Next Article in Journal
Optimization of an Open-Cell Foam-Based Ni-Mg-Al Catalyst for Enhanced CO2 Hydrogenation to Methane
Next Article in Special Issue
Steam Treatment Promotion on the Performance of Pt/CeO2 Three-Way Catalysts for Emission Control of Natural Gas-Fueled Vehicles
Previous Article in Journal
Dimension Engineering in Noble-Metal-Based Nanocatalysts
Previous Article in Special Issue
New Insights into the Effect of Ce Doping on the Catalytic Performance and Hydrothermal Stability of Cu-USY Zeolite Catalysts for the Selective Catalytic Reduction of NO with NH3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Low-Temperature Activity and Hydrothermal Stability of Ce-Mn Oxide-Modified Cu-SSZ-39 Catalysts for NH3-SCR of NOx

1
School of Chemistry and Chemical Engineering, Shandong Provincial Key Laboratory of Fluorine Chemistry and Chemical Materials, University of Jinan, Jinan 250022, China
2
National Engineering Laboratory for Mobile Source Emission Control Technology, China Automotive Technology & Research Center Co., Ltd., Tianjin 300300, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(1), 10; https://doi.org/10.3390/catal14010010
Submission received: 30 November 2023 / Revised: 18 December 2023 / Accepted: 19 December 2023 / Published: 21 December 2023

Abstract

:
Cu-SSZ-39 zeolite with an AEI structure exhibits excellent hydrothermal stability and can be a potential alternative to Cu-SSZ-13 zeolite SCR catalysts for NOx removal in diesel vehicles. However, the inferior low-temperature performance of Cu-SSZ-39 leads to substantial NOx emissions during the cold-start period, impeding its practical application. In this study, Ce-Mn oxide-modified Cu-SSZ-39 catalysts (CeMnOx/Cu-SSZ-39) and references (CeO2/Cu-SSZ-39 and MnOx/Cu-SSZ-39) were prepared by the ion-exchange of Cu ions followed by impregnation of the oxide precursors, with the aim of enhancing the NH3-SCR performance at low temperatures. The modified catalysts exhibited improved low-temperature activity and hydrothermal stability compared to the unmodified counterpart. In particular, CeMnOx/Cu-SSZ-39 showed the highest activity among the three catalysts and achieved NOx conversions above 90% within the temperature range of 180 °C to 600 °C, even after undergoing hydrothermal aging at 800 °C. Experimental results indicated that the synergistic effect between Ce and Mn in CeMnOx improves the redox properties and acidity of the catalyst due to the presence of Ce3+, Mn4+, and abundant adsorbed oxygen species, which facilitate low-temperature SCR reactions. Furthermore, the interaction of CeMnOx with Cu-SSZ-39 stabilizes the zeolite framework and hinders the agglomeration of Cu species during the hydrothermal aging process, contributing to its exceptional hydrothermal stability. The kinetics and NO oxidation experiments demonstrated that CeMnOx provides access to fast SCR reaction pathways by oxidizing NO to NO2, resulting in a significant increase in low-temperature activity. This study provides novel guidelines for the design and preparation of Cu-SSZ-39 zeolite with outstanding SCR performance over a wide temperature range.

Graphical Abstract

1. Introduction

Nitrogen oxides (NOx), emitted in particular by mobile sources, are one of the major precursors of urban air pollution, such as haze and ozone [1]. Currently, selective catalytic reduction (SCR) of NOx with ammonia (NH3) has been proven to be one of the most efficient technologies for NOx purification [2]. Cu-based small-pore zeolite SCR catalysts (especially Cu-SSZ-13 zeolite with a CHA structure) have been widely adopted in diesel engines owing to their wide active temperature window and hydrothermal stability [3,4,5]. In practical applications, the periodically increased exhaust temperature required to stimulate diesel particulate filter (DPF) regeneration necessitates SCR catalysts with high hydrothermal stability [6]. Nevertheless, the Cu-SSZ-13 catalyst exhibits a considerable decrease in deNOx activity after hydrothermal aging at 800 °C, particularly for Al-rich Cu-SSZ-13 catalysts [7,8]. In recent years, many other small-pore zeolites with variable stable framework structures have attracted extensive attention in an attempt to increase hydrothermal stability of the catalysts, including LTA [9,10], KFI [11], AFX [12,13], RTH [14], and AEI [15,16]. In view of its catalytic performance and the availability of synthetic precursors, Cu-SSZ-39 zeolite with an AEI structure is considered to be an excellent substitute for Cu-SSZ-13, possessing a bright future in industrial applications [17,18,19,20,21]. However, Cu-SSZ-39 catalysts frequently exhibit insufficient activity at low exhaust temperatures, leading to a critical problem: up to 80% of NOx emissions occur during the cold-start period before the SCR catalysts achieve the light-off temperature [5]. Therefore, optimizing the SCR performance (especially the low-temperature activity) of Cu-SSZ-39 catalysts is urgently needed.
Recently, more and more researchers have concentrated on enhancing the SCR performance of Cu-SSZ-39 catalysts, and two strategies have been proposed. One is modulating Cu2+ ions, which is the fundamental approach to optimize the SCR performance since the Cu2+ ions are the active sites of Cu-SSZ-39 [22]. Fu et al. increased the Cu content in Cu-SSZ-39, which led to a significant improvement in low-temperature SCR activity by endowing the catalyst with a large amount of highly stable, isolated Cu2+ ions [23]. Additionally, the framework tetrahedral Al atoms were stabilized by the Cu2+ ions located on the face of the double six-membered ring (d6r) units of the AEI framework, thereby enhancing their hydrothermal stability. Du and colleagues regulated Cu2+ ions by varying the precursors during the synthesis of Na-SSZ-39, and the ZSM-5 precursor provided more Cu(OH)+ species as well as a higher Si/Al ratio in the obtained Cu-SSZ-39, which guaranteed superior low-temperature activity and hydrothermal stability in comparison to the three other catalysts prepared using Y zeolite, Beta zeolite, colloidal silica, and sodium aluminate precursors, respectively [24]. Distinctively, Du et al. discovered that a hydrothermal aging treatment (HTA) allowed some of the Cu2+ ions near d6r to transform into CuO species, which promoted the production of crucial nitrate intermediates in SCR reactions, resulting in boosted low-temperature activity [18]. Similarly, Chen et al. observed a comparable phenomenon in the phosphorus-poisoned Cu-SSZ-39 catalyst, where some of the deactivated Cu2+ ions were recovered due to the partial decomposition of Cu-P species after an HTA [19].
In addition to modulating Cu2+ ions, the introduction of heteroatoms is another effective way to improve the SCR performance of Cu-SSZ-39 catalysts [25,26]. Transition and rare-earth metals with variable oxidation states are commonly employed to modify Cu-SSZ-39 catalysts due to their outstanding redox properties and reactivity. For instance, Y3+ ions were introduced to induce more Cu2+ ions to locate in the stable ion-exchange (d6r) sites of SSZ-39 zeolite, thus inhibiting the deactivation of Cu-SSZ-39 catalysts even during hydrothermal aging at 900 °C [27]. Furthermore, Y3+ ions could serve as a sacrificial agent for SO2 poisoning to protect active Cu2+ ions from inactivation in harsh conditions [28]. Notably, Wang et al. demonstrated that incorporating Ce or Mn into Cu-SSZ-39 catalysts (i.e., CeCu-SSZ-39 or MnCu-SSZ-39) has a beneficial effect on their low-temperature activity and hydrothermal stability [29]. This was achieved through the formation of highly dispersed active CuO crystallites in the inner channel and the enhancement of acidity. Based on the reaction mechanisms, they proposed a fast SCR mechanism in addition to the Langmuir–Hinshelwood and Eley–Rideal mechanisms, which contributed to the improved low-temperature activity of the MnCu-SSZ-39 catalyst after hydrothermal aging at 850 °C. Moreover, NO2 was discovered to have a stimulating impact on the SCR activity of Cu-SSZ-39 [30]. Consequently, this provides us with the possibility to enhance the low-temperature activity by utilizing the fast SCR pathway for Cu-SSZ-39, while this is not applicable to Cu-SSZ-13 [31,32].
Inspired by previous studies on optimizing the SCR performance of small-pore zeolite catalysts and developing highly effective oxide catalysts [33,34,35,36,37,38,39,40], we selected Ce-Mn oxide to modify the Cu-SSZ-39 catalyst, not only because it possesses excellent low-temperature SCR activity [41], but also because it provides access to the fast SCR reaction pathway via NO2 generation [42]. Herein, Ce-Mn oxide-modified Cu-SSZ-39 catalysts (CeMnOx/Cu-SSZ-39) as well as references (CeO2/Cu-SSZ-39 and MnOx/Cu-SSZ-39) are prepared, and the activity and hydrothermal aging performance of pure and oxide-modified Cu-SSZ-39 catalysts are evaluated for SCR reactions. Particularly, the interaction of oxides with Cu-SSZ-39 and the synergetic effect between Ce and Mn in Ce-Mn oxides are systematically investigated and the possible mechanisms of increased low-temperature activity and hydrothermal stability are speculated on.

