Next Article in Journal
The Isocyanurate-Carbamate-Bridged Hybrid Mesoporous Organosilica: An Exceptional Anchor for Pd Nanoparticles and a Unique Catalyst for Nitroaromatics Reduction
Next Article in Special Issue
Enhanced Catalytic Soot Oxidation by Ce-Based MOF-Derived Ceria Nano-Bar with Promoted Oxygen Vacancy
Previous Article in Journal
Isoquinolone Syntheses by Annulation Protocols
Previous Article in Special Issue
Investigation of Solid Deposit Inside L-Type Urea Injector and NOx Conversion in a Heavy-Duty Diesel Engine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insight into the Promoting Role of Er Modification on SO2 Resistance for NH3-SCR at Low Temperature over FeMn/TiO2 Catalysts

1
Marine Engineering College, Dalian Maritime University, Dalian 116026, China
2
Liaoning Research Center for Marine Internal Combustion Engine Energy-Saving, Marine Engineering College, Dalian Maritime University, Dalian 116026, China
3
School of Naval Architecture and Ocean Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
4
School of Electronic and Information Technology, Guangdong Ocean University, Zhanjiang 524088, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work and should be considered co-first authors.
Submission received: 13 April 2021 / Revised: 8 May 2021 / Accepted: 10 May 2021 / Published: 11 May 2021

Abstract

:
Er-modified FeMn/TiO2 catalysts were prepared through the wet impregnation method, and their NH3-SCR activities were tested. The results showed that Er modification could obviously promote SO2 resistance of FeMn/TiO2 catalysts at a low temperature. The promoting effect and mechanism were explored in detail using various techniques, such as BET, XRD, H2-TPR, XPS, TG, and in-situ DRIFTS. The characterization results indicated that Er modification on FeMn/TiO2 catalysts could increase the Mn4+ concentration and surface chemisorbed labile oxygen ratio, which was favorable for NO oxidation to NO2, further accelerating low-temperature SCR activity through the “fast SCR” reaction. As fast SCR reaction could accelerate the consumption of adsorbed NH3 species, it would benefit to restrain the competitive adsorption of SO2 and limit the reaction between adsorbed SO2 and NH3 species. XPS results indicated that ammonium sulfates and Mn sulfates formed were found on Er-modified FeMn/TiO2 catalyst surface seemed much less than those on FeMn/TiO2 catalyst surface, suggested that Er modification was helpful for reducing the generation or deposition of sulfate salts on the catalyst surface. According to in-situ DRIFTS the results of, the presence of SO2 in feeding gas imposed a stronger impact on the NO adsorption than NH3 adsorption on Lewis acid sites of Er-modified FeMn/TiO2 catalysts, gradually making NH3-SCR reaction to proceed in E–R mechanism rather than L–H mechanism.

Graphical Abstract

1. Introduction

Nitrogen oxides (NOx) emitted from diesel engines and industrial processes are significant atmospheric pollutants, which can lead to a number of environmental problems, such as acid rain, photochemical smog, ozone depletion, and greenhouse effects [1,2,3,4]. During the past decades, a number of NOx removal technologies have been developed with the purpose of abating NOx emissions. Among them, selective catalytic reduction (SCR) with NH3 (or urea) as a reductant is an effective and economic NOx removal technique. It has been widely used in stationary and mobile sources [5,6,7]. In SCR systems, catalysts play a critical role in creating an efficient reaction and controlling the construction cost [8]. At present, V2O5-WO3/TiO2 and V2O5-MoO3/TiO2 catalysts have been used extensively in commercial projects. However, there are still several disadvantages, such as the narrow catalytic temperature window (300–400 °C) and toxic effect of vanadium species, limiting their further applications [9,10]. Therefore, it is of great significance to develop vanadium-free SCR catalysts with excellent catalytic activities at low temperatures (<250 °C).
In recent years, more and more researchers around the world are trying to develop high-activity low-temperature SCR catalysts. Several kinds of transition metals, such as Mn, Co, Fe, Cu, and Ni, have been extensively studied and exhibited promising SCR performance [11,12,13,14,15]. It is worthy to point out that FeMn/TiO2 catalysts have been investigated extensively due to their advantages of excellent low-temperature SCR performance, environmentally friendly nature, and low cost [16,17,18]. However, there are still some challenging problems for FeMn/TiO2 catalysts to overcome, including poor N2 selectivity and weak tolerance to SO2 and H2O.
Generally, when a certain concentration of SO2 gas co-exists in flue gas, ammonium sulfates and/or metal sulfates are unavoidably formed on the SCR catalyst surface. They block and destroy active sites on SCR catalyst surface, causing the deactivation of low-temperature catalysts. To enhance SO2 resistance of FeMn/TiO2 catalysts, a promising method is to modify the catalysts with other metal elements, further inducing some synergetic effects between the dopants and Fe-Mn-Ti oxides [19,20,21]. Jiang et al. demonstrated that doping Zr in FeMn/TiO2 catalyst could enhance the formation of NO2 in the SCR reaction, thus improve the tolerance of SO2 [19]. Hou et al. found that the addition of proper amount of praseodymium (Pr) to FeMn/TiO2 catalyst could restrain the deposition of ammonium sulfates, thus improve the catalyst stability against SO2 poisoning [22]. A similar conclusion was drawn from their other work in whichSO2 resistance for NH3-SCR at low temperature was enhanced by the doping of lanthanum (La) [23]. The adsorption and oxidation of SO2 on the catalyst were inhibited by La modification.
Erbium (Er) is a kind of rare earth element. It has been applied as a doping additive in various fields due to the incompletely occupied 4f and empty 5d orbitals [24]. For example, Er-doped ZnO (EZO) have been widely studied and expected as one of the promising materials for ZnO based optoelectronic device. The incorporation of Er could affect the band gap and optical constants of ZnO thin films [25]. Er was also used to enhance the sonocatalytic performance of organic dye degradation [26]. The prepared Er3+:YAlO3/TiO2-Fe2O3 composite demonstrated a good sonocatalytic activity. Er3+ ion is an attractive optical activator and displays the richest spectra in the UV-vis-IR range due to its rich energy level structure.
The application of Er in the field of NH3-SCR catalysts is rare. However, it has been used as doping additives to enhance the NH3-SCR performance of V and Ce based catalysts. Vargas et al. found that Er modification on V2O5-WO3/TiO2 catalyst could improve the catalytic activity of catalysts after ageing [27], and the addition of Er caused modifications in symmetric deformation modes of ammonia coordinated on Lewis acid sites. Casanova et al. found that Er mixed with FeVOx supported on TiO2-WO3-SiO2 could enhance the structural and textural stability of the system by hindering the transformation of anatase to rutile, thus enhance activity after thermal treatment [28]. Kim et al. found that Er doping in CeVOx catalyst could enhance redox features and improve the quantities of acid sites and defects. This led to the high-performance of NH3-SCR reaction [29]. Jin et al. found that Er doping could increase oxygen storage capacities and oxygen vacancy concentrations of CeZr/TiO2 catalysts, resulting in excellent SCR activity and SO2 resistance [30]. It was considered that the introduction of Er into SCR catalysts was in favor of enhancing the structural and textural stability, redox property, improving the mounts of acid sites and oxygen vacancy, and blocking anatase transformation to rutile. It was possible to improve the activity and SO2 resistance of FeMn/TiO2 catalysts by doping with Er. However, to the best of our knowledge, the promotional effect of Er incorporation on SO2 resistance for NH3-SCR at low temperature over FeMn/TiO2 catalysts has not been investigated yet.
In the present study, a series of Er-modified FeMn/TiO2 catalysts with different Er/Mn molar ratios were successfully synthesized via the wet impregnating method. NH3-SCR activity and SO2 resistance over the prepared catalysts were tested. The promotional effects of Er modification on SO2 resistance were studied systematically by various characterization techniques with the fresh and used catalysts. Finally, the possible mechanisms for the excellent SO2 tolerance over Er-modified FeMn/TiO2 catalysts were discussed.

