Next Article in Journal
Investigation of Co3O4 and LaCoO3 Interaction by Performing N2O Decomposition Tests under Co3O4-CoO Transition Temperature
Previous Article in Journal
Effect of the Ni/Al Ratio on the Performance of NiAl2O4 Spinel-Based Catalysts for Supercritical Methylcyclohexane Catalytic Cracking
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving the Performance of Gd Addition on Catalytic Activity and SO2 Resistance over MnOx/ZSM-5 Catalysts for Low-Temperature NH3-SCR

1
School of Environment and Energy, South China University of Technology, Guangzhou Higher Education Mega Centre, Guangzhou 510006, China
2
Guangdong Provincial Key Laboratory of Atmospheric Environment and Pollution Control, South China University of Technology, Guangzhou Higher Education Mega Centre, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(3), 324; https://doi.org/10.3390/catal11030324
Submission received: 20 January 2021 / Revised: 25 February 2021 / Accepted: 26 February 2021 / Published: 3 March 2021

Abstract

:
SO2 poisoning is a great challenge for the practical application of Mn-based catalysts in low-temperature selective catalytic reduction (SCR) reactions of NOx with NH3. A series of Gadolinium (Gd)-modified MnOx/ZSM-5 catalysts were synthesized via a citric acid–ethanol dispersion method and evaluated by low-temperature NH3-SCR. Among them, the GdMn/Z-0.3 catalyst with the molar ratio of Gd/Mn of 0.3 presented the highest catalytic activity, in which a 100% NO conversion could be obtained in the temperature range of 120–240 °C. Furthermore, GdMn/Z-0.3 exhibited good SO2 resistance compared with Mn/Z in the presence of 100 ppm SO2. The results of Brunauer–Emmett–Teller (BET), X-ray photoelectron spectroscopy (XPS), temperature-programmed reduction of H2 (H2-TPR) and temperature-programmed desorption of NH3 (NH3-TPD) illustrated that such catalytic performance was mainly caused by large surface area, abundant Mn4+ and surface chemisorbed oxygen species, strong reducibility and the suitable acidity of the catalyst. The in situ diffuse reflectance infrared Fourier transform spectra (DRIFTS) results revealed that the addition of Gd greatly inhibited the reaction between the SO2 and MnOx active sites to form bulk manganese sulfate, thus contributing to high SO2 resistance. Moreover, in situ DRIFTS experiments also shed light on the mechanism of low-temperature SCR reactions over Mn/Z and GdMn/Z-0.3, which both followed the Langmuir–Hinshelwood (L–H) and Eley–Rideal (E–R) mechanism.

Graphical Abstract

1. Introduction

Nitrogen oxides (i.e., NOx, mainly in the forms of NO and NO2) emitted from stationary sources have caused serious environmental problems, including photochemical smog, acid rain, ozone depletion and the greenhouse effect, thus they are a kind of extremely harmful air pollutants to human beings [1,2,3,4]. Selective catalytic reduction (SCR), using ammonia as a reductant, is considered to be one of the most prospective technologies in the abatement of NOx from stationary sources, which is attributed to the core of catalysts with economic and technological efficiency. Although the commercial catalysts (V2O5-WO3(MoO3)/TiO2) in traditional SCR techniques made great contributions in the removal of NOx, they suffered extreme challenges, such as a narrow temperature window (300–400 °C), the biological toxicity (from vanadium) of the active component, the over-oxidation of SO2 or NH3, the deactivation in the high concentration of dust and SO2 poisoning [5,6]. Therefore, some vanadium-free catalysts with high activity and good stability, as well as SO2-resistance ability for NOx removal in the low-temperature SCR (<250 °C) process should be urgently explored.
In the past few decades, plenty of transition metal oxides (FeOx, MnOx, CuOx and CoOx, etc.) supported on different carriers were widely studied for the SCR of NOx [1,2,3]. Among them, Mn-based catalysts are universally considered as suitable alternatives for substituting vanadium-based SCR catalysts due to the properties of variable valence and redox [7], such as MnOx/SAPO-34 [8], MnOx/ZSM-5 [9], MnOx/TiO2 [10,11], MnOx/Al2O3 [12] and MnOx loaded on carbon nanotubes [13]. Particularly, owing to the high surface area, abundant acid sites and the thermal stability, ZSM-5 zeolite has attracted great attention in Mn-based SCR reactions [14,15,16]. Specifically, ZSM-5 can facilitate the adsorption of NH3 and activate the adsorbed NH3 to intermediate species in SCR reactions, which is beneficial to the NH3-SCR [17]. Besides, both the large surface area and abundant porous structure contribute to the high dispersion of active components. For example, Shi et al. reported that the Fe-ZSM-5 catalyst exhibited the superior performance, i.e., widened active temperature window, good SO2 resistance and hydrothermal stability, which is ascribed to the good dispersion of active iron species on ZSM-5 [18]. Therefore, ZSM-5 is considered as a promising support candidate for enhancing SCR performance. However, SO2 poisoning has been a critical issue for the Mn-based SCR catalysts, which limited their practical application and greatly drew our interest [19]. Research on promoting the SO2 resistance of catalysts can be divided into three aspects, including reducing the adsorption of SO2, blocking the oxidization of SO2 to SO3 and lowering the deposition of active metal sulfate or ammonium sulfate [20]. Currently, the introduction of many rare earth metals into Mn-based catalysts exhibits positive effects on improving SO2 resistance, which can be attributed to the strong interaction between the additives and active components to suppress SO2 poisoning [21,22,23,24]. You et al. reported that Ce-modified MnOx supported by graphene presented good activity and high-resistance ability to SO2 and H2O [25]. Liu et al. found that the introduction of Eu into a Mn/TiO2 catalyst was advantageous to prevent the formation of sulfate species, and thus contributed to an excellent SO2 tolerance for the SCR reaction process [26]. According to the ion polarization theory, Yu et al. revealed that the introduction of Pr into MnOx/SAPO-34 could enhance SO2 resistance, which made the sample difficult to be sulfated and restrained the generation of sulfates compared with Ce-doping [27]. Nevertheless, the enhanced mechanism of Mn-based catalysts modified by the different additives on catalytic activity and SO2 resistance is not very clear, and few studies have been done on the mechanism and the promotion of SO2 resistance in low-temperature SCR reactions over Mn-based catalysts modified by rare earth metals. Therefore, it is worth exploring further.
Inspired by the superior properties of rare earth metals, we were very interested in investigating the effect of their addition on activity and SO2 resistance of low-temperature SCR catalysts. So far numerous studies have been developed to improve the catalytic performance of the catalysts, but explorations of the addition of Gd into MnOx/ZSM-5 catalysts (Gd-MnOx/ZSM-5) are still rare [28]. Herein, MnOx/ZSM-5 and Gd-MnOx/ZSM-5 catalysts with different molar ratios of Gd/Mn were synthesized and used for NO removal in the presence of SO2 in low-temperature NH3-SCR. In comparison with MnOx/ZSM-5, the addition of an appropriate amount of Gd (Gd/Mn = 0.3) could not only improve the activity, but also significantly promote the SO2 tolerance of the sample. Meanwhile, the relationship between material properties and catalytic performance was studied through a series of characterizations. Moreover, the key role of Gd in catalytic activity and SO2 resistance was further investigated by in situ diffuse reflectance infrared Fourier transform spectra (DRIFTS) experiments, providing powerful support for further improvement and optimization of catalysts in the future.

2. Results

2.1. Low-Temperature NH3-SCR Performance

The SCR activity of GdMn/Z-x (Gd-modified MnOx/ZSM-5 denoted as GdMn/Z-x) catalysts was evaluated for NO conversion and the results are shown in Figure 1. All the synthesized catalysts presented excellent activity in low-temperature NH3-SCR. Notably, the NO conversion over the GdMn/Z-0.3 catalyst was higher than that of Mn/Z (MnOx/ZSM-5 denoted as Mn/Z) in the range of 80–140 °C, and nearly 100% conversion was obtained at as low as 120 °C. In particular, the catalytic activity of the GdMn/Z-0.3 was competitive or even better than previously reported catalysts under similar condition [29,30], as 90% NO conversion was obtained below 120 °C in this work. The NO conversion decreased with the increase of Gd/Mn, and the worst performance was achieved at 0.6, because the main active sites were covered by the excess GdOx particles on the surface of the catalyst [31]. Moreover, the N2 selectivity (Figure 1b) and stability tests (Figure S1) were also performed to evaluate the practical value of the GdMn/Z-0.3 catalyst. The results displayed that GdMn/Z-0.3 consistently presented above 96% N2 selectivity in low temperatures, and 97%–99% NO conversion was retained in 24 h at 180 °C. It could be concluded that among all the as-prepared catalysts for the NH3-SCR of NO, the GdMn/Z-0.3 catalyst exhibited the optimal catalytic performance with a broadening active temperature window and high N2 selectivity, as well as good stability.