2. Results and Discussion

2.1. Structural and Morphological Characterizations

Figure 1a shows the X-ray diffraction (XRD) patterns of the obtained Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts, which present the characteristic diffraction peaks of SSZ-39 zeolites with AEI topology [15,23]. The structure of Cu-SSZ-39 did not change after supporting Ce/Mn or Ce-Mn oxides, while the intensity of peaks belonging to SSZ-39 dropped. Moreover, no peak was ascribed to Ce or Mn oxides, indicating that the supported oxides might be evenly distributed. After hydrothermal aging at 800 °C (Figure 1b), a negligible change in AEI structure was observed for all the catalysts, suggesting the hydrothermal stability of SSZ-39 zeolite. The N2 adsorption–desorption results revealed that all the catalysts displayed the typical isotherms of microporous materials (Figure S1).
Table 1 illustrates that the BET surface areas and pore volumes of the Cu-SSZ-39 catalyst decreased slightly due to oxide supporting and hydrothermal aging, which could be attributed to the loaded oxide covering some micropores of the catalysts. On the other hand, loaded oxides with different textural properties can contribute to the mesopores, leading to the fact that a higher specific surface area does not necessarily correspond to a larger pore volume. The SEM images and corresponding EDS mappings depicted that the fresh and hydrothermally aged samples consisted of aggregated crystals with rectangular morphology and a size of ~700 nm, and highly dispersed Cu, Ce, and Mn species were observed on the catalysts (Figure 2 and Figure S2). Table 1 also summarizes the quantified Si/Al ratios and Cu contents of the catalysts. Compared with Cu-SSZ-39, there was no significant change in the Si/Al ratio and Cu content after supporting oxides or hydrothermal aging.

2.2. SCR Performance

NOx conversion and N2 selectivity as a function of reaction temperature between 150 and 600 °C in the Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts are shown in Figure 3a and Figure S3a. Cu-SSZ-39 showed a so-called “seagull shape” curve resulting from its low Cu loading [18], while MnOx/Cu-SSZ-39 presented an evident decline in NOx conversion over 400 °C. The presence of CeO2 and CeMnOx significantly benefitted the low-temperature NOx reduction performance of the catalysts, particularly for CeMnOx/Cu-SSZ-39, whose active temperature window is from 180 to 600 °C. The higher catalytic performance of CeMnOx/Cu-SSZ-39 in comparison with those of CeO2/Cu-SSZ-39 and MnOx/Cu-SSZ-39 indicated the existence of synergetic effects between Ce and Mn in CeMnOx in improving the low-temperature activity of Cu-SSZ-39 [18,41], which is consolidated by varying Mn/(Ce + Mn) ratios (Figure S4). After hydrothermal aging at 800 °C, NOx conversions of the Cu-SSZ-39-A, CeO2/Cu-SSZ-39-A, and MnOx/Cu-SSZ-39-A catalysts declined, and their active temperature window became narrow (Figure 3b). As for Cu-SSZ-39-A, the NOx conversion decreased to ~90% and remained within the temperature range of 300–600 °C. Therefore, the SCR activity deteriorated seriously after the hydrothermal aging treatment. In contrast, CeO2/Cu-SSZ-39-A and MnOx/Cu-SSZ-39-A still showed similarly high deNOx activity as fresh counterparts despite a slight narrowing in active temperature windows (225–600 °C and 200–400 °C, respectively). Excitingly, CeMnOx/Cu-SSZ-39-A even exhibited a slight increase in low-temperature activity in comparison with CeMnOx/Cu-SSZ-39, presumably due to the interaction of CeMnOx and the aggregated CuO species which facilitated the low-temperature SCR performance [18]. The active temperature window of CeMnOx/Cu-SSZ-39-A was broadened to 175–600 °C, suggesting CeMnOx/Cu-SSZ-39 is a promising catalyst in a wide temperature window. In addition, the N2 selectivity of all the catalysts maintained levels above 95% even after hydrothermal aging (Figure S3). Therefore, the above results indicated the enhanced low-temperature activity and hydrothermal stability of Ce-Mn oxide-modified Cu-SSZ-39 catalysts.

