290.1460.1%16.6%14.7%
Ce0.1-Mn0.32-58.9%12.6%6.9%
Mn-Ti0.1-0.1655.7%13.2%-
b Surface atom concentration detected by XPS analysis.
Table 3. H2-TPR results of the Cex-Mn-Tiy catalysts.
Table 3. H2-TPR results of the Cex-Mn-Tiy catalysts.
Sampleα (T/°C)Curve Area
β (T/°C)
α/βH2 Consumption (mmol/g)
Ce0.1-Mn-Ti0.12.79 (338)2.56 (467)1.085.35
Ce0.1-Mn3.65 (328)2.91 (434)1.256.56
Mn-Ti0.13.46 (351)3.12 (469)1.116.58
Table 4. NH3-TPD and NO-TPD results.
Table 4. NH3-TPD and NO-TPD results.
SampleCurve Area of NH3-TPD × 103Area/SBET(NH3)Curve Area of NO-TPD × 107Area/SBET(NO) × 105
Ce0.1-Mn-Ti0.19.61831.83.5
Ce0.1-Mn6.01292.04.3
Mn-Ti0.13.51181.65.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhu, Q.; Wang, A.; Zhang, J.; Guo, Y.; Guo, Y.; Wang, L.; Zhan, W. Low-Temperature NH3-SCR on Cex-Mn-Tiy Mixed Oxide Catalysts: Improved Performance by the Mutual Effect between Ce and Ti. Catalysts 2022, 12, 471. https://doi.org/10.3390/catal12050471

AMA Style

Zhu Q, Wang A, Zhang J, Guo Y, Guo Y, Wang L, Zhan W. Low-Temperature NH3-SCR on Cex-Mn-Tiy Mixed Oxide Catalysts: Improved Performance by the Mutual Effect between Ce and Ti. Catalysts. 2022; 12(5):471. https://doi.org/10.3390/catal12050471

Chicago/Turabian Style

Zhu, Qianwen, Aiyong Wang, Jinshui Zhang, Yanglong Guo, Yun Guo, Li Wang, and Wangcheng Zhan. 2022. "Low-Temperature NH3-SCR on Cex-Mn-Tiy Mixed Oxide Catalysts: Improved Performance by the Mutual Effect between Ce and Ti" Catalysts 12, no. 5: 471. https://doi.org/10.3390/catal12050471

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