2. Results

2.1. Low Temperature SCR Activity and SO2 Resistance

Figure 1a and Figure S1a show the activities of ErxFeMn/TiO2 catalysts prepared with different Er/Mn molar ratios across a temperature window of 60–240 °C. Compared to the FeMn/TiO2 catalyst, Er0.01FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts exhibited higher NOx conversion efficiencies over the whole temperature range. But the activity at low temperature (<120 °C) decreased sharply when further increasing Er doping amount. The result indicated that, the low-temperature catalytic performance of FeMn/TiO2 catalyst could be enhanced by doping only a small amount of Er. As shown in Figure 1a, Er0.05FeMn/TiO2 catalyst possessed more than 95% NOx conversion efficiency in the temperature range of 120–240 °C. Since the NOx conversion performance of Er0.05FeMn/TiO2 catalyst was better than those of other catalysts over the whole temperature range, a proper Er doping molar ratio of 0.05 (Er/Mn) was chose for further investigation in this work.
The N2 selectivity of the prepared catalysts in the temperature range of 60–240 °C is illustrated in Figure S1b. It could be seen that the N2 selectivity of FeMn/TiO2 catalyst at 100 °C was 92.4%, but quickly decreased to 18.8% as reaction temperature increased to 240 °C. N2 selectivity of the prepared FeMn/TiO2 catalyst was not so good, and this result was similar to that of some previous work [31]. The poor N2 selectivity could be ascribed to the formation of unwanted N2O through some side reactions [32]. When the Er doping molar ratio was 0.01, the effect of Er doping on N2 selectivity in the whole temperature range was negligible. However, when further increasing Er doping molar ratio, N2 selectivity of Er-modified catalysts improved obviously. The higher the Er doping molar ratio was, the better N2 selectivity was. Here, the enhancement effect on N2 selectivity was mainly ascribed to the inhibition of N2O formation by the introduction of Er element.
In order to investigate the SO2 tolerance of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts, the SO2 resistance of these two catalysts was tested in the presence of 100 ppm SO2. As shown in Figure 1b, the NOx conversion efficiency of FeMn/TiO2 catalyst decreased quickly from 100% to 56.0% after introducing SO2 in feeding gas for 6 h, and it could be restored to 61.0% after stopping SO2 for 2 h. Compared with FeMn/TiO2 catalyst, Er0.05FeMn/TiO2 catalyst displayed much better SO2 resistance performance. After the introduction of SO2 in feeding gas, NOx conversion efficiency of Er0.05FeMn/TiO2 catalyst decreased slowly, and it could be maintained at 90.8% after exposure to 100 ppm SO2 for 6 h. After stopping the injection of SO2 for 2 h, NOx conversion efficiency of Er0.05FeMn/TiO2 catalyst could recover to 95.5%. This result indicated that the introduction of Er element could greatly improve SO2 resistance performance of FeMn/TiO2 catalysts.
Generally, a small amount of H2O vapor will exist in the flue gas of stationary and mobile sources. Here, 5% H2O was introduced in the feeding gas to test the H2O resistance of the selected catalysts. As shown in Figure 1c, the NOx conversion efficiency of FeMn/TiO2 catalyst decreased gradually from 100% to 88.0% after introducing 5% H2O for 6 h, and it could be recovered to 100% quickly by stopping H2O injection. This indicated that the deactivation of FeMn/TiO2 catalyst by H2O was reversible. By contrast, Er0.05FeMn/TiO2 catalyst exhibited excellent H2O resistance performance. The introduction of 5% H2O had little effect on the NOx conversion efficiency of Er0.05FeMn/TiO2 catalyst during the whole reaction process.
The effect of coexistence of 100 ppm SO2 and 5% H2O in feeding gas on NOx conversion performance of FeMn/TiO2 catalysts with and without Er modification was also evaluated, and the results are illustrated in Figure 1d. It was clear that coexistence of SO2 and H2O imposed a much greater impact on FeMn/TiO2 catalyst than that for Er0.05FeMn/TiO2 catalyst. NOx conversion efficiency of FeMn/TiO2 catalyst decreased quickly from 100% to 18.0% after introducing SO2 and H2O for 6 h. As to Er0.05FeMn/TiO2 catalyst, the change trend of NOx conversion efficiency was similar to the case exposed to single SO2. NOx conversion efficiency of Er0.05FeMn/TiO2 catalyst decreased slowly from 100% to 84.7% after injection of SO2 and H2O for 6 h. Therefore, a conclusion could be drawn that the resistance to SO2 and H2O of FeMn/TiO2 catalyst could be enhanced obviously by Er modification.
Previous studies reported that there were two main reasons for the deactivation of catalysts suffered from SO2 poisoning [33,34,35]. On one hand, NH3 used as NO reductant would react with SO2 at low temperature, and the generated ammonium sulfates would cover on the catalyst surface which would block the active sites. One the other hand, the active species of the catalysts would also react with SO2 to form inactive metal sulfates, which were destructive to the NH3-SCR reaction. According to the SO2 resistance tests in this work, the introduction of Er element could greatly improve SO2 resistance performance of FeMn/TiO2 catalysts at 200 °C. It is known that ammonium sulfates can be decomposed at 230–400 °C [36], while metal sulfates can be decomposed at a much higher temperature [37]. In view of this, it was possible that Er modification restrained the formation of ammonium sulfates and inactive metal sulfates to some extent, thus preventing the active sites being blocked or deactivated. The mechanisms of the enhancement effect of Er modification on SO2 resistance were further analyzed through the following characterizations.

2.2. BET Results

N2 adsorption-desorption isotherms of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts are shown in Figure 2. According to IUPAC classification, the adsorption isotherms of all the catalysts were type IV, suggesting that the mesoporous structures of the prepared catalysts were not changed after SO2 resistance tests [38]. The specific surface areas, pore volumes, and pore sizes of these catalysts were shown in Table S1. The specific surface areas of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts were 58.6 and 59.5 m2/g, respectively. After SO2 resistance tests, the specific surface area of FeMn/TiO2-S catalyst decreased by 10.9%, while that of Er0.05FeMn/TiO2-S catalyst decreased by 3.5% only. It implied that a certain amount of ammonium sulfates had been formed on the surface of both kinds of catalysts due to the reaction between NH3 species and SO2 [33]. As the decrease in specific surface area of Er0.05FeMn/TiO2 catalyst was much less than that of FeMn/TiO2 catalyst, it demonstrated that Er modification could effectively restrain the formation of ammonium sulfates on the surface of FeMn/TiO2 catalyst.

2.3. XRD Analysis

Figure 3 presents the XRD patterns of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts. The diffraction peaks of 2θ at 25.3°, 36.9°, 38.6°, 48.0°, 53.9°, 55.1°, and 62.7° could be assigned to anatase TiO2 characteristic peaks (ICCD PDF-2, 21-1272), and the diffraction peaks of 2θ at 33.0°, 38.2°, and 55.1°could be assigned to Mn2O3 characteristic peaks (ICCD PDF-2, 65-7467). It could be seen that all catalysts exhibited distinct anatase TiO2 diffraction peaks, and no obvious diffraction peaks of rutile TiO2 were found. The diffraction peaks of MnOx in the prepared FeMn/TiO2 catalyst matched well with the standard powder diffraction pattern of Mn2O3, implying Mn2O3 was the main crystal phase. The diffraction peaks of Mn2O3 in Er-modified FeMn/TiO2 catalysts were not obvious. This suggests that Er modification promoted the transformation of MnOx from crystalline state to highly dispersed amorphous state, which was beneficial to enhance the catalytic performance at low temperature [39,40]. After the SO2 resistance test, there were more peaks appeared in the diffraction pattern of the used FeMn/TiO2 catalyst, which could be ascribed to the formation of Ti(SO4)2 on catalyst surface (ICCD PDF-2, 18-1406) [41]. However, the same diffraction peak of Ti(SO4)2 was not observed in the diffraction pattern of Er0.05FeMn/TiO2-S catalyst. Since sulfates would cover active sites on the surface of FeMn/TiO2 catalysts, it might be the main reason for the FeMn/TiO2 catalyst’s low SCR activity. This result also agreed well with the aforementioned SO2 resistance test results. Furthermore, it could also demonstrate that Er doping in FeMn/TiO2 catalysts was in favor of inhibiting the formation of sulfates on the catalyst surface, thus improving SO2 tolerance of Er-modified FeMn/TiO2 catalysts.

2.4. H2-TPR Analysis

The redox properties of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts were investigated via H2-TPR, and the results are shown in Figure 4. The reduction peak temperatures and H2 consumption values were listed in Table S2. As for the FeMn/TiO2 catalyst, there were three reduction peaks: the low temperature reduction peak centered at 313 °C could be assigned to the reduction of MnO2 to Mn2O3, the peak centered at 389 °C belonged to the reduction of Mn2O3 to MnO or Fe2O3 to Fe, and the peak centered at 582 °C belonged to the reduction of Mn3O4 to MnO [18,42]. Compared to ErFeMn/TiO2 catalyst, all the reduction peaks of Er0.05FeMn/TiO2 catalyst were shifted toward low temperature direction, indicating that Er0.05FeMn/TiO2 catalyst had a better redox ability at low temperature. As shown in Table S2, the total H2 consumption value of Er0.05FeMn/TiO2 catalyst was much more than that of FeMn/TiO2 catalyst, indicating the number of reductive species on the surface of Er0.05FeMn/TiO2 catalyst was much more than that of FeMn/TiO2 catalyst. The low reduction temperature and high H2 consumption value demonstrated that Er modification could improve the low-temperature redox ability of FeMn/TiO2 catalysts [33]. This result also agreed well with the improved SCR activity of Er-modified FeMn/TiO2 catalysts at a low temperature.
After SO2 resistance tests, the reduction peaks of the used catalysts shifted toward a higher temperature zone. As shown in Figure 4b, there were three reduction peaks in the profiles of these two kinds of catalysts. Their strongest peaks were centered at 457 ºC and 460 °C, respectively, which were related to the reduction of sulfate species (Mn sulfates) [33]. The corresponding H2 consumption value (151.6 cm3/g STP) for the strongest reduction peak of Er0.05FeMn/TiO2-S catalyst was less than that (194.8 cm3/g STP) of FeMn/TiO2-S catalyst. It suggested that less amount of sulfate species had been formed on the surface of Er0.05FeMn/TiO2 catalyst. The peaks centered at 379 and 360 °C could be assigned to the reduction of MnO2 and Mn2O3 [38], and the corresponding H2 consumption value (39.3 cm3/g STP) for Er0.05FeMn/TiO2-S catalyst was much more than that (24.1 cm3/g STP) of FeMn/TiO2-S catalyst. It indicated that Er modification was favorable to maintain the redox ability of FeMn/TiO2 catalyst when SO2 coexisted in feeding gas. It was possible that Er modification on FeMn/TiO2 catalyst could alleviate the deactivation due to the reaction between active species and SO2, thus inhibiting the formation of Mn sulfate species on the catalyst surface.