2.3. Structural-Activity Relationships

In order to clarify how CeMnOx enhances the SCR performance of Cu-SSZ-39 catalyst, the catalysts were comprehensively characterized as below. First, X-ray photoelectron spectroscopy (XPS) was employed to investigate the chemical state of surface elements of the catalysts, and the results are summarized in Figure 4 and Table 2. As shown in Figure 4(a1), each Cu 2p3/2 XPS spectrum of the fresh catalysts can be deconvoluted into two peaks at ~935.8 and ~933.4 eV, which are attributed to Cu2+ and CuO species [43,44]. After the addition of Mn/Ce or Ce-Mn oxides, the peaks of Cu2+ and CuO species in the oxide-modified Cu-SSZ-39 catalysts all shift towards lower binding energies due to electron transfer from the Mn/Ce oxides to Cu species because of the higher electronegativity of Cu2+ ions [43]. After the hydrothermal treatment, the Cu2+ ratios in the aged catalysts decreased due to the aggregation of isolated Cu species to surface CuO species (Figure 4(a2) and Table 2). Notably, the binding energies of the Cu species in CeO2/Cu-SSZ-39-A, CeMnOx/Cu-SSZ-39-A, and MnOx/Cu-SSZ-39-A catalysts are all lower than that in Cu-SSZ-39-A, suggesting that there still exist strong interactions between the supported oxides and Cu species in Cu-SSZ-39 even after hydrothermal aging. Furthermore, CeMnOx/Cu-SSZ-39-A maintained the highest amount of Cu2+ ions among the oxide-modified catalysts after hydrothermal aging (Table 2), which revealed its superior SCR activity (Figure 3b). This also indicated CeMnOx could effectively stabilize Cu-SSZ-39 zeolite and interact with the generated CuO species, contributing to its excellent hydrothermal resistance [27].
Figure 4(b1,b2) shows the O 1s XPS spectra of the catalysts. The spectra could be fitted into three peaks, corresponding to the oxygen in the adsorbed hydroxyl group on the surface (Osur), chemisorbed surface oxygen (Oads), and lattice oxygen (Olat), respectively [45]. A new peak centered at ~530 eV appeared in CeO2/Cu-SSZ-39 and CeMnOx/Cu-SSZ-39, which is associated with lattice oxygen of CeO2 [34]. The calculated Oads concentrations in CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 were higher than that in Cu-SSZ-39, which facilitated an increase in the mobility of oxygen and the redox ability, and thus could accelerate low-temperature SCR reactions [37,46]. The increased Oads species could be attributed to the addition of CeO2, MnOx, or CeMnOx with excellent redox properties, which provided more vacancies for Oads. Furthermore, CeMnOx/Cu-SSZ-39, with the largest amount of Oads, exhibited the highest low-temperature activity among the oxide-modified Cu-SSZ-39 catalysts (Figure 3a). However, the amount of Osur decreased with the addition of oxides, indicating the hydroxyl group as the Brønsted acid site was covered by the supported oxides. After the hydrothermal aging treatment, the ratios of Oads species decreased due to partial sintering of the oxides.
As for the Ce 3d spectra, eight peaks were detected (Figure 4c), marked as u′ (903.5–904.0 eV) and v′ (885.1–885.5 eV) representing Ce3+, and with others marked as u‴ (916.7–917.0 eV), u″ (907.6–908.1 eV), u (900.5–901.2 eV), v‴ (898.1–898.5 eV), v″ (888.7–889.0 eV), and v (882.1–882.4 eV), attributed to Ce4+ [37]. Clearly, the surface Ce3+ ratio of CeMnOx/Cu-SSZ-39 was higher than that of CeO2/Cu-SSZ-39, and a similar trend was maintained even after hydrothermal aging. This may be a result of the redox cycles between Mnn+ and Cen+ as followed by Mn3+ + Ce4+ ↔ Mn4+ + Ce3+ [38]. The presence of Ce3+ may cause an environment with imbalanced charges, resulting in an increase in oxygen vacancies. This, in turn, enhances the activation of surface oxygen species and improves the SCR activity [47,48].
The Mn 2p3/2 spectra in Figure 4d were fitted into three peaks, corresponding to Mn4+ species (644.0–644.5 eV), Mn3+ species (642.3–642.7 eV), and Mn2+ species (640.9–641.6 eV), respectively [37]. As shown in Table 2, Mn4+ species are the dominant species in the catalysts, especially in CeMnOx/Cu-SSZ-39 and CeMnOx/Cu-SSZ-39-A. The higher ratio of Mn4+ species in CeMnOx/Cu-SSZ-39 and CeMnOx/Cu-SSZ-39-A indicates that the synergistic effect between Ce and Mn promoted the formation of Mn4+ species [45,49]. According to previous publications [38,50], the higher surface content of Mn4+ species can improve the catalytic activity at low temperatures, which is consistent with the SCR performance of the catalysts.
To sum up, the XPS analysis indicated that the synergistic effect between Ce and Mn resulted in an increase in active Cu2+ sites, Oads, Ce3+, and Mn4+ species in Ce-Mn oxide-modified Cu-SSZ-39 catalysts with different Mn/(Ce + Mn) ratios (Figure S5 and Table S1). These factors contributed to the improved redox property and low-temperature SCR activity.
To investigate the redox property of the catalysts, which is intimately related to their SCR performance, a temperature programmed reduction of hydrogen (H2-TPR) was carried out, and the results are shown in Figure 5. Three peaks are observed in the H2-TPR profile of Cu-SSZ-39 (Figure 5a). The peak at ~230 °C is ascribed to the reduction of tri-coordinated Cu2+ to Cu+, i.e., [Cu(OH)]+ ions that are located in the window of eight-membered rings balanced with one framework negative charge (denoted as ZCuOH) [43]. As the XPS results confirmed the existence of CuO species in Cu-SSZ-39 (Figure 4(a1)), the broad peak ranging from 300 to 500 °C is attributed to the reduction of CuO and Cu2+ ions located in the window of six-membered rings and coordinated with 2Al sites (denoted as Z2Cu) [28]. The peak above 500 °C is assigned to the reduction of Cu+ ions to Cu0 [51]. After supporting the oxides, the H2-TPR profiles became more complicated, and some new H2 consumption peaks were detected. In the case of CeO2/Cu-SSZ-39, the emergence of broad reduction peaks centered at ~350 and 670 °C is ascribed to the reduction of CeO2 (Ce4+ to Ce3+) in a mixed oxide phase [34,52]. For MnOx/Cu-SSZ-39, two prominent peaks at ~300 and 650 °C were detected, which could be assigned to the reduction of MnOx (Mn4+ to Mn3+) and Mn2O3 (Mn3+ to Mn2+), respectively [38,53]. In the H2-TPR profile of CeMnOx/Cu-SSZ-39, the reduction peaks of ZCuOH, MnOx, Mn2O3, and CeO2 shifted to lower temperatures, suggesting the promoted redox property after supporting CeMnOx on Cu-SSZ-39. A similar phenomenon was observed in Ce-Mn oxide-modified Cu-SSZ-39 catalysts with different Mn/(Ce + Mn) ratios (Figure S6). These results indicate that the interaction of CeMnOx with Cu-SSZ-39 and the synergistic effect between Ce and Mn in CeMnOx improve the redox property, and thus contribute to the enhanced SCR activity.
After the hydrothermal aging treatment, the reduction peaks of ZCuOH species almost disappeared in the catalysts, while the peaks of CuO species became more pronounced due to the generation of CuO species with the deconstruction of the zeolite framework (Figure 5b). Particularly for CeMnOx/Cu-SSZ-39-A, the reduction peak of CuO species shifted to lower temperatures, suggesting that the interaction between CeMnOx and CuO promoted the redox property and, thus, maintained the excellent SCR activity even after hydrothermal aging (Figure 3b). Additionally, this interaction effectively prevented the aggregation of CuO as well as the further exacerbation of the zeolite collapse.
The acidity of the catalysts also played a crucial role in the analysis of NH3 adsorption and activation during the SCR reaction [54]. Temperature programmed desorption of NH3 (NH3-TPD) was conducted to determine the acid strength of the catalysts, and the results are illustrated in Figure 6. The curves of the fresh catalysts show similar features with three desorption peaks (Figure 6a). The literature suggests that the peak at approximately 220 °C (Peak α) is attributed to the weakly adsorbed NH3, mainly terminal hydroxyl species. The medium-temperature peak (Peak β) is ascribed to the NH3 adsorbed on the strong Lewis acid sites created by the metal ions at around 305 °C. Finally, the high-temperature desorption peak (Peak γ) above 480 °C is attributed to the NH3 desorption from the Brønsted acid sites [17,25,29,55]. To quantify the number of acid sites, the acidity of fresh and aged catalysts was calculated and is summarized in Table 3. The total acid amounts in CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 are lower than that in Cu-SSZ-39, which is probably due to the acid sites being partially covered by the supported oxides, consistent with XPS results. Notably, the oxide-modified Cu-SSZ-39 catalysts had larger sums of weak and medium acidity compared with Cu-SSZ-39, which indicated that the existence of Ce/Mn or Ce-Mn oxide in Cu-SSZ-39 could promote the adsorption and activation of NH3 at low and medium temperatures, thus facilitating the SCR reaction [56].
Compared to the fresh catalysts, the desorption temperatures of the peaks over the aged ones shifted towards lower values and the NH3 desorption amount significantly decreased (Figure 6b and Table 3), suggesting that both the strength and amount of acid sites are affected by the hydrothermal aging treatment. The decrease in Peaks α and γ can be attributed to the collapse of the zeolite framework and the loss of active sites; in contrast, the increased intensity in Peak β indicates that the additional Lewis acid sites arise from the transformation of Cu2+ ions and the aggregation of oxides during hydrothermal aging [4]. Interestingly, CeO2/Cu-SSZ-39-A, CeMnOx/Cu-SSZ-39-A, and MnOx/Cu-SSZ-39-A catalysts exhibited a slightly lower decrease in Peaks α and γ compared to Cu-SSZ-39-A. The lesser disruptive effect on the acid sites can be attributed to the interaction of the supported oxides with Cu-SSZ-39, which results in their higher activity and hydrothermal stability, as indicated in the SCR performance of the aged catalysts (Figure 3b).
In order to reveal the stabilizing effect of CeMnOx on the zeolite framework of Cu-SSZ-39, 29Si, and 27Al, MAS NMR spectra were measured to determine the local Si and Al environment in fresh and hydrothermally aged Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts. As illustrated in Figure 7a, the chemical shifts at −101, −106, and −112 ppm were assigned to Si (4Si), Si (3Si, 1Al), and Si (2Si, 2Al), respectively. As shown in Table 4, the = Si/Al ratios calculated using 29Si MAS NMR were in good agreement with those of Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts derived from EDS mappings (Table 1). However, the hydrothermal aging treatment resulted in an increase in Si/Al ratios (calculated using NMR), suggesting significant dealumination of the zeolite framework. Notably, CeMnOx/Cu-SSZ-39-A exhibited a much lower increase in the Si/Al ratio compared to Cu-SSZ-39-A, indicating its higher hydrothermal stability. The 27Al MAS NMR spectra (Figure 7b) showed features at 58 and 0 ppm, which corresponded to tetra-coordinated framework Al and hexa-coordinated extra-framework Al, respectively [23]. After hydrothermal aging, the feature at 33 ppm, arising from the penta-coordinated extra-framework Al, emerged, indicating that Al ions detached from the framework [43]. In order to quantify differently coordinated Al species, 27Al MAS NMR spectra were deconvoluted, and the results are also listed in Table 4. It is worth noting that CeMnOx/Cu-SSZ-39-A preserved a greater amount of tetrahedrally coordinated framework Al than Cu-SSZ-39-A, suggesting that it possesses a higher hydrothermal stability. Combining the results of NMR and SCR activity, the supported CeMnOx on Cu-SSZ-39 could effectively inhibit the dealumination from the zeolite framework and thus guarantee the excellent hydrothermal stability and SCR activity of CeMnOx/Cu-SSZ-39-A.