2.2. Physicochemical Properties of the Fresh Catalysts

2.2.1. X-ray Diffraction (XRD) and Brunauer–Emmett–Teller (BET) Analyses

The characteristic diffraction peaks of all the as-prepared catalysts are shown in Figure 2, and all these peaks corresponded well to ZSM-5 zeolite with the standard card of PDF#44-0638. Notably, no obvious diffraction peaks of GdOx or MnOx could be observed, suggesting that the components might be highly dispersed or in the amorphous state [15,32]. Moreover, with the increase of Gd/Mn, the diffraction intensities in the XRD patterns of the catalysts became weakened and followed the order: Mn/Z > GdMn/Z-0.1 > GdMn/Z-0.3 > GdMn/Z-0.6. These results revealed that the addition of Gd could inhibit the formation of large MnOx bulks to some extent and cause the poor crystallization of the Mn/Z catalyst [33]. The textural properties of the samples are summarized in Table 1. It has been universally acknowledged that the high specific surface area greatly contributes to the adsorption of reactants, further resulting in high activity by offering more active sites [34]. The GdMn/Z-0.3 catalyst had the largest specific surface area of 273.4 m2/g, corresponding to the optimal low-temperature NH3-SCR activity (Figure 1). It was interesting that all the catalysts possessed the mesopores structure, indicating that the inherent micropores of ZSM-5 disappeared. This might be due to the fact that after the loading process, the surface structure could be reconstructed. The results revealed that the Mn/Z catalyst modified by a certain amount of Gd (Gd/Mn = 0.3) could increase the specific surface area and pore volume as well as average pore size, which was advantageous to improve the catalytic performance in low-temperature SCR reactions.

2.2.2. Microstructure Analysis

Scanning electron microscopy (SEM) images of the catalysts are shown in Figure S2. It can be seen that all catalysts mainly exhibited the typical morphology of the ZSM-5 zeolite. With the addition of Gd, the morphology of the Mn/Z (Figure S2a) catalyst was changed significantly. The original hexahedral block structure of Mn/Z was decomposed into a polyhedral block structure with different sizes (Figure S2c,e,g), while the smaller hexahedral block morphology could still be observed in all three Gd-containing samples, which might be related to the decrease of crystallinity, as observed in the XRD patterns (Figure 2). In order to further characterize the morphology and microstructure of the active component loaded on the support of Mn/Z and optimal GdMn/Z-0.3 catalysts, transmission electron microscopy (TEM) analysis was performed, and the results are displayed in Figure 3. It was apparent that many small component particles were well dispersed on the surface or edge of the support for Mn/Z (Figure 3a,b) and GdMn/Z-0.3 (Figure 3d,e). In contrast, the particles of active components on Mn/Z were bigger and more dispersed than those on GdMn/Z-0.3. It can be clearly seen from the high-resolution TEM (HR-TEM) image of Mn/Z (Figure 3c) that the lattice fringe of Mn/Z and the corresponding lattice space were 0.24, 0.31, 0.33 and 0.49 nm, which could be ascribed to the (101) plane, (111) plane of MnO2, (220) plane of Mn2O3 and (101) facet of Mn3O4, respectively [35,36,37], confirming the presence of MnO2, Mn2O3 and Mn3O4. As for the GdMn/Z-0.3 (Figure 3f), there were more amorphous and fuzzy areas, except for a small area where the lattice fringe corresponding to the (101) facet of MnO2 was clear. Additionally, the energy dispersive spectroscopy (EDS) elemental mappings over the GdMn/Z-0.3 catalyst are also shown in Figure 3h–k. The results confirmed that Gd and Mn species were dispersed uniformly on the surface of the support. Moreover, it was possible that Gd tended to enrich on the surface and decrease the particle size of MnOx.

2.2.3. X-ray Photoelectron Spectroscopy (XPS) Analysis

The surface atomic information for the as-prepared catalysts were further analyzed by XPS. Figure 4 exhibits the XPS spectra of Mn 2p and O 1s. The Mn 2p mainly had two peaks of Mn 2p1/2 (649–657 eV) and Mn 2p3/2 (636–646 eV), and the Mn 2p3/2 peak could be deconvoluted to three valence state peaks in Figure 4a, which were assigned to Mn2+ (641 ± 0.5 eV), Mn3+ (642 ± 0.5 eV) and Mn4+ (643 ± 0.5 eV), respectively [38,39]. The relative concentration of Mn2+/Mnn+, Mn3+/Mnn+ and Mn4+/Mnn+ calculated from the XPS spectra are listed in Table 2. Obviously, GdMn/Z-0.3 possessed the highest relative ratios of Mn4+ and Mn3+ species. It has been known that Mn4+ is the most active species to accelerate the fast SCR reaction, owing to the oxidation of NO to NO2, which contributes to the excellent activity in low-temperature SCR [24,40]. The highest concentration of Mn4+ in GdMn/Z-0.3 could be invoked to explain its superior activity compared to the other catalysts with the proper amount of Gd addition. In Figure 4b, two types of surface oxygen were observed in the XPS spectra of O 1s on the catalysts. The low-binding-energy peaks at 529.5 ± 0.5 eV could be attributed to the lattice oxygen O2− (OL), while the high-binding-energy peaks at 532.0 ± 0.5 eV could be attributed to the surface chemisorbed oxygen originating from defect oxide and hydroxyl-like groups, such as O22− and O (OS), respectively [41]. Generally, the surface chemisorbed oxygen species were highly active species compared with lattice oxygen [31]. GdMn/Z-0.3 exhibited a relatively high ratio of Os/(OL + OS) among all the catalysts and further resulted in a fast SCR reaction, which was beneficial for the low-temperature SCR activity [42]. Consequently, the excellent SCR activity of GdMn/Z-0.3 was ascribed to the high concentration of active Mn4+ and abundant surface chemisorbed oxygen species affected by the addition of Gd.

2.2.4. Temperature-Programmed Reduction of H2 (H2-TPR) and Temperature-Programmed Desorption of NH3 (NH3-TPD) Analyses

The redox properties of the catalysts were characterized by H2-TPR experiments, and the results are shown in Figure 5a and Table 3. The area of the reduction peak was correlated to the consumption of H2, and the results (Figure 5a) suggested that the GdMn/Z-0.3 catalyst possessed high-reduction capability. Both the Mn/Z and GdMn/Z-x catalysts mainly exhibited one wide reduction peak in the temperature range of 260–420 °C, which could be ascribed to the two main reduction steps of MnO2 and Mn2O3 to Mn3O4 [9,43]. Compared with the Mn/Z catalyst, the reduction peak of GdMn/Z-0.3 obviously shifted to a lower temperature, corresponding to the excellent low-temperature SCR activity. Meanwhile, combined with the results of XPS, the rich and high-valence state of Mn species, such as Mn4+ species, could play a significant role in the redox reaction on the GdMn/Z-0.3 catalyst surface because of the strong interaction between MnOx and GdOx [25,44]. Therefore, a certain amount of Gd could improve the redox ability of Mn/Z catalysts.
The surface acidity of a catalyst is an essential factor for NH3 adsorption, thus contributing to the high activity of the catalyst in the SCR reaction process [20]. The amount of acid sites and the acid strength of all the catalysts were investigated by NH3-TPD, and the results are shown in Figure 5b. It can be seen that NH3-TPD profiles of the catalysts presented two wide desorption peaks. The desorption peaks in the low-temperature range of 100–300 °C could be assigned to the adsorption of NH3 on weak acid sites, while the desorption peaks in the high-temperature range of 350–650 °C could be ascribed to the medium-strong acid sites [45,46]. As shown in Table 3, the total acid amount of the catalysts was ranked by Mn/Z < GdMn/Z-0.6 < GdMn/Z-0.3 < GdMn/Z-0.1. Both the weak acid and medium-strong acid of Mn/Z could be improved by the introduction of Gd. Interestingly, although the total acid amount of the GdMn/Z-0.3 catalyst was not the maximum, it still exhibited excellent activity. The results indicated that a suitable amount of weak acid and medium-strong acid were conducive to the adsorption and activation of NH3, which could enhance the catalytic performance of catalysts in NH3-SCR reactions.

2.3. In Situ DRIFTS Studies for NH3-SCR

2.3.1. Adsorption of NH3

Figure 6 exhibits the in situ DRIFT spectra of the Mn/Z and GdMn/Z-0.3 catalysts during the reaction of 800 ppm NH3 at 180 °C as a function of time. Several bands were presented at 3660, 3605, 3369, 3279, 3181, 1729, 1606, 1579, 1508, 1433, 1299, 1185 and 1040 cm−1 for NH3 adsorption over the Mn/Z catalyst, as shown in Figure 6a. On the basis of the previous literature [47,48], the negative bands at 3660 and 3605 cm-1 were in the O–H stretching vibration modes, owing to the interaction of the surface hydroxyls with NH3. The bands at 3369–3181 cm−1 were derived from the N–H stretching vibration modes of coordinated NH3 on Lewis acid sites. The features at 1729 and 1433 cm−1 could be due to the asymmetric and symmetric bending vibrations of NH4+ species on Brønsted acid sites, respectively [24,49]. The bands at 1606, 1579, 1299, 1185 and 1040 cm−1 could be attributed to the coordinated NH3 on Lewis acid sites [49,50,51], while the band at 1508 cm−1 could be ascribed to the scissoring vibration mode of amide (−NH2) species [50]. It could be seen that all these bands were strengthened along with the increase of NH3 feeding time, and reached a maximum value in 15 min. Similarly, NH3 adsorption on Brønsted acid sites (1737, 1468 and 1436 cm−1) and Lewis acid sites (3356–3187, 1605, 1568, 1273, 1181 and 1045 cm−1) were also observed on the GdMn/Z-0.3 catalyst (Figure 6b). All these NH3 adsorption bands reached a maximum value at around 15 min. Differently, the bands’ intensities of NH3 adsorption on the GdMn/Z-0.3 catalyst were slightly stronger than those on the Mn/Z catalyst, demonstrating that the introduction of Gd into the Mn/Z catalyst enhanced the surface acidity of the Mn/Z catalyst, which was also evidenced by the NH3-TPD results.