2.5. XPS Analysis

The surface compositions and elementary oxidation states of the fresh and used catalysts (FeMn/TiO2 and Er0.05FeMn/TiO2) were analyzed by X-ray photoelectron spectrometer (XPS). The spectra of Mn 2p, O 1s and S 2p are shown in Figure 5, and the analysis results of surface atomic concentration ratios are presented in Table S3. As shown in Figure 5a, two main peaks assigned to Mn 2p3/2 and Mn 2p1/2 were observed in the spectra of the catalysts. The overlapping peaks of Mn 2p3/2 could be divided into several sub-peaks with Shirley-type background. The obtained peaks could be ascribed to Mn2+ (640.9–641.7 eV), Mn3+ (642.0–643.0 eV) and Mn4+ (643.3–644.5 eV), respectively [37]. The atomic ratio of Mn4+ to Mnn+ was decided according to the peak area. As shown in Table S3, the atomic ratios of Mn4+/Mnn+ for FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts were 31.9% and 33.9%, respectively. It suggested that more surface Mn4+ species were generated with the introduction of Er. This result was in accordance with the result of H2-TPR, in which the area of the reduction peak ascribed to MnO2 to Mn2O3 in the profile of Er0.05FeMn/TiO2 catalyst was larger than that of FeMn/TiO2 catalyst. Kapteijn et al. found that NO removal efficiency at low temperature for manganese oxides decreased in an order of MnO2 > Mn5O8 > Mn2O3 > Mn3O4 [43]. The high catalytic activity of MnO2 (Mn4+) was ascribed to its lattice defects [44]. Here, it could be inferred that the formation of more surface Mn4+ species was conducive for achieving a high NOx conversion efficiency over Er0.05FeMn/TiO2 catalyst, as shown in Figure 1a. Compared with the fresh FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts, the peaks of Mn 2p3/2 for the used catalysts were shifted toward higher binding energy (from 641.8 and 642.1 to 642.3 and 642.5 eV, respectively). It indicated that Mn atoms in the used catalysts were not only bonded to O atoms, but also to the atoms with lower electronegativity [38]. According to S 2p spectra in the XPS and the analysis results of H2-TPR, it was speculated the atoms with lower electronegativity might be originated from Mn sulfates [37,38]. The shift of binding energy of the used FeMn/TiO2 catalyst before and after SO2 resistance test was more obvious than that for Er0.05FeMn/TiO2 catalyst. It implied that Er modification inhibited the formation of sulfates to a certain extent. In addition, Mn atomic concentration on the surface of FeMn/TiO2 catalyst decreased by 4.4% after SO2 resistant test, while that of Er0.05FeMn/TiO2 catalyst decreased by only 0.5%. It demonstrated that Er modification could also inhibit the loss of Mn active species on the surface of catalysts, and thus improving the SO2 resistance.
XPS spectra of O 1s over the fresh and used catalysts were divided into two sub-peaks, and the results are presented in Figure 5b. The sub-peaks centered at 529.7–530.3 eV could be assigned to lattice oxygen O2− (denoted as Oβ). The sub-peaks centered at 530.9–531.9 eV could be assigned to surface chemisorbed labile oxygen (denoted as Oα), including defect oxide (O22–) and hydroxyl-like group (O) [45,46]. The surface chemisorbed labile oxygen (Oα) was not only beneficial for oxidizing NO into NO2, but also contributive to the generation of more active intermediate species such as –NH2 and adsorbed NO2 during the NH3-SCR process [31,47]. Thus, the surface chemisorbed labile oxygen could exhibit better catalytic activity than lattice oxygen. The ratio of Oα to (Oβ + Oα) was calculated according to the relative peak areas. As shown in Table S3, Oα/(Oβ + Oα) ratio of Er0.05FeMn/TiO2 catalyst was 45.7% which was much higher than that of FeMn/TiO2 catalyst (20.9%). It indicated that Er modification was favorable for forming more surface chemisorbed labile oxygen species on the catalyst surface. This result was in accordance with that of catalytic activity tests in which Er0.05FeMn/TiO2 catalyst exhibited higher NOx conversion efficiency than that of FeMn/TiO2 catalyst. Here the increase of chemisorbed labile oxygen ratio was ascribed to the interaction between Er and Ti, which could enhance the formation of oxygen vacancies on the catalyst surface [30]. After SO2 resistance tests, both binding energy peaks of surface chemisorbed labile oxygen (Oα) and lattice oxygen (Oβ) were shifted toward higher values compared to those of the fresh catalysts. It implied that O and O2 might be bonded with other substances when the catalysts were poisoned by SO2. Despite that, the surface chemisorbed labile oxygen (Oα) concentration of Er0.05FeMn/TiO2-S catalyst was still much higher than that of the FeMn/TiO2-S catalyst.
The XPS spectra of S 2p for the used catalysts are shown in Figure 5c. The two sub-peaks could be ascribed to S6+ in the form of SO42– (168.6–168.9 eV) and S4+ in the form of SO32– (169.7–170.1 eV), respectively [33,48,49]. It demonstrated that sulfate and sulfite salts had been formed on the catalyst surface after SO2 resistance tests, possibly as ammonium or Mn salts. This result could not only explain the shift of binding energy found in XPS, but also confirm Ti(SO4)2 diffraction peak detected in XRD. In addition, as shown in Table S3, the contents of S and N on the surface of Er0.05FeMn/TiO2 catalyst were 2.8% and 1.6%, respectively, which were much less than those of FeMn/TiO2 catalyst (5.3% and 2.2%), suggesting that Er modification could inhibit the formation of sulfate and/or sulfite salts on the catalyst surface when SO2 coexisted in feeding gas.

2.6. TG Analysis

In order to further investigate the surface compounds of the catalysts poisoned by SO2, thermogravimetric analysis of FeMn/TiO2-S and Er0.05FeMn/TiO2-S catalysts were performed, and the TG-DTG curves are presented in Figure 6. Four temperature steps could be found in the weight loss curves of the catalysts. In the temperature range from room temperature to 200 °C (Step 1), the weight losses were assigned to the evaporation of physically absorbed H2O on the catalysts [50]. As (NH4)2SO4 and NH4HSO4 would be decomposed completely when heating temperatures were above 230 and 350 °C, respectively [51], the weight losses in the temperature range of 200 to 400 °C (Step 2) was mainly ascribed to the decomposition of (NH4)2SO4 and NH4HSO4. In the temperature range of 500–650 °C (Step 3), the weight losses were related to the phase transformation of metal oxides [52]. The weight losses in the temperature range of 650–900 °C (Step 4) were mainly attributed to the decomposition of MnSO4 [37]. The weight loss in the second step for Er0.05FeMn/TiO2-S catalyst was 0.281% which was obviously less than that for FeMn/TiO2-S catalyst (0.671%). The intensities of DTG curves (Step 2) for Er0.05FeMn/TiO2-S catalyst were lower than those of FeMn/TiO2-S catalyst. It indicated that less (NH4)2SO4 and NH4HSO4 were formed on Er-modified catalyst compared with FeMn/TiO2 catalyst. These results agreed well with the XPS spectra of S 2p. The TG-DTG results further proved that less sulfates formed on the surface of Er-modified FeMn/TiO2 catalysts during SCR reaction in the presence of SO2.

2.7. In-Situ DRIFTS

Aiming to get further insight into the adsorption behavior and the reaction mechanism during NH3-SCR process in the presence of SO2 at low temperature, in-situ DRIFTS experiments were performed at relatively low temperature (200 °C).

2.7.1. Adsorption of SO2

Figure 7 shows in-situ DRIFTS spectra of SO2 adsorption over FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts. After pretreatment of each catalyst sample, 500 ppm SO2 (50 mL/min) was introduced for 30 min, and then IR spectra were recorded. Several bands were observed in the ranges of 800–2000 cm−1. The bands at 1060, 1162, 1284 cm−1 were ascribed to the mononuclear bidentate sulfate complex [38]. The band around 1613 cm−1 was attributed to the sulfates formed by the weak adsorption of SO2, while the band at 1344 cm−1 was ascribed to the asymmetric stretching vibration of O=S=O in sulfates [38]. It could be seen that the intensities of all the bands over Er0.05FeMn/TiO2 catalyst were similar with those over FeMn/TiO2 catalyst, implying that Er modification had no obvious effect on the SO2 adsorption over the FeMn/TiO2 catalyst.