2.4. Reaction Kinetics and Possible Mechanisms

To further investigate the essential influence of CeMnOx on the SCR performance of Cu-SSZ-39, CeMnOx/Cu-SSZ-39 and Cu-SSZ-39 were measured in an SCR reaction under kinetics-controlled differential conditions (Figure S7), and the Arrhenius plots are shown in Figure 8. The apparent activation energies (Ea) depend on the reaction mechanisms of the catalyst being employed, and the Ea of Cu-SSZ-39 was 50.2 kJ·mol−1, which is comparable to that reported in previous studies [30]. Distinctively, the Ea of CeMnOx/Cu-SSZ-39 (17.1 kJ·mol−1) was much lower than that of Cu-SSZ-39, suggesting that Ce-Mn oxide could decrease the energy barriers and accelerate the SCR reaction rate, especially at low temperatures [38]. Moreover, the huge difference in Ea values indicates that the SCR reaction in CeMnOx/Cu-SSZ-39 probably proceeded through another pathway, i.e., fast SCR [55]. The literature suggests that the formation of NO2 from NO oxidation is a crucial process for low-temperature activity since NO2 can initiate fast SCR reactions [20]. Consequently, NO oxidation experiments were conducted, and the results are shown in Figure 9. The NO conversion of Cu-SSZ-39 was low and there was almost no change above 200 °C, indicating that NO was hardly oxidized into NO2, which rules out the possibility that the formation of NO2 in Cu-SSZ-39 catalysts facilitated NOx conversion through the fast SCR pathway. Notably, NO oxidation in CeMnOx/Cu-SSZ-39 was much higher than that in Cu-SSZ-39 and increased rapidly with increasing temperatures, due to the improved redox property derived from CeMnOx, which was confirmed by the XPS and H2-TPR results (Figure 4 and Figure 5). Based on the literature, the generated NO2 from NO oxidation participates in fast SCR reactions, and its reaction rate is higher than that of the standard SCR reaction (Figure S7); thus, the low-temperature SCR performance is improved [29,30].
To corroborate this speculation, the performance of Cu-SSZ-39 under fast SCR feed conditions was measured. As shown in Figure 10, the presence of NO2 in the feed increased the NOx conversions at low temperatures for Cu-SSZ-39 in comparison with under the standard SCR condition (Figure S8), confirming that both NO and NO2 took part in the fast SCR reactions. This improved low-temperature activity is derived from the promoting effect of NO2 on the SCR performance of Cu-SSZ-39 catalysts [30,55]. For CeMnOx/Cu-SSZ-39, further enhancement of NO and NO2 conversions was observed compared with Cu-SSZ-39 at low temperatures (Figure 10 and Figure S8), which could be basically attributed to NO oxidation and the low-temperature SCR activity of CeMnOx. Notably, the NO2 conversion of CeMnOx/Cu-SSZ-39 did not increase with temperature as in the case of Cu-SSZ-39, and it had a minimum value at 175 °C. According to previous research, this decreased NO2 conversion is probably due to the reduction of NH4NO3 by NO to generate NO2 [32,34]. As such, the catalytic effect of CeMnOx on NO oxidation provides a further impetus for accessing the fast SCR pathway of Cu-SSZ-39 [53].

3. Materials and Methods

3.1. Preparation

Na-SSZ-39 (Si/Al = 8.7) was synthesized by the inter-zeolite transformations of FAU zeolite according to procedures described in the literature. Y (Si/Al = 3.1) and N,N-dimethyl-3,5-dimethylpiperidinium hydroxide (DMDMPOH) were used as a precursor and an organic structure directing agent (OSDA), respectively [19]. Typically, 15.2 mg NaOH was dissolved in 35.6 g deionized water; then, 10.1 mL DMDMPOH was added and stirred for 30 min. After that, 20.1 mL Na2SiO3 and 1.0 g Y zeolite were added to the above mixture with continuous stirring for 2 h, and the final molar composition of the mixture was 1 SiO2: 0.015 Al2O3: 0.559 NaOH: 0.116 OSDA: 24.1 H2O. The final mixture was transferred to a 100 mL Teflon-lined stainless-steel autoclave and crystallized at 140 °C for 24 h. The as-prepared zeolite was then centrifuged, washed, and dried at 105 °C overnight, followed by calcination at 575 °C for 8 h in air. Subsequently, the resulting Na-SSZ-39 zeolite was ion-exchanged with (NH4)2SO4 and CuSO4 solutions. After drying at 105 °C overnight and calcining at 575 °C for 8 h, the Cu-SSZ-39 catalyst was obtained.
Ce-Mn oxide-modified Cu-SSZ-39 was prepared via the impregnation method by using Ce(NO3)3 and Mn(NO3)2 as precursors. The total component of Ce-Mn composite oxide was 5 wt.% and the molar ratio of Mn/(Ce + Mn) was 0.3. For more detail, Cu-SSZ-39 was slowly added to the Ce(NO3)3 and Mn(NO3)2 mixed solution under ultrasonic conditions. The amount of mixed solution was equal to the pore volumes of Cu-SSZ-39 zeolite. And then, the mixture was dried overnight at 105 °C. After being calcined at 550 °C for 4 h in muffle, the modified catalysts were obtained and labeled as CeMnOx/Cu-SSZ-39. Additionally, Ce-Mn oxide-modified Cu-SSZ-39 catalysts with different Mn/(Ce + Mn) ratios [Mn/(Ce + Mn) = 0.2, 0.4, and 0.5] were also prepared to further demonstrate the synergetic effects between Ce and Mn in Ce-Mn oxides. For comparison, CeO2- and MnOx-modified Cu-SSZ-39 catalysts were prepared in the same way (denoted as CeO2/Cu-SSZ-39 and MnOx/Cu-SSZ-39) and used as references.
Hydrothermal aging of the catalysts was carried out in a flowing wet-air stream containing 10 vol.% H2O and held at 800 °C for 16 h. The obtained aged catalysts were denoted as Cu-SSZ-39-A, CeMnOx/Cu-SSZ-39-A, CeO2/Cu-SSZ-39-A, and MnOx/Cu-SSZ-39-A, respectively.

3.2. Characterizations

XRD measurements were performed on Smartlab SE X-ray diffractometer (Rigaku, Tokyo, Japan) with Cu Kα radiation. The diffraction patterns were taken in the 2θ range of 5–50°. The surface area and pore structure were analyzed using N2 adsorption/desorption measurements obtained with a Micromeritics ASAP 2020 instrument (Micromeritics, Norcross, GA, USA). The samples were degassed at 300 °C for 8 h before analysis. A GeminiSEM 300 microscope (Carl Zeiss, Oberkochen, Germany) was used to perform scanning electron microscopy (SEM) equipped with energy dispersive spectroscopy (EDS). X-ray photoelectron spectroscopy (XPS) was performed on an AXIS Supra instrument (Shimadzu, Kyoto, Japan), calibrating binding energies using a C 1s peak at 284.8 eV. A Bruker AVANCE NEO 600 WB (Bruker, Saarbrucken, Germany) was used to conduct magnetic angle spinning nuclear magnetic resonance (MAS NMR) experiments. The resonance frequencies of 29Si and 27Al are 119.23 and 156.38 MHz, respectively. To record 29Si MAS NMR spectra, a 4 mm MAS probe was used with a spinning frequency of 10 kHz, a recycle delay of 2 s, and a π/2 pulse length of 1.67 μs. For 27Al MAS NMR spectra, a 4 mm MAS probe was used with a spinning frequency of 15 kHz, a pulse delay of 20 s, and a π/2 pulse length of 1.67 μs.
Temperature programmed reduction with H2 (H2-TPR) experiments were performed using a XIANQUAN TP-5080D chemisorption analyzer (XIANQUAN, Tianjin, China). The sample (100 mg) was treated at 500 °C for an hour in an O2 flow before conducting TPR in 5 vol.% H2/N2 at a flow rate of 20 mL∙min−1. To perform the temperature programmed desorption with NH3 (NH3-TPD) experiments, a mass spectrometer (OmniStar GSD 320, Pfeiffer, Germany) was used. Before the experiments, the samples were pretreated at 500 °C for 30 min in 10 vol.% O2/He and cooled to 100 °C. NH3 was adsorbed in 0.4 vol.% NH3/He until saturation. Then, the samples were purified by purging with He to clean the surface. The TPD experiments involved increasing the temperature from 100 to 800 °C at a ramp rate of 10 °C·min−1 under He flow.