2.3.2. Co-Adsorption of NO and O2

In order to investigate the NOx adsorption states on the as-prepared catalysts, in situ DRIFTS of the Mn/Z and GdMn/Z-0.3 catalysts during 800 ppm NO and 5% O2 at 180 °C were performed. As displayed in Figure 7, the NOx absorption bands emerged immediately after feeding the NO and O2 at 180 °C for 1 min and strengthened with the continuous introduction of NO and O2. On the surface of Mn/Z (Figure 7a), the adsorption band of gaseous or weakly adsorbed NO (1885 cm−1) was observed [52]. In the typical scope of surface nitrates species (1500–1700 cm−1), the bands at 1686, 1637, 1594 and 1523 cm−1 could be assigned to the nitro species, adsorbed NO2 molecules, bridged nitrate and bridged monodentate nitrate, respectively [47,52,53]. While the peaks centered at 1348, 1228 and 1085 cm−1 were separately ascribed to the free nitrate, bridged nitrate and linear nitrite species [54,55]. For the GdMn/Z-0.3 catalyst, the linear nitrite (1098 cm−1), bridged nitrate (1282 cm−1), monodentate nitrate (1546 cm−1) and gaseous NO2 molecules (1629 cm−1) were observed at 180 °C (Figure 7b) [52,53,54,55]. The bands’ intensities belonging to the adsorbed NOx species over the GdMn/Z-0.3 catalyst were much stronger than those on the Mn/Z sample, which might be due to the strong redox on the GdMn/Z-0.3 catalyst, facilitating the NO oxidation and NOx adsorption.

2.3.3. Reaction between NO and O2 and Pre-Adsorbed NH3

In consideration of the distinct surface acidity and NOx adsorption capacity of the Mn/Z and GdMn/Z-0.3 catalysts, the surface intermediates and the reaction pathways on the two catalysts might potentially differ during the NH3-SCR process. In situ DRIFTS experiments were performed to study the evolution of surface species during de-NOx. The spectra of Figure 8a were recorded after adsorption of 800 ppm NH3 for 30 min over the Mn/Z catalyst, with Ar purge for 30 min at 180 °C to remove any gaseous residues. It could be clearly seen that some adsorbed ammonia species included NH4+ ions on Brønsted acid sites (1729 and 1433 cm−1), coordinated NH3 on Lewis acid sites (1606, 1579, 1299, 1185 and 1040 cm−1), as well as amide (−NH2) species (1508 cm−1) that appeared on the Mn/Z catalyst [47,49,56]. The evolution of the surface species was monitored after the inlet gas was switched to 800 ppm NO and 5% O2. The peak intensities of the adsorbed ammonia species gradually decreased, and some bands disappeared in the end, while some ammonia species still existed after introducing NO and O2 for 30 min. Moreover, two new bands corresponding to the nitro and nitrate species (1675 and 1633 cm−1) were generated on the Mn/Z catalyst [47,52]. This phenomenon indicated that adsorbed NH3 species could react with adsorbed NOx or gaseous NO2 to form N2 and H2O [20]. Importantly, amide (-NH2) species were important intermediates in the Eley–Rideal (E–R) reaction mechanism, signifying that adsorbed NH3 species were first activated to -NH2 species, and then consumed by gaseous NO2 (1633 cm−1). As for the GdMn/Z-0.3 catalyst (Figure 8b), a similar phenomenon occurred during the procedure of reaction between NO, O2 and pre-adsorbed NH3. Differently, the adsorbed ammonia species on the GdMn/Z-0.3 catalyst disappeared absolutely after feeding NO and O2 for 15 min. Meanwhile, some new bands ascribed to the nitrate species appeared (1629, 1546, 1326 and 1223 cm−1), which were stronger than those on the Mn/Z catalyst. The phenomena indicated that the adsorbed ammonia species were easily consumed by the adsorbed NOx species, and NOx was easily adsorbed on the GdMn/Z-0.3 catalyst because of its intrinsic redox property, which resulted in the higher low-temperature SCR activity.

2.3.4. Reaction between NH3 and Pre-Adsorbed NO and O2

Reversing the order of reactants, as shown in Figure 9, after switching the gas to NH3, an apparent difference in the reaction process was observed over the Mn/Z and GdMn/Z-0.3 catalysts. For Mn/Z sample, some nitrate/nitrite species appeared, such as NO2 molecules (1637 cm−1), bridged nitrate (1594 cm−1 and 1228 cm−1), monodentate nitrate (1523 cm−1), free nitrate species (1348 cm−1) and linear nitrite (1085 cm−1), which was mainly resulted from the pre-adsorbed NO and O2 as recorded in Figure 9a [52,53,54,55]. Upon switching the gas to NH3, all these bands got weaker and eventually disappeared for a short time of 3 min. Next, characteristic peaks belonging to ammonia species (3369–3181, 1723, 1606, 1579, 1502, 1433, 1305, 1185 and 1040 cm−1) were observed, indicating that adsorbed NOx could react with adsorbed NH3 [47,49,50,51,56], and finally ammonia species accumulated on the catalyst surface, which followed the Langmuir–Hinshelwood (L–H) mechanism. Besides, some similar phenomena were also observed on GdMn/Z-0.3 (Figure 9b) catalyst. However, the difference was that the NOx adsorption amount on GdMn/Z-0.3 catalyst was significantly higher than that on Mn/Z sample, and the adsorbed NOx disappeared within the same time of NH3 introduction, indicating that the reaction between NH3 and pre-adsorbed NO and O2 on GdMn/Z-0.3 was faster than that on Mn/Z sample.

2.3.5. In Situ DRIFTS of Standard NH3-SCR Reaction on the Catalysts

Figure 10 presents the in situ DRIFTS experiments of the co-adsorption of 800 ppm NO, 800 ppm NH3 and 5% O2 for Mn/Z and GdMn/Z-0.3 at different times. The bands belonging to coordinated NH3 species (3369, 3276, 3188, 1609, 1308, 1186 and 1039 cm−1), and nitro/nitrate species (1685, 1637 and 1584 cm−1) were observed simultaneously on the surface of the Mn/Z sample at 180 °C (Figure 10a). This fact demonstrated that the NH3-SCR reaction over the Mn/Z catalyst followed a L–H mechanism due to the simultaneous adsorption of these reactant molecules (NH3, NO and O2) on the surface of the catalyst. More specifically, the adsorbed NO2 molecules (1637 cm−1) and bridged nitrates (1584 cm−1) vanished within 15 min, while the nitro species appeared at 1685 cm−1, slightly increased after exposure to NO, NH3 and O2. Moreover, all these peaks ascribed to the adsorbed NH3 species were strengthened with the time. It was clearly proven that the nitro species located at 1685 cm−1 was more stable than the NO2 molecules (1637 cm−1) and bridged nitrates (1594 cm−1). As for the GdMn/Z-0.3 (Figure 10b) sample, it could be found that the spectra of NO, NH3 and O2 co-adsorption were different from those of NO and O2 adsorption and NH3 adsorption over the GdMn/Z-0.3 catalyst, as shown in Figure 6 and Figure 7. In particular, only one peak belonging to the stable nitro species (1685 cm−1) could be detected apart from the adsorbed NH3 species (3362–3178, 1732, 1613, 1473, 1428, 1273, 1175 and 1043 cm−1), and all these band intensities were weaker than those observed in Figure 6 and Figure 7, which is because of the occurrence of the NH3-SCR reaction during the co-adsorption of NO, NH3 and O2 process. In other words, the NH3-SCR reaction route over the GdMn/Z-0.3 catalyst not only followed an L–H mechanism, but also an E–R mechanism, which manifested that the introduction of Gd only altered the reaction rate of the SCR reaction, and did not influence the reaction mechanism of NH3-SCR over both catalysts.