2.7.2. Effect of SO2 on NH3 Adsorption on Catalyst Surface

In order to investigate the effect of SO2 on NH3 adsorption on the catalysts, in-situ DRIFTS spectra of the co-adsorption of NH3 + SO2 at 200 °C were analyzed. After pretreatment, the catalysts were pre-adsorbed with 500 ppm NH3 (50 mL/min) for 30 min, then 100 ppm SO2 was introduced, and the IR spectra were measured as a function of time.
As shown in Figure 8a, several bands were found in the spectra of FeMn/TiO2 catalyst after the adsorption of NH3 for 30 min. The strong absorbed bands (1166 and 1605 cm−1) were assigned to coordinate NH3 on Lewis acid sites, and the weak band (1450 cm−1) was assigned to NH4+ species on Brønsted acid sites [37,45,53]. After the introduction of SO2, the absorbance intensities of all the adsorbed bands increased obviously. With the introduction of SO2, the absorbance intensity of the band (1166 cm−1) assigned to NH4+ species on Brønsted acid sites was enhanced gradually, and a new band (1678 cm−1) assigned to NH4+ species on Brønsted acid sites appeared in the spectra. Zhang et al. also reported that SO2 could promote the generation of NH4+ species on Brønsted acid sites [38]. However, the band (1166 cm−1) corresponding to NH3 species on Lewis acid sites decreased gradually, and almost disappeared after 5 min of SO2 injection. At the same time, three bands appeared at 1261, 1144 and 1033 cm−1 which were ascribed to sulfates, and they were also enhanced with the increase of SO2 injection duration [38]. It indicated that the adsorption of NH3 on Lewis acid sites was inhibited by the introduction of SO2, thus reducing the SCR activity at low temperature. As shown in Figure 8b, the change in absorbed bands of Er0.05FeMn/TiO2 catalyst was similar to that of FeMn/TiO2 catalyst. However, the intensity of the band (1169 cm−1) corresponding to NH3 species on Lewis acid sites increased gradually with the introduction of SO2 for 30 min. According to the previous study, adsorbed NH3 species on Lewis acid sites played a key role in SCR activity at low temperature [54]. Here it was concluded that the inhibition of competitive adsorption between SO2 and NH3 species on Lewis acid sites was a contributive factor to high SO2 tolerance of Er-modified FeMn/TiO2 catalysts.
Figure 8c exhibits the in-situ DRIFTS spectra of NH3 adsorption on FeMn/TiO2-S and Er0.05FeMn/TiO2-S catalysts at 200 °C. After adsorption of 500 ppm NH3 for 30 min and N2 purging, several bands ascribed to various NH3 species were found in the spectra of the fresh catalysts. The strong absorbed bands (1166 and 1605 cm−1) were assigned to asymmetric and symmetric bending vibrations of the N–H bonds in NH3 coordinated to Lewis acid sites, and the weak band (1450 cm−1) was assigned to the asymmetric bending vibrations of ionic NH4+ bonded to the Brønsted acid sites [37,45,53,55]. The intensities of all of bands over Er0.05FeMn/TiO2 catalyst were higher than those over the FeMn/TiO2 catalyst. It demonstrated that there were more Lewis acid sites and Brønsted acid sites existed on the surface of Er0.05FeMn/TiO2 catalyst. As to the used catalysts, the intensity of the bands (1609 and 1084 cm−1) ascribed to NH3 species adsorbed on Lewis acid sites decreased obviously due to the formation of new bands (1260 and 1032 cm−1) ascribed to sulfate species [37,38]. However, the intensities of the bands (1684 and 1443 cm−1) ascribed to NH4+ species adsorbed on Brønsted acid sites increased remarkably due to the formation of sulfate species during the SO2 resistance tests. Some previous studies reported that the formation of sulfates on the catalyst surface could result in an increase in the amount of NH4+ species adsorbed on Brønsted acid sites and a decrease in the amount of NH3 species adsorbed on Lewis acid sites [17,55,56]. Here the decrement of adsorbed NH3 species on Lewis acid sites and the increment of adsorbed NH4+ species on Brønsted acid sites over Er0.05FeMn/TiO2-S catalyst were much less compare to FeMn/TiO2-S catalyst. It demonstrated that Er modification improved the NH3 adsorption on Lewis acid sites of Er0.05FeMn/TiO2-S catalyst, and less sulfate species formed on the surface of Er0.05FeMn/TiO2-S catalyst. As a result, Er modification on FeMn/TiO2 catalyst improved the SO2 tolerance during the NH3-SCR process at a low temperature.

2.7.3. Effect of SO2 on NO + O2 Adsorption on Catalyst Surface

To understand the effect of SO2 on NOx adsorption on FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts, each catalyst was exposed to 500 ppm NO and 5% O2 at 200 ºC for 30 min, then 100 ppm SO2 was introduced. As shown in Figure 8a, with the introduction of NO + O2 for 30 min, several bands of adsorbed NOx species appeared at 1575, 1380, 1366, and 1262 cm−1 in the spectra of FeMn/TiO2 catalyst, which were ascribed to bidentate nitrate (1575 cm−1) [53], M-NO2 nitro compounds (1380 and 1366 cm−1) [53], and monodentate nitrate (1262 cm−1) [53], respectively. With the introduction of SO2, all bands in the spectra of FeMn/TiO2 catalyst exhibited an obvious decrease in intensity. Moreover, several new bands (1602, 1270, 1159 and 1069 cm−1) ascribed to sulfates appeared with the injection of SO2, and the absorbance intensities of these bands increased gradually. It indicated that SO2 could restrain the adsorption of NOx species on FeMn/TiO2 catalysts. In other words, there was obvious competitive adsorption between NO and SO2 which would inhibit the SCR reaction under L-H mechanism. It might be the reason for the low NOx conversion efficiency in the presence of SO2 for FeMn/TiO2 catalyst [57]. The inhibiting effect of SO2 on NOx adsorption on Er0.05FeMn/TiO2 catalyst surface was quite similar to that for FeMn/TiO2 catalyst. But the absorbance intensity of the band (1601 cm−1) assigned to sulfates was obviously lower than that of FeMn/TiO2 catalyst, indicating that less sulfate species formed on Er0.05FeMn/TiO2 catalyst.
Figure 9c exhibits the in-situ DRIFTS spectra of 500 ppm NO and 5% O2 adsorption on FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts at 200 °C. After adsorption of NO and O2 for 30 min and purging with N2, several bands were found in the spectra of the fresh catalysts, which could be ascribed to bidentate nitrate (1575 cm−1) [53], M-NO2 nitro compounds (1380 and 1366 cm−1) [53], and monodentate nitrate (1262 cm−1) [53], respectively. After SO2 resistance tests, the intensities of the bands ascribed to NOx species in the spectra of both used catalysts decreased dramatically. It implied that the formation of nitrate species could be inhibited by SO2 adsorption. In addition, the negative bands (1326 and 1311 cm−1) ascribed to the displacement of the sulfates were found on FeMn/TiO2-S and Er0.05FeMn/TiO2-S catalysts. This indicated that competitive adsorption between NO and SO2 occurred on the surface of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts [58].
According to the in-situ DRIFTS results mentioned above, it can be confirmed that the adsorption of NO on FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts was deeply inhibited by SO2, while the inhibition effect of SO2 on NH3 adsorption over the Er0.05FeMn/TiO2 catalyst was much weaker than that over the FeMn/TiO2 catalyst. As a result, the reaction through the Langmuir–Hinshelwood (L–H) mechanism for these two catalysts was suppressed severely. While the inhibiting effect on the reaction through Eley–Ridel (E–R) mechanisms for Er0.05FeMn/TiO2 catalyst was obviously weaker than that for FeMn/TiO2 catalyst. Therefore, Er modification promoted SO2 resistance of the FeMn/TiO2 catalyst during the NH3-SCR process at low temperature.