3.3. Catalyst Activity and Reaction Kinetics Measurements

NH3-SCR activity was performed in a fixed-bed reactor with a thermocouple placed inside the catalyst bed (about 0.1 g, 40–60 mesh) and in the temperature range 150–600 °C. The reactant gas conditions were as follows: 500 ppm of NOx (500 ppm of NO for standard SCR or 250 ppm of NO2 + 250 ppm of NO for fast SCR), 500 ppm of NH3, 5.3 vol.% of O2, and He as the balance gas. The total gas flow rate was maintained at 300 mL∙min−1 and the gas hourly space velocity (GHSV) was 100,000 h−1. The concentrations of NO and NO2 were monitored using a chemiluminescence NOx analyzer (Model 42i-HL, Thermo Electron Corporation, Waltham, MA, USA), while N2O and NH3 were recorded using a quadrupole mass spectrometer (OmniStar GSD 301, Pfeiffer, Germany). The NOx conversion and N2 selectivity were calculated using Equations (1) and (2):
NO x   conversion = NO x in [ NO x ] out [ NO x ] in × 100 %
N 2   selectivity = [ NO x ] in + [ NH 3 ] in NO x out - NH 3 out 2 × [ N 2 O ] [ NO x ] in + [ NH 3 ] in [ NO x ] out [ NH 3 ] out × 100 %
The experiment of NO oxidation was conducted using the same procedure as the NH3-SCR test. The inlet gas composition was as follows: [NO] = 500 ppm, [O2] = 5.3 vol.%, balanced by He. We calculated the NO conversion using Equation (3):
NO   conversion = NO in [ NO ] out [ NO ] in × 100 %
To obtain kinetic data within the kinetic regime and eliminate the effects of mass transfer limitations, we selected particles of samples and a total flow rate of 100–200 mesh and 600 mL∙min−1, respectively. Additionally, we estimated the GHSV to be 800,000 h−1 with a decreased amount of catalyst of 0.045 mL. The apparent activation energy (Ea) was calculated based on the Arrhenius plot of the reaction rate (r) versus 1/T (T is the temperature) when the NOx conversion was kept below 15% [57].

4. Conclusions

In this study, Ce-Mn oxide-modified Cu-SSZ-39 catalysts (CeMnOx/Cu-SSZ-39) along with CeO2/Cu-SSZ-39 and MnOx/Cu-SSZ-39 references were prepared through the ion-exchange of Cu ions followed by impregnation of the oxide precursors. The supported oxides could improve the low-temperature SCR activity for Cu-SSZ-39. Especially for CeMnOx/Cu-SSZ-39, the low-temperature SCR activity was considerably improved, achieving above 90% NOx conversion between 180 and 600 °C. Notably, this superior activity was maintained even after hydrothermal aging at 800 °C, indicating excellent hydrothermal stability. The supported oxides rendered more Cu2+, Ce3+/Mn4+, and adsorbed oxygen species in the catalyst, and all of these enhanced the redox properties and acidity, which facilitated the SCR reaction at low temperatures. Furthermore, CeMnOx/Cu-SSZ-39 took advantage of CeMnOx, as well as the strong interaction between CeMnOx and Cu-SSZ-39, resulting in exceptional SCR activity over a wide temperature range. Importantly, the interaction of Cu-SSZ-39 with oxides, especially CeMnOx, contributed to the remarkable hydrothermal stability of the zeolite by effectively inhibiting the dealumination from the zeolite framework and preventing the aggregation of Cu species during the hydrothermal aging process. The NO oxidation and fast SCR activity measurements confirmed that the catalytic effect of CeMnOx on NO oxidation provides an impetus for accessing the fast SCR pathway of Cu-SSZ-39. Therefore, CeMnOx opens up the fast SCR pathway by oxidizing NO to NO2, which significantly enhances the low-temperature activity. Our research results not only demonstrate the importance of synergy in Ce-Mn oxides, but also provide a new approach for developing a highly effective Cu-SSZ-39 zeolite SCR catalyst over a wide temperature range, which is of great significance for achieving efficient NOx removal in diesel vehicles.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14010010/s1, Figure S1: N2 adsorption/desorption isotherms (a) and pore size distribution curves (b) of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts; Figure S2: SEM images and Cu, Ce, Mn element EDS mappings of Cu-SSZ-39 (a,e), CeO2/Cu-SSZ-39 (b,f), CeMnOx/Cu-SSZ-39 (c,g), and MnOx/Cu-SSZ-39 (d,h) catalysts: (a–d) fresh and (e–h) hydrothermally aged samples; Figure S3: N2 selectivity of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermal aged samples; Figure S4: NOx conversion of CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.2], CeMnOx/Cu-SSZ-39, CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.4], CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.5], and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples; Figure S5: XPS spectra of Cu 2p3/2 (a) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve); O 1s (b) (black line was assigned to the original data, dark yellow line was assigned to the fitting curve); Ce 3d (c) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve); and Mn 2p3/2 (d) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve) of CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.2], CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.4], and CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.5] catalysts; Figure S6: H2-TPR profiles of CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.2], CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.4], and CeMnOx/Cu-SSZ-39 [Mn/(Ce + Mn) = 0.5] catalysts; Figure S7: Kinetic catalytic performances of Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts; Figure S8: NOx conversion under fast SCR condition in Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts; Table S1: Cu 2p3/2, O 1s, Ce 3d, and Mn 2p3/2 XPS analysis of the catalysts.

Author Contributions

Conceptualization, Y.X.; formal analysis, A.T., F.Y., Y.X., X.Z., L.Y. and S.L.; investigation, A.T., F.Y., X.Z. and L.Y.; data curation, A.T., F.Y., X.Z., S.L., D.H., J.J. and Y.L.; writing—original draft preparation, A.T.; writing—review and editing, Y.X., Z.L. and Z.Z.; supervision, Y.X. and Z.Z.; funding acquisition, Y.X. and Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Nos. 22376078, 22276070, and 22076062), the China Postdoctoral Science Foundation (No. 2022M711957), the Taishan Scholar Program of Shandong (No. tstp20230628), the Shandong Provincial Natural Science Foundation (No. ZR2023ZD39), the National Engineering Laboratory for Mobile Source Emission Control Technology (No. NELMS2019A14), and the Project of Jinan Municipal Bureau of Science and Technology (No. 2020GXRC021).

Data Availability Statement

The data presented in this study are available.