2.4. SO2 Resistance Studies

The impact of SO2 on the low-temperature NH3-SCR reaction with the initial NO conversion of nearly 100% and 80% over the Mn/Z and GdMn/Z-0.3 catalysts were tested respectively, and the results are shown in Figure 11. It could be clearly seen that the NO conversion of the Mn/Z catalyst decreased dramatically from nearly 100% to 32% during the 3 h test after the SO2 stream was added in. Compared with Mn/Z, the GdMn/Z-0.3 catalyst presented excellent SO2 resistance of about 85% NO conversion after the 3 h SO2 test. Interestingly, when the SO2 steam was cut off, GdMn/Z-0.3 restored to nearly 97% NO conversion while Mn/Z was almost unchanged. Notably, the SO2 resistance of the GdMn/Z-0.3 catalyst in this work was competitive or even better than that reported in the literature under similar condition [57,58]. For example, the high NO conversion of GdMn/Z-0.3 was almost unaffected in the presence of SO2 for 1 h, which was much better than previous studies in which NO conversion decreased from 100% to 60% [46]. In addition, it could be seen that the catalytic activity of the GdMn/Z-0.3 catalyst decreased slowly from the initial 80% NO conversion to 63% when SO2 was introduced, while the Mn/Z catalyst dropped sharply to nearly 25%. It is known that the decline of catalytic activity in the presence of SO2 could be attributed to the competitive adsorption of NO and SO2 [59], indicating that Gd could inhibit the adsorption of SO2 and the reaction between SO2 and active MnOx, which could be confirmed by in situ DRIFTS results. Furthermore, in comparison with the sole SO2 at 180 °C, the NO conversion of two catalysts decreased obviously in the presence of the SO2 and H2O stream (Figure S3). Inspiringly, the initial NO conversion of GdMn/Z-0.3 only decreased to nearly 80%, while Mn/Z sharply fell to nearly 14%, suggesting that Gd could not only significantly enhance the low-temperature catalytic activity, but also improve the SO2 and/or H2O resistance of the Mn/Z catalyst.

2.5. XRD, SEM and XPS Analyses of the Used Samples

Both the properties of Mn/Z and GdMn/Z-0.3, after being treated at 180 °C in the presence of 100 ppm SO2 (denoted as Mn/Z-S and GdMn/Z-0.3-S), were further investigated by XRD, SEM and XPS analyses. Figure S4 exhibits the XRD patterns of the catalysts after the SO2 poisoning, and there were no characteristic peaks of sulfate species observed on Mn/Z-S and GdMn/Z-0.3-S. Compared with the fresh catalysts, the intensities of the two treated catalysts slightly decreased, indicating that some new amorphous species possibly formed and covered on the surface below the detection limit. SEM was performed to investigate the changes of the catalysts’ morphology after the reaction with SO2. As shown in Figure 12, relatively large deposited particles or agglomerations were observed on the surface of the Mn/Z-S sample after the 100 ppm SO2 test. In contrast, there was an almost smooth surface over the GdMn/Z-0.3-S sample, suggesting that metal sulfate or ammonium sulfate species were not easily formed after the addition of Gd, and consequently, the GdMn/Z-0.3 catalyst exhibited outstanding SO2 resistance.
XPS experiments were carried out to investigate the changes of the elements’ existing states on the two catalysts in the presence of SO2, and the results are exhibited in Figure 13. After a 3 h SO2 test, the atomic ratio of Mn4+ and Mn3+ on Mn/Z-S and GdMn/Z-0.3-S was less than that of the fresh catalysts, but the proportion of Mn4+ and Mn3+ over GdMn/Z-0.3-S was still more than that of Mn/Z-S, demonstrating that SO2 could affect the oxidation state of Mn species. In addition, more Mn3+ located on the surface of GdMn/Z-0.3-S might become the Lewis acid center for NH3 adsorption in low-temperature SCR reactions, which was attributed to the high NO removal efficiency in the presence of 100 ppm SO2 [54]. Compared with the fresh catalysts, the ratio of OS/(OL + OS) of the used catalysts obviously decreased, which might account for the falling activity after the SO2 poisoning. Particularly, the concentration of OS on GdMn/Z-0.3-S was higher than on Mn/Z-S, corresponding to the better NO conversion in the SO2 stream.

2.6. In Situ DRIFTS Studies for NH3-SCR with SO2

2.6.1. SO2 Adsorption

In order to study the adsorption capacity of SO2 on the two catalysts, the DRIFTS of 100 ppm SO2 and 5% O2 co-adsorption at 180 °C were conducted to investigate the existence states of sulfates. As shown in Figure 14a, several distinct peaks at 1635, 1589, 1471, 1431, 1343, 1298, 1186 and 1040 cm−1 were detected on the Mn/Z catalyst after introducing SO2 and O2 for 30 min. As reported in the previous literature [60,61,62], the bands at 1040 cm−1 could be attributed to the symmetric stretching vibrations of O–S–O species. The bands at 1298 and 1343 cm−1 were ascribed to the asymmetric stretching frequencies of O–S–O species and coordinated sulfate species, respectively. The bands at 1431 and 1471 cm−1 were usually assigned to the symmetrical stretching vibration of gaseous SO2, while the bands at 1589 and 1635 cm−1 were associated with the adsorbed HSO4 generated by the interaction between SO2 and surface hydroxyl groups. A weak peak at 1186 cm−1 over Mn/Z appeared, which was possibly because of the bulk-like sulfates in MnO2 (Mn(SO4)2) [62,63]. Notably, only three weak bands belonging to the surface sulfate species (1630 and 1295 cm−1) and gaseous SO2 (1473 cm−1) formed on the GdMn/Z-0.3 catalyst (Figure 14b), demonstrating that Gd as a sacrificial site potentially greatly inhibited the reaction between the SO2 and MnOx active sites to form bulk Mn(SO4)2 or Mn2(SO4)3 on the catalyst surface, and further reduced the sulfurization rate [62]. It was reported that the reaction between SO2 and metal oxide active sites could inactivate these active sites, resulting in the catalyst deactivation [23]. Fortunately, the introduction of Gd effectively suppressed the generation of Mn(SO4)2, which contributed to the excellent SO2 resistance of the GdMn/Z-0.3 catalyst.

2.6.2. Reaction between NH3, NO and O2 (SO2) and Pre-Adsorbed SO2 (NH3, NO and O2)

The influence of sulfur species on the adsorption of NH3 and NOx species was further investigated over the Mn/Z and GdMn/Z-0.3 catalysts at 180 °C. As shown in Figure 14 and Figure 15, several peaks belonging to various sulfur species were detected on the Mn/Z and GdMn/Z-0.3 catalysts after adsorption of 100 ppm SO2 and 5% O2 for 30 min [60,61,62]. All these bands remained almost unchanged for both catalysts after Ar purging for 30 min (Figure 15). Compared with the above DRIFT spectra of the NH3, NO and O2 co-adsorption depicted in Figure 10, the adsorption capacity of NOx on Mn/Z obviously decreased after SO2 adsorption (Figure 15a), indicating competitive adsorption between NOx and SO2. The latter had a remarkably stronger adsorption capacity than the former. Thus, the low-temperature NH3-SCR reaction between the adsorbed NOx and NH3 species following the L–H mechanism was repressed by the presence of SO2, causing the serious deactivation of Mn/Z under the atmosphere of SO2. As for the GdMn/Z-0.3 catalyst (Figure 15b), the DRIFT spectra were similar with those exhibited in Figure 10 conducting the standard NH3-SCR reaction, demonstrating that the low-temperature NH3-SCR reaction route over GdMn/Z-0.3 could remain unchanged in the presence of SO2, owing to the small amount of adsorption of sulfur species on the GdMn/Z-0.3 catalyst. Herein, the GdMn/Z-0.3 catalyst exhibited excellent SO2 resistance.
Notably, the attribution of the peaks in Figure 16 was consistent with that in Figure 6 and Figure 7. It could be seen that the original adsorption peaks of NOx and NH3 species located at a low wavenumber (1000–2000 cm−1) over the Mn/Z catalyst almost remained unchanged after feeding SO2, while the peaks at a high wavenumber (3100–3400 cm−1) weakened gradually. In combination with Figure 6, Figure 7 and Figure 14, it could be concluded that the peaks of the sulfur species on the Mn/Z catalyst almost overlapped with the peaks of the adsorbed NH3 and NOx species (Figure 16a). Thus, the strength of these peaks was almost unchanged in the whole process, which was probably caused by the adsorption of the sulfur species. In addition, the adsorbed NH3 species on GdMn/Z-0.3 gradually decreased with the time of SO2 feeding (Figure 16b), and the adsorption peaks of SO2 finally appeared at 1473 and 1295 cm−1, which were weaker than those on Mn/Z. It could be seen from Figure 16 that the adsorption of SO2 on GdMn/Z-0.3 was still weaker than that on Mn/Z, even in the case of NH3 and NOx pre-adsorption.

2.6.3. Co-Adsorption of SO2, NH3, NO and O2

When SO2 was introduced into the standard SCR reaction atmosphere, the position of the bands in the Mn/Z catalyst (Figure 17a) was similar to that described in the standard SCR reaction over the Mn/Z catalyst (Figure 10a). Differently, the bands located at 1583 and 1635 cm−1, corresponding to adsorbed NOx species, did not disappear within 30 min, while they were almost consumed within 10 min in the standard NH3-SCR reaction without SO2, indicating the presence of SO2 could lead to a decrease in the reaction rate, which was also the reason for its decreased activity. In addition, it was surprising that the peak at 1583 cm−1 attributed to bridged nitrates increased slightly with the presence of SO2 in the initial period, owing to the overlapping of the adsorbed sulfates. As for the GdMn/Z-0.3 catalyst (Figure 17b), no bands ascribed to sulfate species were observed, while NOx (1685 cm−1) and NH3 adsorption (1726 cm−1) were strengthened. Compared with the standard NH3-SCR reaction conditions, the reaction rate was hardly changed. Therefore, it could be speculated that the presence of SO2 in the reaction atmosphere had little influence on the reaction path over the GdMn/Z-0.3 catalyst, contributing to its outstanding SO2 resistance.