3. Discussion

In our present work, compared to the FeMn/TiO2 catalyst, Er0.05FeMn/TiO2 exhibited better SCR activity and SO2 tolerance at a low temperature. According to the analysis results of H2-TPR, XPS and in-situ DRIFTS, the increased amounts of Mn4+ (Equations (5) and (6)), surface chemisorbed labile oxygen (Equation (4)) and Lewis acid sites (Equation (1)) were considered as the dominant factors which were beneficial to achieve high SCR activity at low temperature for Er0.05FeMn/TiO2 catalyst. When SO2 was introduced in feeding gas, the adsorbed NH3 species and Mn4+ active species could react with the adsorbed SO2 species (Equations (10) and (11)). The generated (NH4)2SO4 and MnSO4 could not only block the active sites, but also suppress the redox cycles and the relevant reactions [59,60]. Based on the analysis results of BET, XRD, H2-TPR, XPS, TG-DTG and in-situ DRIFTS, the amounts of ammonium sulfates and metal sulfates generated on Er0.05FeMn/TiO2 catalyst were obviously less than those for FeMn/TiO2 catalyst. Thus, the SO2 tolerance of FeMn/TiO2 catalyst was remarkably enhanced by Er modification.
NH 3   ( g ) NH 3   ( ad ) ( Lewis   acid   sites )
NO   ( g ) NO   ( ad )
SO 2   ( g ) SO 2   ( ad )
O 2   ( g ) 2 O ( ad )
NH 3   ( ad ) + Mn 4 + + O ( ad ) NH 2   ( ad ) + Mn 3 + + OH
NO   ( ad ) + Mn 4 + + O ( ad ) Mn 3 + + NO 2 ( ad )
NO   ( g )   +   NH 2 ( ad ) NH 2   NO ( ad )
NH 2 NO ( ad ) N 2 ( g ) + H 2 O ( ad )
2 NH 3 ( ad ) + NO 2   ( ad ) + NO ( g ) 2 N 2 + 3 H 2 O
4 NH 3 ( ad ) + 2 SO 2 ( ad ) + O 2 + 2 H 2 O 2 ( NH 4 ) 2 SO 4
MnO 2 + SO 2 ( ad ) MnSO 4
2 NH 3 + NO + NO 2 2 N 2 + 3 H 2 O
According to the result of in-situ DRIFTS, competitive adsorptions of SO2 with NO and NH3 would occur after the introduction of SO2. Therefore, the adsorption and activation of NO and NH3 on FeMn/TiO2 catalyst would be inhibited greatly (Equations (1) and (2)), further suppressing the subsequent SCR reactions. However, coexisting SO2 only imposed a slight impact on NH3 adsorption on Lewis acid sites of Er0.05FeMn/TiO2 catalyst, thus NH3-SCR reaction through E-R mechanism could proceed in a normal way. It played a key role for achieving excellent SO2 tolerance over the Er0.05FeMn/TiO2 catalyst. In addition, XPS analysis results indicated that the amounts of Mn4+ and surface chemisorbed labile oxygen on the surface of Er0.05FeMn/TiO2 catalyst were much more than those for FeMn/TiO2 catalyst. It was reported that abundant Mn4+ and surface chemisorbed labile oxygen species were beneficial for fast SCR reaction (Equation (12)) [61,62], which would accelerate the consumption of adsorbed NH3 species on Er0.05FeMn/TiO2 catalyst. Therefore, the competitive adsorption of SO2 and NH3 species could be inhibited by Er modification on FeMn/TiO2 catalyst. The reaction between adsorbed SO2 and NH3 species was limited, resulting in less ammonium sulfates formed on the surface of Er-modified FeMn/TiO2 catalyst. According to the results and discussions mentioned above, the promotional effect and mechanism of Er modification on SO2 resistance for NH3-SCR at low temperature over FeMn/TiO2 catalysts are illustrated in Figure 10.

4. Materials and Methods

4.1. Catalysts Preparation

A facile wet impregnating method was used to prepare Er-modified FeMn/TiO2 catalysts. The molar ratio of Fe/Mn/Ti was fixed at 2:6:15 while Er/Mn molar ratios varied in the range of 0–0.2. In a typical preparation process, 1.686 g Fe(NO3)3·9H2O, 3.143 Mn(NO3)2·4H2O and a certain amount of Er(NO3)2·6H2O were dissolved in 100 mL deionized water, successively. 2.5 g TiO2 powder (Degussa P25) was added into the mixture solution, followed by stirring thoroughly at 80 °C for 2 h. Then, the ammonia solution (28 wt.%) was added dropwise until solution pH reached at 10. Subsequently, the precipitate was collected by vacuum filtration, dried at 120 °C overnight, and calcined at 500 °C for 3 h in air. All the catalysts were crushed and sieved to 40–60 mesh for activity evaluation. Here Er-modified FeMn/TiO2 catalysts were denoted as ErxFeMn/TiO2, where x represented various Er/Mn molar ratios (0.01, 0.05, 0.1 and 0.2). For comparison, FeMn/TiO2 catalyst without Er modification was also prepared by using the same method without the addition of Er salt. The FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts after SO2 resistance experiments were denoted as FeMn/TiO2-S and Er0.05FeMn/TiO2-S, respectively.

4.2. Catalyst Activity Test

SCR activity measurements were performed using 1 mL (~0.5 g) of catalyst sample in a fixed-bed quartz reactor (i.d. 6 mm). Before each test, the catalyst was pretreated at 240 °C for 1 h in N2 stream, and then cooled down to room temperature. Then, the catalyst was exposed to a reactant gas mixture containing 500 ppm NO, 500 ppm NH3, 5% O2, 5% H2O (if used), 100 ppm SO2 (if used), and N2 as balance gas. The total flow rate of simulated flue gas was 500 mL/min corresponding to a gas hourly space velocity (GHSV) of 30,000 h−1. After purging for 1 h, reactant gases adsorbed on catalyst surface reached an adsorption-desorption equilibrium. The reaction temperature increased from room temperature to 240 °C. The concentrations of NO, NO2, N2O, SO2, NH3, and H2O were monitored continuously by a FTIR spectrometer (Antaris IGS, Thermo Fisher Scientific, (Waltham, MA, USA) equipped with a heated low-volume multiple-path gas cell (2 m) and a MCT detector cooled by liquid nitrogen. NOx conversion and N2 selectivity were calculated as follows:
NO x   conversion   ( % ) = [ NO x ] in [ NO x ] out [ NO x ] in × 100 %
N 2   selevtivity   ( % ) = ( 1 [ NO 2 ] out + 2 [ N 2 O ] out [ NO x ] in [ NO x ] out + [ NH 3 ] in [ NH 3 ] out )   × 100 %

4.3. Catalyst Characterization

The textural properties of the prepared catalysts were measured by a physisorption analyzer (ASAP 2020 PLUS, Micromeritics, Norcross, GA, USA). Before each test, the catalyst was degassed under vacuum at 350 °C for 4 h, and the N2 adsorption-desorption data was recorded at liquid nitrogen temperature (−196 °C). The specific surface areas were calculated via the Brunauer–Emmett–Teller (BET) equation. The pore sizes and pore volumes were calculated via the Barrett–Joyner–Halenda (BJH) method.
Powder XRD patterns were carried out using an X-ray diffraction meter (Empyrean, PANalytical, Eindhoven, Noord-Brabant, The Netherlands) using Cu Kα as radiation source (λ = 0.15406 nm) at 40 kV and 30 mA. The diffractogram was recorded in a 2θ range of 10–80° with a scanning step size of 0.02°.
H2-temperature programmed reduction (H2-TPR) was tested using a chemisorption analyzer (Autochem II 2920, Micromeritics, Norcross, GA, USA). Prior to each test, the sample was pretreated at 350 °C for 1 h in He stream (50 mL/min), and then cooled down to 50 °C. H2 reduction data were recorded by a TCD detector from 50 to 800 °C at a heating rate of 10 °C/min. A mixture gas of 10% H2/Ar was used as reduction gas with a flow rate of 50 mL/min, and H2O produced in the reduction process was removed by a cold trap.
TG analysis was performed on a thermo gravimetric analyzer (TGA-50, Shimadzu, Tokyo, Honshu, Japan). The weight loss curves of the samples were collected over a temperature range of 0–1000 °C in N2 atmosphere with a programmed heating rate of 10 °C/min.
X-ray photoelectron spectroscopy (XPS) profiles were collected using a surface analysis photoelectron spectrometer (AXIS-ULTRA DLD-600W, Shimadzu, Tokyo, Honshu, Japan), with Al Kα as radiation source at 300 W. All binding energies were standardized to the adventitious C 1s peak at 284.8 eV. The pass energies of survey spectrum and high-resolution spectrum were 100 and 30 eV, respectively. Line shapes with sum of Gaussian and Lorentzian functions and Shirley-type background were applied for fitting XPS spectra.
In-situ DRIFTS measurements were performed on an FTIR spectrometer (Nicolet iS50, Thermo Fisher Scientific, Waltham, MA, USA) equipped with a MCT/A detector cooled by liquid nitrogen and a high-temperature reaction chamber (Praying Mantis, Harrick, Waltham, MA, USA). Prior to each test, the catalyst was pretreated in a N2 flow at 200 °C for 30 min. Then, the background spectra recorded in N2 flow were automatically subtracted from the corresponding spectra. All DRIFTS spectra were collected by accumulating 64 scans with a resolution of 4 cm−1 as a function of time.