Conflicts of Interest

Z.L. was employed by China Automotive Technology & Research Center Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Richter, A.; Burrows, J.; Nüß, H.; Granier, C.; Niemeier, U. Increase in tropospheric nitrogen dioxide over China observed from space. Nature 2005, 437, 129–132. [Google Scholar] [CrossRef]
  2. Guan, B.; Zhan, R.; Lin, H.; Huang, Z. Review of state of the art technologies of selective catalytic reduction of NOx from diesel engine exhaust. Appl. Therm. Eng. 2014, 66, 395–414. [Google Scholar] [CrossRef]
  3. Li, P.; Xin, Y.; Zhang, H.; Yang, F.; Tang, A.; Han, D.; Jia, J.; Wang, J.; Li, Z.; Zhang, Z. Recent progress in performance optimization of Cu-SSZ-13 catalyst for selective catalytic reduction of NOx. Front. Chem. 2022, 10, 1033255. [Google Scholar] [CrossRef]
  4. Yang, F.; Xin, Y.; Zhu, X.; Tang, A.; Yu, L.; Han, D.; Jia, J.; Lu, Y.; Zhang, Z. Hard template-assisted trans-crystallization synthesis of hierarchically porous Cu-SSZ-13 with enhanced NH3-SCR performance. Catalysts 2023, 13, 1217. [Google Scholar] [CrossRef]
  5. Li, J.; Meng, X.; Xiao, F. Zeolites for control of NOx emissions: Opportunities and challenges. Chem Catal. 2022, 2, 253–261. [Google Scholar] [CrossRef]
  6. Simancas, R.; Chokkalingam, A.; Elangovan, S.; Liu, Z.; Sano, T.; Iyoki, K.; Wakihara, T.; Okubo, T. Recent progress in the improvement of hydrothermal stability of zeolites. Chem. Sci. 2021, 12, 7677–7695. [Google Scholar] [CrossRef]
  7. Shan, Y.; Shan, W.; Shi, X.; Du, J.; Yu, Y.; He, H. A comparative study of the activity and hydrothermal stability of Al-rich Cu-SSZ-39 and Cu-SSZ-13. Appl. Catal. B-Environ. 2020, 264, 118511. [Google Scholar] [CrossRef]
  8. Schmieg, S.; Oh, S.; Kim, C.; Brown, D.; Lee, J.; Peden, C.; Kim, D. Thermal durability of Cu-CHA NH3-SCR catalysts for diesel NOx reduction. Catal. Today 2012, 184, 252–261. [Google Scholar] [CrossRef]
  9. Ryu, T.; Ahn, N.; Seo, S.; Cho, J.; Kim, H.; Jo, D.; Park, G.; Kim, P.; Kim, C.; Bruce, E.; et al. Fully copper-exchanged high-silica LTA zeolites as unrivaled hydrothermally stable NH3-SCR catalysts. Angew. Chem. Int. Ed. 2017, 56, 3256–3260. [Google Scholar] [CrossRef]
  10. Ahn, N.; Ryu, T.; Kang, Y.; Kim, H.; Shin, J.; Nam, I.; Hong, S. The origin of an unexpected increase in NH3-SCR activity of aged Cu-LTA catalysts. ACS Catal. 2017, 7, 6781–6785. [Google Scholar] [CrossRef]
  11. Kim, J.; Cho, S.; Kim, D. Facile synthesis of KFI-type zeolite and its application to selective catalytic reduction of NOx with NH3. ACS Catal. 2017, 7, 6070–6081. [Google Scholar] [CrossRef]
  12. Fickel, D.; D’Addio, E.; Lauterbach, J.; Lobo, R. The ammonia selective catalytic reduction activity of copper-exchanged small-pore zeolites. Appl. Catal. B-Environ. 2011, 102, 441–448. [Google Scholar] [CrossRef]
  13. Li, R.; Jiang, X.; Lin, J.; Zhang, Z.; Huang, Q.; Fu, G.; Zhu, Y.; Jiang, J. Understanding the influence of hydrothermal treatment on NH3-SCR of NOx activity over Cux-SSZ-16. Chem. Eng. J. 2022, 441, 136021. [Google Scholar] [CrossRef]
  14. Shan, Y.; Shi, X.; Du, J.; Yu, Y.; He, H. Cu-exchanged RTH-type zeolites for NH3-selective catalytic reduction of NOx: Cu distribution and hydrothermal stability. Catal. Sci. Technol. 2019, 9, 106–115. [Google Scholar] [CrossRef]
  15. Moliner, M.; Franch, C.; Palomares, E.; Grill, M.; Corma, A. Cu-SSZ-39, an active and hydrothermally stable catalyst for the selective catalytic reduction of NOx. Chem. Commum. 2012, 48, 8264–8266. [Google Scholar] [CrossRef]
  16. Martín, N.; Boruntea, C.; Moliner, M.; Corma, A. Efficient synthesis of the Cu-SSZ-39 catalyst for DeNOx applications. Chem. Commum. 2015, 51, 11030–11033. [Google Scholar] [CrossRef]
  17. Xu, H.; Zhang, J.; Wu, Q.; Chen, W.; Lei, C.; Zhu, Q.; Han, S.; Fei, J.; Zheng, A.; Zhu, L.; et al. Direct synthesis of aluminosilicate SSZ-39 zeolite using colloidal silica as a starting source. ACS Appl. Mater. Interfaces 2019, 11, 23112–23117. [Google Scholar] [CrossRef]
  18. Du, J.; Shan, Y.; Sun, Y.; Gao, M.; Liu, Z.; Shi, X.; Yu, Y.; He, H. Unexpected increase in low-temperature NH3-SCR catalytic activity over Cu-SSZ-39 after hydrothermal aging. Appl. Catal. B-Environ. 2021, 294, 120237. [Google Scholar] [CrossRef]
  19. Chen, J.; Shan, Y.; Sun, Y.; Ding, W.; Xue, S.; Han, X.; Du, J.; Yan, Z.; Yu, Y.; He, H. Hydrothermal aging alleviates the phosphorus poisoning of Cu-SSZ-39 catalysts for NH3-SCR reaction. Environ. Sci. Technol. 2023, 57, 4113–4121. [Google Scholar] [CrossRef] [PubMed]
  20. Han, L.; Cai, S.; Gao, M.; Hasegawa, J.; Wang, P.; Zhang, J.; Shi, L.; Zhang, D. Selective catalytic reduction of NOx with NH3 by using novel catalysts: State of the art and future prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef] [PubMed]
  21. Xu, H.; Chen, W.; Wu, Q.; Lei, C.; Zhang, J.; Han, S.; Zhang, L.; Zhu, Q.; Meng, X.; Dai, D.; et al. Transformation synthesis of aluminosilicate SSZ-39 zeolite from ZSM-5 and beta zeolite. J. Mater. Chem. A 2019, 7, 4420–4425. [Google Scholar] [CrossRef]
  22. Ipek, B.; Wulfers, M.; Kim, H.; Göltl, F.; Hermans, I.; Smith, J.; Booksh, K.; Brown, C.; Lobo, R. Formation of [Cu2O2]2+ and [Cu2O]2+ toward C–H bond activation in Cu-SSZ-13 and Cu-SSZ-39. ACS Catal. 2017, 7, 4291–4303. [Google Scholar] [CrossRef]
  23. Fu, G.; Yang, R.; Liang, Y.; Yi, X.; Li, R.; Yan, N.; Zheng, A.; Yu, L.; Yang, X.; Jiang, J. Enhanced hydrothermal stability of Cu/SSZ-39 with increasing Cu contents, and the mechanism of selective catalytic reduction of NO. Micropor. Mesopor. Mat. 2021, 320, 111060. [Google Scholar] [CrossRef]
  24. Du, J.; Han, S.; Huang, C.; Shan, Y.; Zhang, Y.; Shan, W.; He, H. Comparison of precursors for the synthesis of Cu-SSZ-39 zeolite catalysts for NH3-SCR reaction. Appl. Catal. B-Environ. 2023, 338, 123072. [Google Scholar] [CrossRef]
  25. Sun, G.; Yu, R.; Xu, L.; Wang, B.; Zhang, W. Enhanced hydrothermal stability and SO2-tolerance of Cu–Fe modified AEI zeolite catalysts in NH3-SCR of NOx. Catal. Sci. Technol. 2022, 12, 3898–3911. [Google Scholar] [CrossRef]
  26. Sonoda, T.; Maruo, T.; Yamasaki, Y.; Tsunoji, N.; Takamitsu, Y.; Sadakane, M.; Sano, T. Synthesis of high-silica AEI zeolites with enhanced thermal stability by hydrothermal conversion of FAU zeolites, and their activity in the selective catalytic reduction of NOx with NH3. J. Mater. Chem. A 2015, 3, 857–865. [Google Scholar] [CrossRef]
  27. Lin, Q.; Xu, S.; Zhao, H.; Liu, S.; Xu, H.; Dan, Y.; Chen, Y. Highlights on key roles of Y on the hydrothermal stability at 900 °C of Cu/SSZ-39 for NH3-SCR. ACS Catal. 2022, 12, 14026–14039. [Google Scholar] [CrossRef]
  28. Yu, R.; Kong, H.; Zhao, Z.; Shi, C.; Meng, X.; Xiao, F.; De Baerdemaeker, T.; Parvulescu, A.; Müller, U.; Zhang, W. Rare-earth yttrium exchanged Cu-SSZ-39 zeolite with superior hydrothermal stability and SO2-tolerance in NH3-SCR of NOx. ChemCatChem 2022, 14, e202200228. [Google Scholar] [CrossRef]