3. Materials and Methods

3.1. Catalyst Preparation

The Gd-modified MnOx/ZSM-5 catalysts were prepared using a citric acid–ethanol dispersion method. All the reagents were of the required analytical grade and had no further purification. ZSM-5 zeolite (Si/Al = 38, XFNANO Company, Nanjing, China) was used as support material. Manganese acetate tetrahydrate (Mn(CH3COO)2·4H2O, 99.9%, Aladdin, Shanghai, China) and gadolinium nitrate hexahydrate (Gd(NO3)3·6H2O, 99.9%, Aladdin, Shanghai, China) were used as the precursors of the manganese oxides and gadolinium oxides, respectively. First of all, a certain amount of Mn(CH3COO)2·4H2O (Mn loading = 15 wt%) and Gd(NO3)3·6H2O were dissolved in 120 mL ethanol and then stirred vigorously for 30 min to dissolve precursors completely under room temperature. Next, a certain amount of ZSM-5 zeolite was slowly added to the above mixture and stirred continuously. Then, the mixed solution was sonicated for 45 min, followed by the addition of citric acid, and further stirred in an 80 °C water bath until the solvent was completely evaporated to obtain a colloidal solid. Finally, the product was put in a 110 °C oven to dry for 12 h, and then calcined in air at 400 °C for 4 h. The synthesized catalysts were denoted as GdMn/Z-x, where x represented the mole ratio of Gd and Mn (x = 0, 0.1, 0.3 and 0.6). For comparison, the MnOx/ZSM-5 catalyst was also prepared using the same method and denoted as Mn/Z.

3.2. NH3-SCR Activity Test

The NH3-SCR activity evaluation was carried out in a fixed-bed quartz flow tube reactor (12 mm inside diameter and 600 mm in length) under atmospheric pressure in the reaction temperature range of 80–240 °C. Before the tests, all the catalysts were crushed and sieved as 40–60 mesh. The feed gases were designed as follows: 800 ppm NO, 800 ppm NH3, 5 vol.% O2, 100 ppm SO2 (when needed) and 10% H2O (when needed), and balanced by Ar. In order to obtain a gas hourly space velocity (GHSV) of 40,000 h−1, the rate of the total flow was 600 mL min−1, and each sample volume was 0.9 mL. The inlet and outlet concentrations of NOx (NO and NO2) from the fixed-bed quartz flow tube reactor were measured by a NO–NO2–NOx analyzer (model 42i-HL, Thermal Scientific, Waltham, MA, USA). At the same time, the outlet concentration of the product N2 from the reactor was also monitored by a gas chromatograph (GC9560, Huaai, Shanghai, China) with a thermal conductivity detector (TCD) and 5A columns. The conversion of NO and the selectivity of N2 in the low-temperature NH3-SCR reaction were calculated by the following formulas:
NO   conversion = C NO in C NO out C NO in × 100 %
N 2   selectivity = C N 2 out C NO x in C NO x out × 100 %
where C NO x in is the inlet concentration (ppm) of NO and NO2, while C NO x   out and   C N 2 out correspond to the outlet concentration (ppm) of NO, NO2 and N2, respectively.

3.3. Catalyst Characterization

XRD patterns were recorded by a D8 Advance diffractometer (Bruker, Karlsruhe, Germany) with Cu-Kα radiation ( λ = 1.5418 Å). N2 adsorption−desorption experiments were performed by an ASAP 2460 analyzer (Micromeritics, Norcross, GA, USA) at liquid N2 temperature (−196 °C). Prior to the measurements, the samples were degassed in a vacuum at 300 °C for 4 h. The specific surface area was calculated using the BET method at P/P0 = 0.01–0.3. The total pore volume was determined from the desorption branch isotherms by the Barrett–Joyner–Halenda (BJH) method and the average pore size was determined from the desorption average pore diameter (4V/A by BET). The morphology of samples was characterized by field-emission SEM with an instrument of ZEISS Merlin (Carl Zeiss AG, Jena, Germany). TEM and HR-TEM were performed on a JEM-2100F equipped with an EDS analysis to further obtain microstructure information of the samples. XPS with Al-Kα radiation (hυ = 1253.6 eV, Axis Ultra DLD, Kratos, UK) was conducted to analyze the composition and valences of elements on the surface of all catalysts. Both H2-TPR and NH3-TPD analyses were conducted by an Auto ChemII 2920 instrument (Micromeritics, Norcross, GA, USA).
In situ DRIFTS experiments were performed on a Fourier transform infrared spectrometer (FTIR, Nicolet iS50, Thermo Scientific, Waltham, MA, USA) equipped with an MCT/A detector and a Harrick DRIFT chamber in the range of 600–4800 cm−1 at a resolution of 4 cm−1. Before the single-beam background collection, the catalyst was purged in Ar gas at 300 °C for 1 h and then cooled to 180 °C for the aimed tests. A total flow of 100 mL min−1 of the gas stream was balanced with Ar, and the test conditions of the adsorption and the reaction are described in detail in the results section. All the DRIFTS spectra were recorded through subtraction of the corresponding background reference by collecting 64 scans.

4. Conclusions

In summary, several GdMn/Z-x catalysts were synthesized by a citric acid–ethanol dispersion method and used for the NH3-SCR of NO removal in the presence of SO2. Among all the catalysts, the activity of NO conversion followed the order of GdMn/Z-0.3 > GdMn/Z-0.6 > Mn/Z > GdMn/Z-0.1. The results exhibited that the addition of a proper amount of Gd (Gd/Mn = 0.3) into the Mn/Z catalyst could remarkably enhance its catalytic activity and SO2 resistance for NH3-SCR. The excellent performance of GdMn/Z-0.3 could be attributed to the existence of a large surface area, abundant Mn4+ and chemisorbed oxygen species, strong reducibility and suitable surface acidity. In addition, the Mn/Z catalyst showed extremely poor SO2 resistance due to its strong SO2 adsorption ability. A large number of stable sulfate species, especially bulk manganese sulfate, were formed on the Mn/Z surface, which reduced the NO adsorption ability and covered the active sites. However, Gd significantly inhibited SO2 adsorption on Mn/Z, and the bulk manganese sulfate was not observed in the DRIFT spectra of SO2 adsorption, thus leading to excellent SO2 resistance over GdMn/Z-0.3. In situ DRIFTS experiment results revealed that the reaction mechanism of low-temperature SCR over Mn/Z and GdMn/Z-0.3 followed the L–H and E–R mechanism simultaneously, while the reaction path over Mn/Z was greatly affected by SO2, resulting in the poor activity. Fortunately, the reaction path on GdMn/Z-0.3 was almost unaffected by SO2, thus possessing excellent SO2 resistance. Consequently, GdMn/Z-0.3 as a high-effective potential catalyst can be applied in the remediation of harmful NOx pollutants.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/11/3/324/s1; Figure S1: The long-term stability test of GdMn/Z-0.3 catalyst; Figure S2: SEM images of the (a,b) Mn/Z, (c,d) GdMn/Z-0.1, (e,f) GdMn/Z-0.3 and (g,h) GdMn/Z-0.6 catalysts; Figure S3: The SO2 and H2O resistance tests over the Mn/Z and GdMn/Z-0.3 catalysts; Figure S4: XRD patterns of fresh and used Mn/Z and GdMn/Z-0.3 catalysts; Figure S5: The single beam background of Mn/Z and GdMn/Z-0.3 catalysts and (b) the spectra of the catalysts prior to any SCR reactants dosage.