5. Conclusions

In this work, the effect and mechanism of Er modification on SO2 resistance over FeMn/TiO2 catalysts were investigated systematically. The results showed that Er modification on FeMn/TiO2 catalysts could effectively improve SO2 tolerance at low temperature. After exposure to 100 ppm SO2 for 6 h, Er0.05FeMn/TiO2 catalyst still exhibited more than 90% NOx conversion efficiency at 200 °C. BET results indicated that Er doping could alleviate the decrease of specific surface area of FeMn/TiO2 catalyst in the presence of SO2. Less sulfates were confirmed on Er0.05FeMn/TiO2-S catalyst compared with those for FeMn/TiO2-S catalyst. The XPS results indicated that Er modification could result in an increase in the concentrations of surface Mn4+ species and chemisorbed liable oxygen on both fresh and used catalyst. In-situ DRIFTS spectra revealed that Er modification could effectively alleviate the inhibiting effect of coexisting SO2 on NH3 adsorption. NH3-SCR reaction for Er-modified FeMn/TiO2 catalysts could proceed through E-R mechanism in almost a normal way rather than L-H mechanism. As a result, superior NOx conversion performance could be achieved on Er-modified FeMn/TiO2 catalyst, even in the presence of SO2.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal11050618/s1, Figure S1: (a) NOx conversion, (b) N2 selectivity in the NH3-SCR reaction (Reaction conditions: 1 mL catalyst, [NO] = [NH3] = 500 ppm, [O2] = 5 vol.%, [H2O] = 5 vol.% (when used), [SO2] = 100 ppm (when used), balanced with N2, GSHV = 30,000 h−1). Figure S2: EDS mapping of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts. Figure S3: XPS survey spectrum of Er0.05FeMn/TiO2 catalyst. Table S1: Textural properties of the fresh and used catalysts. Table S2: H2-TPR results of the fresh and used catalysts. Table S3: Calculation results of surface atomic concentration ratios of Fe, Mn and O in the fresh and used catalysts. Table S4: Surface element contents of FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.

Author Contributions

Conceptualization, Z.H.; validation, Z.H.; supervision, Z.H.; investigation, Z.H. and H.D.; formal analysis, X.W., C.L. and Y.G.; methodology, J.D. and L.S.; data curation, S.Y.; funding acquisition, Z.H., J.D., L.S., S.Y. and X.P.; writing—original draft, H.D.; writing—review & editing, Z.H.; project administration, Z.H. and X.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (51779024, 51809100, 52071046, 51979022, 51979045), Natural Science Foundation of Liaoning Province of China (2020MS130), National Key Research and Development Program of China (2017YFC1404600), and Fundamental Research Funds for the Central Universities (3132019330).