  29. Wang, Y.; Li, G.; Zhang, S.; Zhang, X.; Zhang, X.; Hao, Z. Promoting effect of Ce and Mn addition on Cu-SSZ-39 zeolites for NH3-SCR reaction: Activity, hydrothermal stability, and mechanism study. Chem. Eng. J. 2020, 393, 124782. [Google Scholar]
  30. Zhu, N.; Shan, Y.; Shan, W.; Sun, Y.; Liu, K.; Zhang, Y.; He, H. Distinct NO2 effects on Cu-SSZ-13 and Cu-SSZ-39 in the selective catalytic reduction of NOx with NH3. Environ. Sci. Technol. 2020, 54, 15499–15506. [Google Scholar] [CrossRef]
  31. Shan, Y.; He, G.; Du, J.; Sun, Y.; Liu, Z.; Fu, Y.; Liu, F.; Shi, X.; Yu, Y.; He, H. Strikingly distinctive NH3-SCR behavior over Cu-SSZ-13 in the presence of NO2. Nat. Commun. 2022, 13, 4606. [Google Scholar] [CrossRef] [PubMed]
  32. Cui, Y.; Gao, F. Cu loading dependence of fast NH3-SCR on Cu/SSZ-13. Emiss. Control Sci. Technol. 2019, 5, 124–132. [Google Scholar] [CrossRef]
  33. Liu, Q.; Fu, Z.; Ma, L.; Niu, H.; Liu, C.; Li, J.; Zhang, Z. MnOx-CeO2 supported on Cu-SSZ-13: A novel SCR catalyst in a wide temperature range. Appl. Catal. A-Gen. 2017, 547, 146–154. [Google Scholar] [CrossRef]
  34. Andana, T.; Rappé, K.; Nelson, N.; Gao, F.; Wang, Y. Selective catalytic reduction of NOx with NH3 over Ce-Mn oxide and Cu-SSZ-13 composite catalysts–Low temperature enhancement. Appl. Catal. B-Environ. 2022, 316, 121522. [Google Scholar] [CrossRef]
  35. Huang, Y.; Zhao, H.; Wu, X.; Huang, Z.; Shen, H.; Jing, G. Enhanced low-temperature SCR performance of metal oxide/Cu-SSZ-13 composite catalysts: The role of oxygen species in metal oxide. Chem. Eng. J. 2023, 475, 146407. [Google Scholar] [CrossRef]
  36. Andana, T.; Rappé, K.; Gao, F.; Szanyi, J.; Pereira-Hernandez, X.; Wang, Y. Recent advances in hybrid metal oxide–zeolite catalysts for low-temperature selective catalytic reduction of NOx by ammonia. Appl. Catal. B-Environ. 2021, 291, 120054. [Google Scholar] [CrossRef]
  37. Chen, L.; Ren, S.; Liu, L.; Su, B.; Yang, J.; Chen, Z.; Wang, M.; Liu, Q. Catalytic performance over Mn-Ce catalysts for NH3-SCR of NO at low temperature: Different zeolite supports. J. Environ. Chem. Eng. 2022, 10, 107167. [Google Scholar] [CrossRef]
  38. Chen, L.; Ren, S.; Jiang, Y.; Liu, L.; Wang, M.; Yang, J.; Chen, Z.; Liu, W.; Liu, Q. Effect of Mn and Ce oxides on low-temperature NH3-SCR performance over blast furnace slag-derived zeolite X supported catalysts. Fuel 2022, 320, 123969. [Google Scholar] [CrossRef]
  39. Damma, D.; Ettireddy, P.; Reddy, B.; Smirniotis, P. A review of low temperature NH3-SCR for removal of NOx. Catalysts 2019, 9, 349. [Google Scholar] [CrossRef]
  40. Zhang, N.; Li, L.; Guo, Y.; He, J.; Wu, R.; Song, L.; Zhang, G.; Zhao, J.; Wang, D.; 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]
  41. Wang, F.; Li, S.; You, R.; Han, Z.; Yuan, W.; Zhu, B.; Gao, Y.; Yang, H.; Wang, Y. Unraveling the synergy between MnOx and CeO2 in MnOx-CeO2 SCR catalysts based on experimental and DFT studies. Appl. Surf. Sci. 2023, 638, 158124. [Google Scholar] [CrossRef]
  42. Nelson, N.; Andana, T.; Rappé, K.; Wang, Y. Mechanistic insight into low temperature SCR by ceria–manganese mixed oxides incorporated into zeolites. Catal. Sci. Technol. 2023, 13, 1111–1118. [Google Scholar] [CrossRef]
  43. Chen, M.; Li, J.; Xue, W.; Wang, S.; Han, J.; Wei, Y.; Mei, D.; Li, Y.; Yu, J. Unveiling secondary-ion-promoted catalytic properties of Cu-SSZ-13 zeolites for selective catalytic reduction of NOx. J. Am. Chem. Soc. 2022, 144, 12816–12824. [Google Scholar] [CrossRef] [PubMed]
  44. Wan, J.; Chen, J.; Zhao, R.; Zhou, R. One-pot synthesis of Fe/Cu-SSZ-13 catalyst and its highly efficient performance for the selective catalytic reduction of nitrogen oxide with ammonia. J. Environ. Sci. 2021, 100, 306–316. [Google Scholar] [CrossRef] [PubMed]
  45. Li, T.; Li, W.; Wey, M. Cation exchange reaction drives CeMn-montmorillonite catalyst with high dispersion and abundant acidic sites for low-temperature NH3-SCR. J. Environ. Chem. Eng. 2023, 11, 109518. [Google Scholar] [CrossRef]
  46. Yang, L.; Wang, P.; Yao, L.; Meng, X.; Jia, C.; Jiang, X.; Jiang, W. Copper doping promotion on Ce/CAC-CNT catalysts with high sulfur dioxide tolerance for low-temperature NH3-SCR. ACS Sustain. Chem. Eng. 2021, 9, 987–997. [Google Scholar] [CrossRef]
  47. Huang, X.; Dong, F.; Zhang, G.; Tang, Z. Constructing TiO2@CeMnOx nanocages by self-sacrificial hydrolytic etching MIL-125 for efficient wide-temperature selective catalytic reduction of nitrogen oxides. Chem. Eng. J. 2022, 432, 134236. [Google Scholar] [CrossRef]
  48. Zhang, N.; Tong, J.; Miyazaki, S.; Zhao, S.; Kubota, H.; Jing, Y.; Mine, S.; Toyao, T.; Shimizu, K. Mechanism of NH3-SCR over P/CeO2 catalysts investigated by operando spectroscopies. Environ. Sci. Technol. 2023, 57, 16289–16295. [Google Scholar] [CrossRef]
  49. Venkataswamy, P.; Rao, K.; Jampaiah, D.; Reddy, B. Nanostructured manganese doped ceria solid solutions for CO oxidation at lower temperatures. Appl. Catal. B-Environ. 2015, 162, 122–132. [Google Scholar] [CrossRef]
  50. Chen, L.; Yang, J.; Ren, S.; Chen, Z.; Zhou, Y.; Liu, W. Effects of Sm modification on biochar supported Mn oxide catalysts for low-temperature NH3-SCR of NO. J. Energy Inst. 2021, 98, 234–243. [Google Scholar] [CrossRef]
  51. Chen, M.; Zhao, W.; Wei, Y.; Han, J.; Li, J.; Sun, C.; Mei, D.; Yu, J. La ions-enhanced NH3-SCR performance over Cu-SSZ-13 catalysts. Nano Res. 2023, 16, 12126–12133. [Google Scholar] [CrossRef]
  52. Gevers, L.; Enakonda, L.; Shahid, A.; Ould-Chikh, S.; Silva, C.; Paalanen, P.; Aguilar-Tapia, A.; Hazemann, J.; Hedhili, M.; Wen, F.; et al. Unraveling the structure and role of Mn and Ce for NOx reduction in application-relevant catalysts. Nat. Commun. 2022, 13, 2960. [Google Scholar] [CrossRef] [PubMed]
  53. Salazar, M.; Becker, R.; Grünert, W. Hybrid catalysts–an innovative route to improve catalyst performance in the selective catalytic reduction of NO by NH3. Appl. Catal. B-Environ. 2015, 165, 316–327. [Google Scholar] [CrossRef]
  54. Zhang, N.; Qian, Y.; Toyao, T.; Shimizu, K. Continuous unsteady-state de-NOx system via tandem water–gas shift, NH3 synthesis, and NH3–SCR under periodic lean/rich conditions. Environ. Sci. Technol. 2023, 57, 19584–19593. [Google Scholar] [CrossRef] [PubMed]
  55. Zhu, N.; Shan, Y.; Shan, W.; Lian, Z.; Du, J.; He, H. Reaction pathways of standard and fast selective catalytic reduction over Cu-SSZ-39. Environ. Sci. Technol. 2021, 55, 16175–16183. [Google Scholar] [CrossRef]
  56. Du, J.; Wang, J.; Shi, X.; Shan, Y.; Zhang, Y.; He, H. Promoting effect of Mn on in situ synthesized Cu-SSZ-13 for NH3-SCR. Catalysts 2020, 10, 1375. [Google Scholar] [CrossRef]
  57. Xin, Y.; Cheng, L.; Lv, Y.; Jia, J.; Han, D.; Zhang, N.; Wang, J.; Zhang, Z.; Cao, X. Experimental and theoretical insight into the facet-dependent mechanisms of NO oxidation catalyzed by structurally diverse Mn2O3 nanocrystals. ACS Catal. 2022, 12, 397–410. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples.
Figure 1. XRD patterns of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples.
Catalysts 14 00010 g001