Author Contributions

Funding acquisition, B.H.; Investigation, J.G.; Writing—original draft, J.G.; Writing—review and editing, J.G., L.Z., W.L., D.H., J.W. and B.H. 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, grant number NSFC-51478191 and the Key Research and Development Plan of Guangdong Province, grant number 2019B110207001.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kwak, J.H.; Tonkyn, R.G.; Kim, D.H.; Szanyi, J.; Peden, C.H.F. Excellent activity and selectivity of Cu-SSZ-13 in the selective catalytic reduction of NOx with NH3. J. Catal. 2010, 275, 187–190. [Google Scholar] [CrossRef]
  2. Jiang, H.; Wang, Q.; Wang, H.; Chen, Y.; Zhang, M. Temperature effect on the morphology and catalytic performance of Co-MOF-74 in low-temperature NH3-SCR process. Catal. Commun. 2016, 80, 24–27. [Google Scholar] [CrossRef]
  3. 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]
  4. Gholami, F.; Tomas, M.; Gholami, Z.; Vakili, M. Technologies for the nitrogen oxides reduction from flue gas: A review. Sci. Total Environ. 2020, 714. [Google Scholar] [CrossRef] [PubMed]
  5. Heck, R.M. Catalytic abatement of nitrogen oxides-stationary applications. Catal. Today 1999, 53, 519–523. [Google Scholar] [CrossRef]
  6. Roy, S.; Baiker, A. NOx storage-reduction catalysis: From mechanism and materials properties to storage-reduction performance. Chem. Rev. 2009, 109, 4054–4091. [Google Scholar] [CrossRef] [PubMed]
  7. Fang, D.; Xie, J.; Hu, H.; Yang, H.; He, F.; Fu, Z. Identification of MnOx species and Mn valence states in MnOx/TiO2 catalysts for low temperature SCR. Chem. Eng. J. 2015, 271, 23–30. [Google Scholar] [CrossRef]
  8. Yu, C.; Huang, B.; Dong, L.; Chen, F.; Liu, X. In situ FT-IR study of highly dispersed MnOx/SAPO-34 catalyst for low-temperature selective catalytic reduction of NOx by NH3. Catal. Today 2017, 281, 610–620. [Google Scholar] [CrossRef]
  9. Lou, X.; Liu, P.; Li, J.; Li, Z.; He, K. Effects of calcination temperature on Mn species and catalytic activities of Mn/ZSM-5 catalyst for selective catalytic reduction of NO with ammonia. Appl. Surf. Sci. 2014, 307, 382–387. [Google Scholar] [CrossRef]
  10. Xie, J.L.; Fang, D.; He, F.; Chen, J.F.; Fu, Z.B.; Chen, X.L. Performance and mechanism about MnOx species included in MnOx/TiO2 catalysts for SCR at low temperature. Catal. Commun. 2012, 28, 77–81. [Google Scholar] [CrossRef]
  11. Wei, L.; Cui, S.P.; Guo, H.X.; Ma, X.Y.; Wan, Y.Q.; Yu, S.S. The mechanism of the deactivation of MnOx/TiO2 catalyst for low-temperature SCR of NO. Appl. Surf. Sci. 2019, 483, 391–398. [Google Scholar] [CrossRef]
  12. Kijlstra, W.S.; Biervliet, M.; Poels, E.K.; Bliek, A. Deactivation by SO2 of MnOx/Al2O3 catalysts used for the selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 1998, 16, 327–337. [Google Scholar] [CrossRef]
  13. Su, Y.X.; Fan, B.X.; Wang, L.S.; Liu, Y.F.; Huang, B.C.; Fu, M.L.; Chen, L.M.; Ye, D.Q. MnOx supported on carbon nanotubes by different methods for the SCR of NO with NH3. Catal. Today 2013, 201, 115–121. [Google Scholar] [CrossRef]
  14. Carja, G.; Kameshima, Y.; Okada, K.; Madhusoodana, C.D. Mn–Ce/ZSM5 as a new superior catalyst for NO reduction with NH3. Appl. Catal. B Environ. 2007, 73, 60–64. [Google Scholar] [CrossRef]
  15. Lv, G.; Bin, F.; Song, C.L.; Wang, K.P.; Song, J.N. Promoting effect of zirconium doping on Mn/ZSM-5 for the selective catalytic reduction of NO with NH3. Fuel 2013, 107, 217–224. [Google Scholar] [CrossRef]
  16. Xu, W.; Zhang, G.; Chen, H.; Zhang, G.; Han, Y.; Chang, Y.; Gong, P. Mn/beta and Mn/ZSM-5 for the low-temperature selective catalytic reduction of NO with ammonia: Effect of manganese precursors. Chin. J. Catal. 2018, 39, 118–127. [Google Scholar] [CrossRef]
  17. Chen, P.; Jabłońska, M.; Weide, P.; Caumanns, T.; Weirich, T.; Muhler, M.; Moos, R.; Palkovits, R.; Simon, U. Formation and Effect of NH4+ Intermediates in NH3–SCR over Fe-ZSM-5 Zeolite Catalysts. ACS Catal. 2016, 6, 7696–7700. [Google Scholar] [CrossRef]
  18. Shi, J.; Zhang, Y.; Fan, Z.; Chen, M.; Zhang, Z.; Shangguan, W. Widened Active Temperature Window of a Fe-ZSM-5 Catalyst by an Impregnation Solvent for NH3-SCR of NO. Ind. Eng. Chem. Res. 2018, 57, 13703–13712. [Google Scholar] [CrossRef]
  19. Kim, Y.J.; Kwon, H.J.; Heo, I.; Nam, I.-S.; Cho, B.K.; Choung, J.W.; Cha, M.-S.; Yeo, G.K. Mn–Fe/ZSM5 as a low-temperature SCR catalyst to remove NOx from diesel engine exhaust. Appl. Catal. B Environ. 2012, 126, 9–21. [Google Scholar] [CrossRef]
  20. 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] [PubMed]
  21. Ilieva, L.; Pantaleo, G.; Ivanov, I.; Nedyalkova, R.; Venezia, A.M.; Andreeva, D. NO reduction by CO over gold based on ceria, doped by rare earth metals. Catal. Today 2008, 139, 168–173. [Google Scholar] [CrossRef]
  22. 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] [PubMed]
  23. Jin, R.; Liu, Y.; Wang, Y.; Cen, W.; Wu, Z.; Wang, H.; Weng, X. 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]
  24. Meng, D.; Zhan, W.; Guo, Y.; Guo, Y.; Wang, L.; Lu, G. A Highly Effective Catalyst of Sm-MnOx for the NH3-SCR of NOx at Low Temperature: Promotional Role of Sm and Its Catalytic Performance. ACS Catal. 2015, 5, 5973–5983. [Google Scholar] [CrossRef]
  25. You, X.C.; Sheng, Z.Y.; Yu, D.Q.; Yang, L.; Xiao, X.; Wang, S. Influence of Mn/Ce ratio on the physicochemical properties and catalytic performance of graphene supported MnOx-CeO2 oxides for NH3-SCR at low temperature. Appl. Surf. Sci. 2017, 423, 845–854. [Google Scholar] [CrossRef]
  26. 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]
  27. Yu, C.; Huang, B.; Dong, L.; Chen, F.; Yang, Y.; Fan, Y.; Yang, Y.; Liu, X.; Wang, X. Effect of Pr/Ce addition on the catalytic performance and SO2 resistance of highly dispersed MnOx/SAPO-34 catalyst for NH3-SCR at low temperature. Chem. Eng. J. 2017, 316, 1059–1068. [Google Scholar] [CrossRef]
  28. Liu, C.; Shi, J.-W.; Gao, C.; Niu, C. Manganese oxide-based catalysts for low-temperature selective catalytic reduction of NOx with NH3: A review. Appl. Catal. A Gen. 2016, 522, 54–69. [Google Scholar] [CrossRef]
  29. Panahi, P.N.; Salari, D.; Niaei, A.; Mousavi, S.M. NO reduction over nanostructure M-Cu/ZSM-5 (M: Cr, Mn, Co and Fe) bimetallic catalysts and optimization of catalyst preparation by RSM. J. Ind. Eng. Chem. 2013, 19, 1793–1799. [Google Scholar] [CrossRef]
  30. Zhao, X.; Mao, L.; Dong, G. Mn-Ce-V-WOx/TiO2 SCR Catalysts: Catalytic Activity, Stability and Interaction among Catalytic Oxides. Catalysts 2018, 8, 76. [Google Scholar] [CrossRef] [Green Version]
  31. Fan, Z.; Shi, J.-W.; Gao, C.; Gao, G.; Wang, B.; Wang, Y.; He, C.; Niu, C. 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]
  32. Chen, H.F.; Xia, Y.; Fang, R.Y.; Huang, H.; Gan, Y.P.; Liang, C.; Zhang, J.; Zhang, W.K.; Liu, X.S. Effects of Nd-modification on the activity and SO2 resistance of MnOx/TiO2 catalysts for low-temperature NH3-SCR. New J. Chem. 2018, 42, 12845–12852. [Google Scholar] [CrossRef]
  33. Yang, G.; Zhao, H.T.; Luo, X.; Shi, K.Q.; Zhao, H.B.; Wang, W.K.; Chen, Q.H.; Fan, H.; Wu, T. Promotion effect and mechanism of the addition of Mo on the enhanced low temperature SCR of NOx by NH3 over MnOx/gamma-Al2O3 catalysts. Appl. Catal. B Environ. 2019, 245, 743–752. [Google Scholar] [CrossRef] [Green Version]
  34. Liu, Z.; Yi, Y.; Zhang, S.; Zhu, T.; Zhu, J.; Wang, J. Selective catalytic reduction of NOx with NH3 over Mn-Ce mixed oxide catalyst at low temperatures. Catal. Today 2013, 216, 76–81. [Google Scholar] [CrossRef]