Data Availability Statement

The study did not report any extra data.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Su, Z.H.; Ren, S.; Chen, Z.C.; Yang, J.; Zhou, Y.H.; Jiang, L.J.; Yang, C. Deactivation effect of CaO on Mn-Ce/AC catalyst for SCR of NO with NH3 at low temperature. Catalysts 2020, 10, 873. [Google Scholar] [CrossRef]
  2. Zhu, N.; Shan, W.P.; Lian, Z.H.; Zhang, Y.; Liu, K.; He, H. A superior Fe-V-Ti catalyst with high activity and SO2 resistance for the selective catalytic reduction of NOx with NH3. J. Hazard. Mater. 2020, 382, 120970. [Google Scholar] [CrossRef]
  3. Zhu, J.; Liu, Z.D.; Xu, L.; Ohnishi, T.; Yanaba, Y.; Ogura, M.; Wakihara, T.; Okubo, T. Understanding the high hydrothermal stability and NH3-SCR activity of the fast-synthesized ERI zeolite. J. Catal. 2020, 391, 346–356. [Google Scholar] [CrossRef]
  4. Zhou, G.Y.; Maitarad, P.; Wang, P.L.; Han, L.P.; Yan, T.T.; Li, H.R.; Zhang, J.P.; Shi, L.Y.; Zhang, D.S. Alkali-resistant NOx reduction over SCR catalysts via boosting NH3 adsorption rates by in situ constructing the sacrificed sites. Environ. Sci. Technol. 2020, 54, 13314–13321. [Google Scholar] [CrossRef]
  5. Ye, D.; Wang, X.X.; Liu, H.; Wang, H.N. Insights into the effects of sulfate species on CuO/TiO2 catalysts for NH3-SCR reactions. Mol. Catal. 2020, 496, 111191. [Google Scholar] [CrossRef]
  6. Zhao, X.Y.; Ma, S.B.; Li, Z.B.; Yuan, F.L.; Niu, X.Y.; Zhu, Y.J. Synthesis of CenTiOx flakes with hierarchical structure and its enhanced activity for selective catalytic reduction of NOx with NH3. Chem. Eng. J. 2020, 392, 123801. [Google Scholar] [CrossRef]
  7. 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 Environ. 2020, 270, 118860. [Google Scholar] [CrossRef]
  8. Fan, Y.M.; Ling, W.; Huang, B.C.; Dong, L.F.; Yu, C.L.; Xi, H.X. The synergistic effects of cerium presence in the framework and the surface resistance to SO2 and H2O in NH3-SCR. J. Ind. Eng. Chem. 2017, 56, 108–119. [Google Scholar] [CrossRef]
  9. Ye, L.M.; Lu, P.; Chen, X.B.; Fang, P.; Peng, Y.; Li, J.H.; Huang, H.B. The deactivation mechanism of toluene on MnOx-CeO2 SCR catalyst. Appl. Catal. B Environ. 2020, 277, 119257. [Google Scholar] [CrossRef]
  10. Yan, T.; Liu, Q.; Wang, S.H.; Xu, G.; Wu, M.H.; Chen, J.J.; Li, J.H. Promoter rather than inhibitor: Phosphorus incorporation accelerates the activity of V2O5-WO3/TiO2 catalyst for selective catalytic reduction of NOx by NH3. ACS Catal. 2020, 10, 2747–2753. [Google Scholar] [CrossRef]
  11. Yan, R.; Lin, S.X.; Li, Y.L.; Liu, W.M.; Mi, Y.Y.; Tang, C.J.; Wang, L.; Wu, P.; Peng, H.G. Novel shielding and synergy effects of Mn-Ce oxides confined in mesoporous zeolite for low temperature selective catalytic reduction of NOx with enhanced SO2/H2O tolerance. J. Hazard. Mater. 2020, 396, 122592. [Google Scholar] [CrossRef]
  12. Wang, Z.Y.; Guo, R.T.; Shi, X.; Liu, X.Y.; Qin, H.; Liu, Y.Z.; Duan, C.P.; Guo, D.Y.; Pan, W.G. The superior performance of CoMnOx catalyst with ball-flowerlike structure for low-temperature selective catalytic reduction of NOx by NH3. Chem. Eng. J. 2020, 381, 122753. [Google Scholar] [CrossRef]
  13. Lyu, Z.K.; Niu, S.L.; Lu, C.M.; Zhao, G.J.; Gong, Z.Q.; Zhu, Y. A density functional theory study on the selective catalytic reduction of NO by NH3 reactivity of α-Fe2O3 (001) catalyst doped by Mn, Ti, Cr and Ni. Fuel 2020, 267, 117147. [Google Scholar] [CrossRef]
  14. Li, R.; Wang, P.Q.; Ma, S.B.; Yuan, F.L.; Li, Z.B.; Zhu, Y.J. Excellent selective catalytic reduction of NOx by NH3 over Cu/SAPO-34 with hierarchical pore structure. Chem. Eng. J. 2020, 379, 122376. [Google Scholar] [CrossRef]
  15. Zhao, J.J.; Yu, Y.B.; Han, X.; He, H. Fuel reforming over Ni-based catalysts coupled with selective catalytic reduction of NOx. Chin. J. Catal. 2013, 34, 1407–1417. [Google Scholar] [CrossRef]
  16. Du, H.; Han, Z.T.; Wang, Q.M.; Gao, Y.; Gao, C.; Dong, J.M.; Pan, X.X. Effects of ferric and manganese precursors on catalytic activity of Fe-Mn/TiO2 catalysts for selective reduction of NO with ammonia at low temperature. Environ. Sci. Pollut. Res. 2020, 27, 40870–40881. [Google Scholar] [CrossRef] [PubMed]
  17. Jiang, B.Q.; Wu, Z.B.; Liu, Y.; Lee, S.C.; Ho, W.K. DRIFT study of the SO2 effect on low-temperature SCR reaction over Fe-Mn/TiO2. J. Phys. Chem. C 2010, 114, 4961–4965. [Google Scholar] [CrossRef]
  18. Putluru, S.S.R.; Schill, L.; Jensen, A.D.; Siret, B.; Tabaries, F.; Fehrmann, R. Mn/TiO2 and Mn-Fe/TiO2 catalysts synthesized by deposition precipitation-promising for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2015, 165, 628–635. [Google Scholar] [CrossRef]
  19. Jiang, B.Q.; Deng, B.Y.; Zhang, Z.Q.; Wu, Z.B.; Tang, X.J.; Yao, S.L.; Lu, H. Effect of Zr addition on the low-temperature SCR activity and SO2 tolerance of Fe-Mn/Ti catalysts. J. Phys. Chem. C 2012, 118, 14866–14875. [Google Scholar] [CrossRef]
  20. Choi, H.; Jeong, Y.E.; Kumar, P.A.; Lee, K.Y.; Ha, H.P. Sb modified Fe-Mn/TiO2 catalyst for the reduction of NOx with NH3 at low temperatures. Res. Chem. Intermed. 2018, 44, 3737–3751. [Google Scholar] [CrossRef]
  21. Zhu, Y.W.; Zhang, Y.P.; Xiao, R.; Huang, T.J.; Shen, K. Novel holmium-modified Fe-Mn/TiO2 catalysts with a broad temperature window and high sulfur dioxide tolerance for low-temperature SCR. Catal. Commun. 2017, 88, 64–67. [Google Scholar] [CrossRef]
  22. Hou, X.X.; Chen, H.P.; Liang, Y.H.; Yang, X.; Wei, Y.L. Pr-doped modified Fe-Mn/TiO2 catalysts with a high activity and SO2 tolerance for NH3-SCR at low-temperature. Catal. Lett. 2020, 150, 1041–1048. [Google Scholar] [CrossRef]
  23. Hou, X.X.; Chen, H.P.; Liang, Y.H.; Wei, Y.L.; Li, Z.Q. La modified Fe-Mn/TiO2 catalysts to improve SO2 resistance for NH3-SCR at low-temperature. Catal. Surv. Asia 2020, 24, 291–299. [Google Scholar] [CrossRef]
  24. Yang, P.; Lu, C.; Hua, N.P.; Du, Y.K. Titanium dioxide nanoparticles Co-doped with Fe3+ and Eu3+ ions for photocatalysis. Mater. Lett. 2002, 57, 794–801. [Google Scholar] [CrossRef]
  25. Mao, C.Y.; Li, W.J.; Wu, F.; Dou, Y.Y.; Fang, L.; Ruan, H.B.; Kong, C.Y. Effect of Er doping on microstructure and optical properties of ZnO thin films prepared by sol-gel method. J. Mater. Sci.-Mater. Electron. 2015, 26, 8732–8739. [Google Scholar] [CrossRef]
  26. Gao, J.Q.; Jiang, R.Z.; Wang, J.; Wang, B.X.; Li, K.; Kang, P.L.; Li, Y.; Zhang, X.D. Sonocatalytic performance of Er3+:YAlO3/TiO2-Fe2O3 in organic dye degradation. Chem. Eng. J. 2011, 168, 1041–1048. [Google Scholar] [CrossRef]
  27. Vargas, M.A.L.; Casanova, M.; Trovarelli, A.; Busca, G. An ir study of thermally stable V2O5-WO3-TiO2 SCR catalysts modified with silica and rare-earths (Ce, Tb, Er). Appl. Catal. B Environ. 2007, 75, 303–311. [Google Scholar] [CrossRef]
  28. Casanova, M.; Llorca, J.; Sagar, A.; Schermanz, K.; Trovarelli, A. Mixed iron-erbium vanadate NH3-SCR catalysts. Catal. Today 2015, 241, 159–168. [Google Scholar] [CrossRef]
  29. Kim, J.; Lee, S.; Kwon, D.W.; Ha, H.P. Er composition (X)-mediated catalytic properties of Ce1-xErxVO4 surfaces for selective catalytic NOx reduction with NH3 at elevated temperatures. Catal. Today 2021, 359, 65–75. [Google Scholar] [CrossRef]
  30. Jin, Q.J.; Shen, Y.S.; Zhu, S.M.; Li, X.H.; Hu, M. Promotional effects of Er incorporation in CeO2(ZrO2)/TiO2 for selective catalytic reduction of NO by NH3. Chin. J. Catal. 2016, 37, 1521–1529. [Google Scholar] [CrossRef]
  31. Mu, J.C.; Li, X.Y.; Sun, W.B.; Fan, S.Y.; Wang, X.Y.; Wang, L.; Qin, M.C.; Gan, G.Q.; Yin, Z.F.; Zhang, D.K. Enhancement of low-temperature catalytic activity over a highly dispersed Fe-Mn/Ti catalyst for selective catalytic reduction of NOx with NH3. Ind. Eng. Chem. Res. 2018, 57, 10159–10169. [Google Scholar] [CrossRef]
  32. Yang, J.; Ren, S.; Chou, Y.H.; Su, Z.H.; Yao, L.; Cao, J.; Jiang, L.J.; Hu, G.; Kong, M.; Yang, J.; et al. In situ IR comparative study on N2O formation pathways over different valence states manganese oxides catalysts during NH3-SCR of NO. Chem. Eng. J. 2020, 397, 125446. [Google Scholar] [CrossRef]
  33. Jiang, L.J.; Liu, Q.C.; Ran, G.L.; Kong, M.; Ren, S.; Yang, J.; Li, J.L. 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]
  34. 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]
  35. Kwon, D.W.; Nam, K.B.; Hong, S.C. The role of ceria on the activity and SO2 resistance of catalysts for the selective catalytic reduction of NOx by NH3. Appl. Catal. B Environ. 2015, 166–167, 37–44. [Google Scholar] [CrossRef]
  36. Chang, H.Z.; Chen, X.Y.; Li, J.H.; Ma, L.; Wang, C.Z.; Liu, C.X.; Schwank, J.W.; Hao, J.M. Improvement of activity and SO2 tolerance of Sn-modified MnOx-CeO2 catalysts for NH3-SCR at low temperatures. Environ. Sci. Technol. 2013, 47, 5294–5301. [Google Scholar] [CrossRef]
  37. Fan, Z.Y.; Shi, J.W.; Gao, C.; Gao, G.; Wang, B.R.; Wang, Y.; He, C.; Niu, C.M. Gd-modified MnOx for the selective catalytic reduction of NO by NH3: The promoting effect of Gd on the catalytic performance and sulfur resistance. Chem. Eng. J. 2018, 348, 820–830. [Google Scholar] [CrossRef]
  38. Zhang, Y.P.; Li, G.B.; Wu, P.; Zhuang, K.; Shen, K.; Wang, S.; Huang, T.J. Effect of SO2 on the low-temperature denitrification performance of Ho-modified Mn/Ti catalyst. Chem. Eng. J. 2019, 400, 122597. [Google Scholar] [CrossRef]
  39. Gan, L.N.; Chen, J.J.; Peng, Y.; Yu, J.; Tran, T.; Li, K.Z.; Wang, D.; Xu, G.W.; Li, J.H. NOx removal over V2O5/WO3-TiO2 prepared by a grinding method: Influence of the precursor on vanadium dispersion. Ind. Eng. Chem. Res. 2018, 57, 150–157. [Google Scholar] [CrossRef]
  40. Chen, H.F.; Xia, Y.; Huang, H.; Gan, Y.P.; Tao, X.Y.; Liang, C.; Luo, J.M.; Fang, R.Y.; Zhang, J.; Zhang, W.K.; 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]