Figure 2. SEM images of ((a1,a2),(e1,e2)) Cu-SSZ-39, ((b1,b2),(f1,f2)) CeO2/Cu-SSZ-39, ((c1,c2),(g1,g2)) CeMnOx/Cu-SSZ-39, and ((d1,d2),(h1,h2)) MnOx/Cu-SSZ-39 catalysts: ((a1,a2)–(d1,d2)) fresh samples and ((e1,e2)–(h1,h2)) hydrothermally aged samples.
Figure 2. SEM images of ((a1,a2),(e1,e2)) Cu-SSZ-39, ((b1,b2),(f1,f2)) CeO2/Cu-SSZ-39, ((c1,c2),(g1,g2)) CeMnOx/Cu-SSZ-39, and ((d1,d2),(h1,h2)) MnOx/Cu-SSZ-39 catalysts: ((a1,a2)–(d1,d2)) fresh samples and ((e1,e2)–(h1,h2)) hydrothermally aged samples.
Catalysts 14 00010 g002
Figure 3. NOx conversion of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples.
Figure 3. NOx conversion of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples.
Catalysts 14 00010 g003
Figure 4. XPS spectra of Cu 2p3/2 (a1,a2) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve); O 1s (b1,b2) (black line was assigned to the original data, dark yellow line was assigned to the fitting curve); Ce 3d (c) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve); and Mn 2p3/2 (d) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve) of the catalysts.
Figure 4. XPS spectra of Cu 2p3/2 (a1,a2) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve); O 1s (b1,b2) (black line was assigned to the original data, dark yellow line was assigned to the fitting curve); Ce 3d (c) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve); and Mn 2p3/2 (d) (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve) of the catalysts.
Catalysts 14 00010 g004
Figure 5. H2-TPR profiles of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples.
Figure 5. H2-TPR profiles of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples.
Catalysts 14 00010 g005
Figure 6. NH3-TPD curves of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve, red fitting peaks belonged to weak acidity, blue fitting peaks belonged to medium acidity, and yellow fitting peaks belonged to strong acidity).
Figure 6. NH3-TPD curves of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts: (a) fresh samples and (b) hydrothermally aged samples (gray line was assigned to the original data, dark yellow line was assigned to the fitting curve, red fitting peaks belonged to weak acidity, blue fitting peaks belonged to medium acidity, and yellow fitting peaks belonged to strong acidity).
Catalysts 14 00010 g006
Figure 7. 29Si MAS NMR (a) and 27Al MAS NMR (b) spectra with curve fittings of Cu-SSZ-39, CeMnOx/Cu-SSZ-39, Cu-SSZ-39-A, and CeMnOx/Cu-SSZ-39-A catalysts (gray line was assigned to the original data, other different color lines and dotted lines were assigned to the fitting curves and fitting peaks, respectively, for which the corresponding catalysts were noted in the figure).
Figure 7. 29Si MAS NMR (a) and 27Al MAS NMR (b) spectra with curve fittings of Cu-SSZ-39, CeMnOx/Cu-SSZ-39, Cu-SSZ-39-A, and CeMnOx/Cu-SSZ-39-A catalysts (gray line was assigned to the original data, other different color lines and dotted lines were assigned to the fitting curves and fitting peaks, respectively, for which the corresponding catalysts were noted in the figure).
Catalysts 14 00010 g007
Figure 8. Arrhenius plots of Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts.
Figure 8. Arrhenius plots of Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts.
Catalysts 14 00010 g008
Figure 9. NO conversion (a) and NO2 concentration (b) of Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts in NO oxidation reaction.
Figure 9. NO conversion (a) and NO2 concentration (b) of Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts in NO oxidation reaction.
Catalysts 14 00010 g009
Figure 10. NO (a) and NO2 (b) conversion under fast SCR conditions in Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts.
Figure 10. NO (a) and NO2 (b) conversion under fast SCR conditions in Cu-SSZ-39 and CeMnOx/Cu-SSZ-39 catalysts.
Catalysts 14 00010 g010
Table 1. Textural properties and quantitative analysis of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts before and after hydrothermal aging.
Table 1. Textural properties and quantitative analysis of Cu-SSZ-39, CeO2/Cu-SSZ-39, CeMnOx/Cu-SSZ-39, and MnOx/Cu-SSZ-39 catalysts before and after hydrothermal aging.
SamplesSurface Area (m2∙g−1)Pore Volume
(cm3∙g−1)
Pore Size
(nm)
Si/Al
Ratio a
Cu a
(wt.%)
Cu-SSZ-39664.80.281.668.71.6
CeO2/Cu-SSZ-39560.50.271.918.81.5
CeMnOx/Cu-SSZ-39568.00.261.848.41.6
MnOx/Cu-SSZ-39585.10.241.668.61.6
Cu-SSZ-39-A613.80.261.708.51.6
CeO2/Cu-SSZ-39-A555.90.271.918.61.5
CeMnOx/Cu-SSZ-39-A547.50.241.748.61.7
MnOx/Cu-SSZ-39-A551.80.221.578.41.7
a Derived from EDS mappings.
Table 2. Cu 2p3/2, O 1s, Ce 3d, and Mn 2p3/2 XPS analysis of the catalysts.
Table 2. Cu 2p3/2, O 1s, Ce 3d, and Mn 2p3/2 XPS analysis of the catalysts.
SampleCu (%)O (%)Ce (%)Mn (%)
Cu2+CuOOsurOadsOlatCe3+Ce4+Mn2+Mn3+Mn4+
Cu-SSZ-3964.635.431.343.924.8-----
CeO2/Cu-SSZ-39643617.354.628.118.781.3---
CeMnOx/Cu-SSZ-39712918.76120.322.277.825.33341.7
MnOx/Cu-SSZ-3963.636.41852.829.2--30.534.535
Cu-SSZ-39-A38.861.230.536.333.2-----
CeO2/Cu-SSZ-39-A42.457.616.651.831.618.581.5---
CeMnOx/Cu-SSZ-39-A48.551.516.658.724.721.678.42735.137.9
MnOx/Cu-SSZ-39-A44.655.417.647.435--34.539.126.4
Table 3. Quantification of NH3 desorption amount based on NH3-TPD results.
Table 3. Quantification of NH3 desorption amount based on NH3-TPD results.
SampleNH3 Desorption Amount (μmol·gcat−1)
Weak (Peak α)Medium (Peak β)Strong (Peak γ)Total
Cu-SSZ-3955.3110.6624.4790.3
CeO2/Cu-SSZ-3946.1138.1473.7657.9
CeMnOx/Cu-SSZ-3960.5204.2491.6756.3
MnOx/Cu-SSZ-39153.1145.7429.9728.7
Cu-SSZ-39-A18.6121.0170.6310.2
CeO2/Cu-SSZ-39-A26.4201.9227.5455.8
CeMnOx/Cu-SSZ-39-A34.5222.7236.1493.3
MnOx/Cu-SSZ-39-A87.8294.0177.2559.0
Table 4. The deconvolution results of 29Si and 27Al MAS NMR spectra.
Table 4. The deconvolution results of 29Si and 27Al MAS NMR spectra.
SampleSi/Al RatioProportion of Si (nSi, mAl) (%)Tetra-Coordinated AlPenta-Coordinated AlHexa-Coordinated Al
Si (4Si)Si (3Si, 1Al)Si (2Si, 2Al)
Cu-SSZ-399.363.430.46.291.7-8.3
CeMnOx/Cu-SSZ-399.162.5316.596.2-3.8
Cu-SSZ-39-A25.387.78.83.567.122.810.1
CeMnOx/Cu-SSZ-39-A16.581.213.35.587.710.51.8
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tang, A.; Yang, F.; Xin, Y.; Zhu, X.; Yu, L.; Liu, S.; Han, D.; Jia, J.; Lu, Y.; Li, Z.; et al. Enhanced Low-Temperature Activity and Hydrothermal Stability of Ce-Mn Oxide-Modified Cu-SSZ-39 Catalysts for NH3-SCR of NOx. Catalysts 2024, 14, 10. https://doi.org/10.3390/catal14010010

AMA Style

Tang A, Yang F, Xin Y, Zhu X, Yu L, Liu S, Han D, Jia J, Lu Y, Li Z, et al. Enhanced Low-Temperature Activity and Hydrothermal Stability of Ce-Mn Oxide-Modified Cu-SSZ-39 Catalysts for NH3-SCR of NOx. Catalysts. 2024; 14(1):10. https://doi.org/10.3390/catal14010010

Chicago/Turabian Style

Tang, Ahui, Fuzhen Yang, Ying Xin, Xiaoli Zhu, Long Yu, Shuai Liu, Dongxu Han, Junxiu Jia, Yaning Lu, Zhenguo Li, and et al. 2024. "Enhanced Low-Temperature Activity and Hydrothermal Stability of Ce-Mn Oxide-Modified Cu-SSZ-39 Catalysts for NH3-SCR of NOx" Catalysts 14, no. 1: 10. https://doi.org/10.3390/catal14010010

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