  35. Tang, W.; Deng, Y.; Li, W.; Li, S.; Wu, X.; Chen, Y. Restrictive nanoreactor for growth of transition metal oxides (MnO2, Co3O4, NiO) nanocrystal with enhanced catalytic oxidation activity. Catal. Commun. 2015, 72, 165–169. [Google Scholar] [CrossRef]
  36. Wang, Y.; Li, D.; Li, K.; Farrauto, R. Enhanced propane and carbon monoxide oxidation activity by structural interactions of CeO2 with MnOx/Nb2O5-x catalysts. Appl. Catal. B Environ. 2020, 267. [Google Scholar] [CrossRef]
  37. Wang, F.; Xiao, M.; Ma, X.; Wu, S.; Ge, M.; Yu, X. Insights into the transformations of Mn species for peroxymonosulfate activation by tuning the Mn3O4 shapes. Chem. Eng. J. 2021, 404. [Google Scholar] [CrossRef]
  38. Yang, S.; Wang, C.; Li, J.; Yan, N.; Ma, L.; Chang, H. Low temperature selective catalytic reduction of NO with NH3 over Mn-Fe spinel: Performance, mechanism and kinetic study. Appl. Catal. B Environ. 2011, 110, 71–80. [Google Scholar] [CrossRef]
  39. Zhang, X.; Wu, Q.; Diao, Q.; Wang, J.; Xiao, K.; Yang, B.; Wu, X. Performance study for NH3-SCR at low temperature based on different methods of Mnx/SEP catalyst. Chem. Eng. J. 2019, 370, 364–371. [Google Scholar] [CrossRef]
  40. Wang, T.; Wan, Z.; Yang, X.; Zhang, X.; Niu, X.; Sun, B. Promotional effect of iron modification on the catalytic properties of Mn-Fe/ZSM-5 catalysts in the Fast SCR reaction. Fuel Process. Technol. 2018, 169, 112–121. [Google Scholar] [CrossRef]
  41. Chang, H.Z.; Li, J.H.; Yuan, J.; Chen, L.; Dai, Y.; Arandiyan, H.; Xu, J.Y.; Hao, J.M. Ge, Mn-doped CeO2-WO3 catalysts for NH3-SCR of NOx: Effects of SO2 and H-2 regeneration. Catal. Today 2013, 201, 139–144. [Google Scholar] [CrossRef]
  42. Meng, D.; Zhan, W.; Guo, Y.; Guo, Y.; Wang, Y.; Wang, L.; Lu, G. 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 Chem. 2016, 420, 272–281. [Google Scholar] [CrossRef]
  43. Gao, C.; Shi, J.-W.; Fan, Z.; Yu, Y.; Chen, J.; Li, Z.; Niu, C. Eu-Mn-Ti mixed oxides for the SCR of NOx with NH3: The effects of Eu-modification on catalytic performance and mechanism. Fuel Process. Technol. 2017, 167, 322–333. [Google Scholar] [CrossRef]
  44. Liu, D.J.; Zhou, W.G.; Wu, J. Effect of Ce and La on the activity of CuO/ZSM-5 and MnOx/ZSM-5 composites for elemental mercury removal at low temperature. Fuel 2017, 194, 115–122. [Google Scholar] [CrossRef]
  45. Zhang, D.; Zhang, L.; Shi, L.; Fang, C.; Li, H.; Gao, R.; Huang, L.; Zhang, J. In situ supported MnOx-CeOx on carbon nanotubes for the low-temperature selective catalytic reduction of NO with NH3. Nanoscale 2013, 5, 1127–1136. [Google Scholar] [CrossRef]
  46. Shao, J.; Cheng, S.; Li, Z.; Huang, B. Enhanced Catalytic Performance of Hierarchical MnOx/ZSM-5 Catalyst for the Low-Temperature NH3-SCR. Catalysts 2020, 10, 311. [Google Scholar] [CrossRef] [Green Version]
  47. Zhou, G.Y.; Zhong, B.C.; Wang, W.H.; Guan, X.J.; Huang, B.C.; Ye, D.Q.; Wu, H.J. In situ DRIFTS study of NO reduction by NH3 over Fe-Ce-Mn/ZSM-5 catalysts. Catal. Today 2011, 175, 157–163. [Google Scholar] [CrossRef]
  48. Wang, F.; Zhu, Y.; Li, Z.; Shan, Y.; Shan, W.; Shi, X.; Yu, Y.; Zhang, C.; Li, K.; Ning, P.; et al. Promoting effect of acid sites on NH3-SCO activity with water vapor participation for Pt-Fe/ZSM-5 catalyst. Catal. Today 2020. [Google Scholar] [CrossRef]
  49. Long, R.Q.; Yang, R.T. Reaction Mechanism of Selective Catalytic Reduction of NO with NH3 over Fe-ZSM-5 Catalyst. J. Catal. 2002, 207, 224–231. [Google Scholar] [CrossRef]
  50. Shu, Y.; Sun, H.; Quan, X.; Chen, S. Enhancement of Catalytic Activity over the Iron-Modified Ce/TiO2 Catalyst for Selective Catalytic Reduction of NOx with Ammonia. J. Phys. Chem. C 2012, 116, 25319–25327. [Google Scholar] [CrossRef]
  51. Sun, C.; Chen, W.; Jia, X.; Liu, A.; Gao, F.; Feng, S.; Dong, L. Comprehensive understanding of the superior performance of Sm-modified Fe2O3 catalysts with regard to NO conversion and H2O/SO2 resistance in the NH3-SCR reaction. Chin. J. Catal. 2021, 42, 417–430. [Google Scholar] [CrossRef]
  52. Mu, J.; Li, X.; Sun, W.; Fan, S.; Wang, X.; Wang, L.; Qin, M.; Gan, G.; Yin, Z.; Zhang, D. Inductive Effect Boosting Catalytic Performance of Advanced Fe1–xVxOδ Catalysts in Low-Temperature NH3 Selective Catalytic Reduction: Insight into the Structure, Interaction, and Mechanisms. ACS Catal. 2018, 8, 6760–6774. [Google Scholar] [CrossRef]
  53. Chen, Y.; Zhang, Z.; Liu, L.; Mi, L.; Wang, X. In situ DRIFTS studies on MnO nanowires supported by activated semi-coke for low temperature selective catalytic reduction of NO with NH3. Appl. Surf. Sci. 2016, 366, 139–147. [Google Scholar] [CrossRef]
  54. Kijlstra, W.S.; Brands, D.S.; Poels, E.K.; Bliek, A. Mechanism of the selective catalytic reduction of NO by NH3 over MnOx/Al2O3. 1. Adsorption and desorption of the single reaction components. J. Catal. 1997, 171, 208–218. [Google Scholar] [CrossRef]
  55. Fei, Z.; Yang, Y.; Wang, M.; Tao, Z.; Liu, Q.; Chen, X.; Cui, M.; Zhang, Z.; Tang, J.; Qiao, X. Precisely fabricating Ce-O-Ti structure to enhance performance of Ce-Ti based catalysts for selective catalytic reduction of NO with NH3. Chem. Eng. J. 2018, 353, 930–939. [Google Scholar] [CrossRef]
  56. Wang, Y.; Shi, X.; Shan, Y.; Du, J.; Liu, K.; He, H. Hydrothermal Stability Enhancement of Al-Rich Cu-SSZ-13 for NH3 Selective Catalytic Reduction Reaction by Ion Exchange with Cerium and Samarium. Ind. Eng. Chem. Res. 2020, 59, 6416–6423. [Google Scholar] [CrossRef]
  57. Dong, L.; Fan, Y.; Ling, W.; Yang, C.; Huang, B. Effect of Ce/Y Addition on Low-Temperature SCR Activity and SO2 and H2O Resistance of MnOx/ZrO2/MWCNTs Catalysts. Catalysts 2017, 7, 181. [Google Scholar] [CrossRef]
  58. Jiang, B.Q.; Deng, B.Y.; Zhang, Z.Q.; Wu, Z.L.; 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 2014, 118, 14866–14875. [Google Scholar] [CrossRef]
  59. Yi, H.H.; Ma, C.B.; Tang, X.L.; Zhao, S.Z.; Yang, K.; Sani, Z.; Song, L.L.; Zhang, X.D.; Han, W. Synthesis of MgO@CeO2-MnOx core shell structural adsorbent and its application in reducing the competitive adsorption of SO2 and NOx in coal-fired flue gas. Chem. Eng. J. 2019, 372, 129–140. [Google Scholar] [CrossRef]
  60. Wei, L.; Cui, S.P.; Guo, H.X.; Ma, X.Y.; Zhang, L.J. DRIFT and DFT study of cerium addition on SO2 of Manganese-based Catalysts for low temperature SCR. J. Mol. Catal. A Chem. 2016, 421, 102–108. [Google Scholar] [CrossRef]
  61. Han, L.; Gao, M.; Hasegawa, J.Y.; Li, S.; Shen, Y.; Li, H.; Shi, L.; Zhang, D. SO2-Tolerant Selective Catalytic Reduction of NOx over Meso-TiO2@Fe2O3@Al2O3 Metal-Based Monolith Catalysts. Environ. Sci. Technol. 2019, 53, 6462–6473. [Google Scholar] [CrossRef] [PubMed]
  62. Tang, X.; Shi, Y.; Gao, F.; Zhao, S.; Yi, H.; Xie, Z. Promotional role of Mo on Ce0.3FeOx catalyst towards enhanced NH3-SCR catalytic performance and SO2 resistance. Chem. Eng. J. 2020, 398. [Google Scholar] [CrossRef]
  63. Wang, F.M.; Shen, B.X.; Zhu, S.W.; Wang, Z. Promotion of Fe and Co doped Mn-Ce/TiO2 catalysts for low temperature NH3-SCR with SO2 tolerance. Fuel 2019, 249, 54–60. [Google Scholar] [CrossRef]
Figure 1. The NH3-SCR performance over GdMn/Z-x catalysts (x = 0, 0.1, 0.3 and 0.6): (a) NO conversion and (b) N2 selectivity; with the reaction conditions: 800 ppm NH3, 800 ppm NO, 5 vol.% O2, Ar to balance and gas hourly space velocity (GHSV) = 40,000 h−1.
Figure 1. The NH3-SCR performance over GdMn/Z-x catalysts (x = 0, 0.1, 0.3 and 0.6): (a) NO conversion and (b) N2 selectivity; with the reaction conditions: 800 ppm NH3, 800 ppm NO, 5 vol.% O2, Ar to balance and gas hourly space velocity (GHSV) = 40,000 h−1.