  41. Lu, Q.; Wang, Z.X.; Guo, H.Q.; Li, K.; Zhang, Z.X.; Cui, M.S.; Yang, Y.P. Selective preparation of monocyclic aromatic hydrocarbons from ex-situ catalytic fast pyrolysis of pine over Ti(SO4)2-Mo2N/HZSM-5 catalyst. Fuel 2019, 243, 88–96. [Google Scholar] [CrossRef]
  42. Ettireddy, P.R.; Ettireddy, N.; Boningari, T.; Pardemann, R.; Smirniotis, P.G. Investigation of the selective catalytic reduction of nitric oxide with ammonia over Mn/TiO2 catalysts through transient isotopic labeling and in situ FT-IR studies. J. Catal. 2012, 292, 53–63. [Google Scholar] [CrossRef]
  43. Kapteijn, F.; Vanlangeveld, A.D.; Moulijn, J.A.; Andreini, A.; Vuurman, M.A.; Turek, A.M.; Jehng, J.M.; Wachs, I.E. Alumina-supported manganese oxide catalysts: I. Characterization: Effect of precursor and loading. J. Catal. 1994, 150, 94–104. [Google Scholar] [CrossRef]
  44. Morales, M.R.; Barbero, B.P.; Cadús, L.E. Total oxidation of ethanol and propane over Mn-Cu mixed oxide catalysts. Appl. Catal. B Environ. 2006, 67, 229–236. [Google Scholar] [CrossRef]
  45. Zhang, Z.P.; Li, R.M.; Wang, M.J.; Li, Y.S.; Tong, Y.M.; Yang, P.P.; Zhu, Y.J. Two steps synthesis of cetiox oxides nanotube catalyst: Enhanced activity, resistance of SO2 and H2O for low temperature NH3-SCR of NOx. Appl. Catal. B Environ. 2021, 282, 119542. [Google Scholar] [CrossRef]
  46. Zhao, X.; Huang, L.; Li, H.R.; Hu, H.; Hu, X.N.; Shi, L.Y.; Zhang, D.S. Promotional effects of zirconium doped CeVO4 for the low-temperature selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2016, 183, 269–281. [Google Scholar] [CrossRef]
  47. Liu, Y.; Gu, T.T.; Weng, X.L.; Wang, Y.; Wu, Z.B.; Wang, H.Q. DRIFT studies on the selectivity promotion mechanism of Ca-modified Ce-Mn/TiO2 catalysts for low-temperature NO reduction with NH3. J. Phys. Chem. C 2012, 116, 16582–16592. [Google Scholar] [CrossRef]
  48. Wang, Y.L.; Li, X.X.; Zhan, L.; Li, C.; Qiao, W.M.; Ling, L.C. Effect of SO2 on activated carbon honeycomb supported CeO2-MnOx catalyst for NO removal at low temperature. Ind. Eng. Chem. Res. 2015, 54, 2274–2278. [Google Scholar] [CrossRef]
  49. Romano, E.J.; Schulz, K.H. A XPS investigation of SO2 adsorption on ceria-zirconia mixed-metal oxides. Appl. Surf. Sci. 2005, 246, 262–270. [Google Scholar] [CrossRef]
  50. Sun, C.Z.; Liu, H.; Chen, W.; Chen, D.Z.; Yu, S.H.; Liu, A.N.; 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]
  51. Wu, Z.B.; Jin, R.B.; Wang, H.Q.; Liu, Y. Effect of ceria doping on SO2 resistance of Mn/TiO2 for selective catalytic reduction of NO with NH3 at low temperature. Catal. Commun. 2009, 10, 935–939. [Google Scholar] [CrossRef]
  52. Gao, L.; Li, C.T.; Li, S.H.; Zhang, W.; Du, X.Y.; Huang, L.; Zhu, Y.C.; Zhai, Y.B.; Zeng, G.M. Superior performance and resistance to SO2 and H2O over CoOx-modified MnOx/biomass activated carbons for simultaneous Hg0 and NO removal. Chem. Eng. J. 2019, 371, 781–795. [Google Scholar] [CrossRef]
  53. Wu, Z.B.; Jiang, B.Q.; Liu, Y.; Wang, H.Q.; Jin, R.B. DRIFT study of manganese/titania-based catalysts for low-temperature selective catalytic reduction of NO with NH3. Environ. Sci. Technol. 2007, 41, 5812–5817. [Google Scholar] [CrossRef]
  54. Zuo, J.L.; Chen, Z.H.; Wang, F.R.; Yu, Y.H.; Wang, L.F.; Li, X.H. Low-temperature selective catalytic reduction of NOx with NH3 over novel Mn-Zr mixed oxide catalysts. Ind. Eng. Chem. Res. 2014, 53, 2647–2655. [Google Scholar] [CrossRef]
  55. Lian, Z.H.; Liu, F.D.; Shan, W.P.; He, H. Improvement of Nb doping on SO2 resistance of VOx/CeO2 catalyst for the selective catalytic reduction of NOx with NH3. J. Phys. Chem. C 2017, 121, 7803–7809. [Google Scholar] [CrossRef]
  56. Liu, J.; Guo, R.T.; Li, M.Y.; Sun, P.; Liu, S.M.; Pan, W.G.; Liu, S.W.; Sun, X. Enhancement of the SO2 resistance of Mn/TiO2 SCR catalyst by Eu modification: A mechanism study. Fuel 2018, 223, 385–393. [Google Scholar] [CrossRef]
  57. Jin, R.B.; Liu, Y.; Wang, Y.; Cen, W.L.; Wu, Z.B.; Wang, H.Q.; Weng, X.L. The role of cerium in the improved SO2 tolerance for NO reduction with NH3 over Mn-Ce/TiO2 catalyst at low temperature. Appl. Catal. B Environ. 2014, 148–149, 582–588. [Google Scholar] [CrossRef]
  58. Han, L.P.; Cai, S.X.; Gao, M.; Hasegawa, J.; 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]
  59. 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-Gen. 2020, 590, 117333. [Google Scholar] [CrossRef]
  60. Zhang, B.L.; Liebau, M.; Suprun, W.; Liu, B.; Zhang, S.G.; Gläser, R. Suppression of N2O formation by H2O and SO2 in the selective catalytic reduction of NO with NH3 over a Mn/Ti-Si catalyst. Catal. Sci. Technol. 2019, 9, 4759–4770. [Google Scholar] [CrossRef]
  61. 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]
  62. Xie, S.Z.; Li, L.L.; Jin, L.J.; Wu, Y.H.; Liu, H.; Qin, Q.J.; Wei, X.L.; Liu, J.X.; Dong, L.H.; Li, B. Low temperature high activity of M (M = Ce, Fe, Co, Ni) doped M-Mn/TiO2 catalysts for MH3-SCR and in situ DRIFTS for investigating the reaction mechanism. Appl. Surf. Sci. 2020, 515, 146014. [Google Scholar] [CrossRef]
Figure 1. NOx conversion (a), resistance to sulfur poisoning test (b), resistance to water vapor poisoning test (c), resistance to water vapor and sulfur poisoning test (d) during the NH3-SCR reaction over FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts. (Reaction conditions: 1 mL catalyst, [NO] = [NH3] = 500 ppm, [O2] = 5 vol.%, [SO2] = 100 ppm (when used), 5% H2O (when used), balanced with N2, GSHV = 30,000 h−1).
Figure 1. NOx conversion (a), resistance to sulfur poisoning test (b), resistance to water vapor poisoning test (c), resistance to water vapor and sulfur poisoning test (d) during the NH3-SCR reaction over FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts. (Reaction conditions: 1 mL catalyst, [NO] = [NH3] = 500 ppm, [O2] = 5 vol.%, [SO2] = 100 ppm (when used), 5% H2O (when used), balanced with N2, GSHV = 30,000 h−1).
Catalysts 11 00618 g001
Figure 2. N2 adsorption/desorption isotherms of the fresh and used FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Figure 2. N2 adsorption/desorption isotherms of the fresh and used FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Catalysts 11 00618 g002
Figure 3. XRD patterns of the fresh and used FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Figure 3. XRD patterns of the fresh and used FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Catalysts 11 00618 g003
Figure 4. H2-TPR profiles of the fresh (a) and used (b) FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Figure 4. H2-TPR profiles of the fresh (a) and used (b) FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Catalysts 11 00618 g004
Figure 5. XPS spectra of Mn 2p (a), O 1s (b) and S 2p (c) over FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Figure 5. XPS spectra of Mn 2p (a), O 1s (b) and S 2p (c) over FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts.
Catalysts 11 00618 g005
Figure 6. TGA-DTG profiles of FeMn/TiO2-S (a) and Er0.05FeMn/TiO2-S (b) catalysts.
Figure 6. TGA-DTG profiles of FeMn/TiO2-S (a) and Er0.05FeMn/TiO2-S (b) catalysts.
Catalysts 11 00618 g006
Figure 7. In-situ DRIFTS spectra of adsorption SO2 over FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts at 200 °C.
Figure 7. In-situ DRIFTS spectra of adsorption SO2 over FeMn/TiO2 and Er0.05FeMn/TiO2 catalysts at 200 °C.
Catalysts 11 00618 g007
Figure 8. In-situ DRIFTS spectra FeMn/TiO2 (a) and Er0.05FeMn/TiO2 (b) catalysts exposed to 500 ppm NH3 followed by the introduction of 100 ppm SO2 at 200 °C, and in-situ DRIFTS spectra of adsorption 500 ppm NH3 over the fresh and used catalysts at 200 °C (c).
Figure 8. In-situ DRIFTS spectra FeMn/TiO2 (a) and Er0.05FeMn/TiO2 (b) catalysts exposed to 500 ppm NH3 followed by the introduction of 100 ppm SO2 at 200 °C, and in-situ DRIFTS spectra of adsorption 500 ppm NH3 over the fresh and used catalysts at 200 °C (c).
Catalysts 11 00618 g008aCatalysts 11 00618 g008b
Figure 9. In-situ DRIFTS spectra FeMn/TiO2 (a) and Er0.05FeMn/TiO2 (b) catalysts exposed to 500 ppm NO + 5% O2 followed by the introduction of 100 ppm SO2 at 200 °C, and in-situ DRIFTS spectra of adsorption 500 ppm NO + 5% O2 over the fresh and used catalysts at 200 °C (c).
Figure 9. In-situ DRIFTS spectra FeMn/TiO2 (a) and Er0.05FeMn/TiO2 (b) catalysts exposed to 500 ppm NO + 5% O2 followed by the introduction of 100 ppm SO2 at 200 °C, and in-situ DRIFTS spectra of adsorption 500 ppm NO + 5% O2 over the fresh and used catalysts at 200 °C (c).
Catalysts 11 00618 g009
Figure 10. The promotional effect and mechanism of Er modification on SO2 resistance for NH3-SCR at low temperature over FeMn/TiO2 catalysts.
Figure 10. The promotional effect and mechanism of Er modification on SO2 resistance for NH3-SCR at low temperature over FeMn/TiO2 catalysts.
Catalysts 11 00618 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Du, H.; Han, Z.; Wu, X.; Li, C.; Gao, Y.; Yang, S.; Song, L.; Dong, J.; Pan, X. Insight into the Promoting Role of Er Modification on SO2 Resistance for NH3-SCR at Low Temperature over FeMn/TiO2 Catalysts. Catalysts 2021, 11, 618. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050618

AMA Style

Du H, Han Z, Wu X, Li C, Gao Y, Yang S, Song L, Dong J, Pan X. Insight into the Promoting Role of Er Modification on SO2 Resistance for NH3-SCR at Low Temperature over FeMn/TiO2 Catalysts. Catalysts. 2021; 11(5):618. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050618

Chicago/Turabian Style

Du, Huan, Zhitao Han, Xitian Wu, Chenglong Li, Yu Gao, Shaolong Yang, Liguo Song, Jingming Dong, and Xinxiang Pan. 2021. "Insight into the Promoting Role of Er Modification on SO2 Resistance for NH3-SCR at Low Temperature over FeMn/TiO2 Catalysts" Catalysts 11, no. 5: 618. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050618

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