Catalysts 11 00324 g001
Figure 2. XRD patterns of the pure ZSM-5 and the catalysts.
Figure 2. XRD patterns of the pure ZSM-5 and the catalysts.
Catalysts 11 00324 g002
Figure 3. TEM and HR-TEM images of (ac) Mn/Z and (df) GdMn/Z-0.3 catalysts, (g) high-angle annular dark field scanning TEM image and (hk) EDS elemental mappings of GdMn/Z-0.3 catalyst.
Figure 3. TEM and HR-TEM images of (ac) Mn/Z and (df) GdMn/Z-0.3 catalysts, (g) high-angle annular dark field scanning TEM image and (hk) EDS elemental mappings of GdMn/Z-0.3 catalyst.
Catalysts 11 00324 g003
Figure 4. XPS spectra of (a) Mn 2p and (b) O 1s of the catalysts.
Figure 4. XPS spectra of (a) Mn 2p and (b) O 1s of the catalysts.
Catalysts 11 00324 g004
Figure 5. (a) H2-TPR profiles and (b) NH3-TPD profiles of the catalysts.
Figure 5. (a) H2-TPR profiles and (b) NH3-TPD profiles of the catalysts.
Catalysts 11 00324 g005
Figure 6. In situ DRIFT spectra of (a) Mn/Z and (b) GdMn/Z-0.3, collected as a function of time and exposed to NH3 at 180 °C.
Figure 6. In situ DRIFT spectra of (a) Mn/Z and (b) GdMn/Z-0.3, collected as a function of time and exposed to NH3 at 180 °C.
Catalysts 11 00324 g006
Figure 7. In situ DRIFT spectra of (a) Mn/Z and (b) GdMn/Z-0.3, collected as a function of time and exposed to NO and O2 at 180 °C.
Figure 7. In situ DRIFT spectra of (a) Mn/Z and (b) GdMn/Z-0.3, collected as a function of time and exposed to NO and O2 at 180 °C.
Catalysts 11 00324 g007
Figure 8. In situ DRIFT spectra of the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts pretreated by exposure to NH3 for 30 min followed by exposure to NO and O2 at 180 °C for various minutes.
Figure 8. In situ DRIFT spectra of the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts pretreated by exposure to NH3 for 30 min followed by exposure to NO and O2 at 180 °C for various minutes.
Catalysts 11 00324 g008
Figure 9. In situ DRIFT spectra of Mn/Z (a) and GdMn/Z-0.3 (b) catalysts pretreated by exposure to NO and O2 for 30 min, followed by exposure to NH3 at 180 °C for various minutes.
Figure 9. In situ DRIFT spectra of Mn/Z (a) and GdMn/Z-0.3 (b) catalysts pretreated by exposure to NO and O2 for 30 min, followed by exposure to NH3 at 180 °C for various minutes.
Catalysts 11 00324 g009
Figure 10. In situ DRIFT spectra of the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts exposed to NH3, NO and O2 at 180 °C for various minutes.
Figure 10. In situ DRIFT spectra of the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts exposed to NH3, NO and O2 at 180 °C for various minutes.
Catalysts 11 00324 g010
Figure 11. The SO2 resistance tests over Mn/Z and GdMn/Z-0.3; reaction conditions: 800 ppm NH3, 800 ppm NO, 5 vol.% O2, 100 ppm SO2, Ar to balance and GHSV = 40,000 h−1. (For the initial 100% NO conversion, the reaction temperature was 180 °C. For the 80% NO conversion, the reaction temperatures were 110 °C and 100 °C for Mn/Z and GdMn/Z, respectively).
Figure 11. The SO2 resistance tests over Mn/Z and GdMn/Z-0.3; reaction conditions: 800 ppm NH3, 800 ppm NO, 5 vol.% O2, 100 ppm SO2, Ar to balance and GHSV = 40,000 h−1. (For the initial 100% NO conversion, the reaction temperature was 180 °C. For the 80% NO conversion, the reaction temperatures were 110 °C and 100 °C for Mn/Z and GdMn/Z, respectively).
Catalysts 11 00324 g011
Figure 12. SEM images of used (a,b) Mn/Z-S and (c,d) GdMn/Z-0.3-S catalysts.
Figure 12. SEM images of used (a,b) Mn/Z-S and (c,d) GdMn/Z-0.3-S catalysts.
Catalysts 11 00324 g012
Figure 13. XPS spectra of Mn/Z-S and GdMn/Z-S: (a) Mn 2p and (b) O 1s.
Figure 13. XPS spectra of Mn/Z-S and GdMn/Z-S: (a) Mn 2p and (b) O 1s.
Catalysts 11 00324 g013
Figure 14. In situ DRIFT spectra of SO2 and O2 adsorption on the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts at 180 °C.
Figure 14. In situ DRIFT spectra of SO2 and O2 adsorption on the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts at 180 °C.
Catalysts 11 00324 g014
Figure 15. Dynamic changes of in situ DRIFT spectra over the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts as a function of time in a flow of NH3, NO and O2 after the catalysts were pre-exposed to a flow of SO2 for 30 min, followed by Ar purging for 30 min.
Figure 15. Dynamic changes of in situ DRIFT spectra over the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts as a function of time in a flow of NH3, NO and O2 after the catalysts were pre-exposed to a flow of SO2 for 30 min, followed by Ar purging for 30 min.
Catalysts 11 00324 g015
Figure 16. Dynamic changes of in situ DRIFT spectra over the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts as a function of time in a flow of SO2 after the catalysts were pre-exposed to a flow of NH3, NO and O2 for 30 min, followed by Ar purging for 30 min.
Figure 16. Dynamic changes of in situ DRIFT spectra over the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts as a function of time in a flow of SO2 after the catalysts were pre-exposed to a flow of NH3, NO and O2 for 30 min, followed by Ar purging for 30 min.
Catalysts 11 00324 g016
Figure 17. Dynamic changes of in situ DRIFT spectra over the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts as a function of time under an atmosphere of SO2, NH3, NO and O2.
Figure 17. Dynamic changes of in situ DRIFT spectra over the (a) Mn/Z and (b) GdMn/Z-0.3 catalysts as a function of time under an atmosphere of SO2, NH3, NO and O2.
Catalysts 11 00324 g017
Table 1. Texture properties of the catalysts.
Table 1. Texture properties of the catalysts.
CatalystsSpecific Surface Area
(m2g−1)
Total Pore Volume
(cm3g−1)
Average Pore Size
(nm)
Mn/Z266.80.112.98
GdMn/Z-0.1241.60.203.27
GdMn/Z-0.3273.40.253.47
GdMn/Z-0.6259.20.193.40
Table 2. XPS results for the surface atomic composition and concentrations of the catalysts.
Table 2. XPS results for the surface atomic composition and concentrations of the catalysts.
SamplesSurface Atomic Concentration (%)Relative Concentration Ratios (%)
MnOGdothers aMn4+Mn3+Mn2+OSOL
Mn/Z5.0977.1917.7228.535.735.872.827.2
GdMn/Z-0.14.4764.591.2429.7027.234.638.171.928.1
GdMn/Z-0.34.9066.463.2225.4234.142.223.774.425.6
GdMn/Z-0.64.0860.975.1229.8331.936.831.369.830.2
Other a elements include Si, Al and C.
Table 3. H2-TPR and NH3-TPD data of the catalysts.
Table 3. H2-TPR and NH3-TPD data of the catalysts.
CatalystsWeak Acid
(mmol·g−1)
Medium-Strong Acid
(mmol·g−1)
Total Acid Amount
(mmol·g−1)
H2-TPR
Curve Area
Mn/Z0.6050.3040.9090.31435
GdMn/Z-0.10.7890.4651.2540.25803
GdMn/Z-0.30.7020.5421.2440.69167
GdMn/Z-0.60.6380.5881.2260.40507
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guan, J.; Zhou, L.; Li, W.; Hu, D.; Wen, J.; Huang, B. Improving the Performance of Gd Addition on Catalytic Activity and SO2 Resistance over MnOx/ZSM-5 Catalysts for Low-Temperature NH3-SCR. Catalysts 2021, 11, 324. https://doi.org/10.3390/catal11030324

AMA Style

Guan J, Zhou L, Li W, Hu D, Wen J, Huang B. Improving the Performance of Gd Addition on Catalytic Activity and SO2 Resistance over MnOx/ZSM-5 Catalysts for Low-Temperature NH3-SCR. Catalysts. 2021; 11(3):324. https://doi.org/10.3390/catal11030324

Chicago/Turabian Style

Guan, Jinkun, Lusha Zhou, Weiquan Li, Die Hu, Jie Wen, and Bichun Huang. 2021. "Improving the Performance of Gd Addition on Catalytic Activity and SO2 Resistance over MnOx/ZSM-5 Catalysts for Low-Temperature NH3-SCR" Catalysts 11, no. 3: 324. https://doi.org/10.3390/catal11030324

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