Next Article in Journal
Egg-Shell-Type MgAl2O4 Pellet Catalyst for Steam Methane Reforming Reaction Activity: Effect of Pellet Preparation Temperature
Next Article in Special Issue
Development of Quinary Layered Double Hydroxide-Derived High-Entropy Oxides for Toluene Catalytic Removal
Previous Article in Journal
Synthesis of Oxygenated Hydrocarbons from Ethanol over Sulfided KCoMo-Based Catalysts: Influence of Novel Fiber- and Powder-Activated Carbon Supports
Previous Article in Special Issue
Enhanced Catalytic Performance and Sulfur Dioxide Resistance of Reduced Graphene Oxide-Promoted MnO2 Nanorods-Supported Pt Nanoparticles for Benzene Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Sulfur Deactivation and Regeneration over Cu-CHA Zeolite Catalyst

1
State Key Laboratory of New Technologies for Comprehensive Utilization of Rare and Precious Metals, Kunming Institute of Precious Metals, Kunming 650106, China
2
Kunming Sino-Platinum Metals Catalyst Co., Ltd., Kunming 650106, China
3
Sino-Platinum Metals Catalyst (Dongying) Co., Ltd., Dongying 257000, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1499; https://doi.org/10.3390/catal12121499
Submission received: 1 October 2022 / Revised: 10 November 2022 / Accepted: 19 November 2022 / Published: 23 November 2022
(This article belongs to the Special Issue Exclusive Papers in Environmentally Friendly Catalysis in China)

Abstract

:
Benefiting from the exceptional selective catalytic reduction of NOx with ammonia (NH3-SCR) activity, excellent N2 selectivity, and superior hydrothermal durability, the Cu2+-exchanged zeolite catalyst with a chabazite structure (Cu-CHA) has been considered the predominant SCR catalyst in nitrogen oxide (NOx) abatement. However, sulfur poisoning remains one of the most significant deterrents to the catalyst in real applications. This review summarizes the NH3-SCR reaction mechanism on Cu-CHA, including the active sites and the nature of hydrothermal aging resistance. On the basis of the NH3-SCR reaction mechanism, the review gives a comprehensive summary of sulfate species, sulfate loading, emitted gaseous composition, and the impact of exposure temperature/time on Cu-CHA. The nature of the regeneration of sulfated catalysts is also covered in this review. The review gives a valuable summary of new insights into the matching between the design of NH3-SCR activity and sulfur resistance, highlighting the opportunities and challenges presented by Cu-CHA. Guidance for future sulfur poisoning diagnosis, effective regeneration strategies, and a design for an efficient catalyst for the aftertreatment system (ATS) are proposed to minimize the deterioration of NOx abatement in the future. Finally, we call for more attention to be paid to the effects of PO43- and metal co-cations with sulfur in the ATS.

1. Introduction

With the aggravation of the global energy crisis, lean-burn combustion diesel engines featuring strong power and economical merits have been widely used in road and off-road machines. Excessive air intake during engine combustion can not only increase fuel combustion efficiency but also reduce the contents of carbon monoxide (CO), hydrocarbons (HC), and particulate matter (PM) in exhaust emissions [1]. However, lean burn combustion tends to emit excessive NOx and ammonia (NH3), which cause damage to the environment and human health. In recent years, stringent regulations have been formulated to control exhaust emissions in various countries and regions. In order to comply with these stringent regulations, the ATS, as shown in Figure 1, consisting of a diesel oxidation catalyst (DOC), a diesel particulate filter (DPF), ammonia selective catalytic reduction (NH3-SCR), and an ammonia slip catalyst (ASC), has been successfully developed as a promising technology for NOx emission control [1,2,3,4]. The implemented CHINA VI emission regulation requires an 80% reduction in NOx and a 40% increase in durability compared with CHINA V in heavy-duty vehicles [4,5]. Therefore, advanced ATS catalysts shall have better hydrothermal aging resistance and NOx abatement ability, especially during the cold start (below 200 °C) and DPF active regeneration operations (above 650 °C) [5,6]. As can be seen, during DPF active regeneration, excessive fuel is injected into the upstream DOC for soot combustion. This requires the SCR catalysts to have superior stability to withstand HC adsorption and HTA resistance in humid and high-temperature environments [6].
Recently, Cu-CHA has been reported as one of the most promising commercial catalysts in the current emission control of diesel engines [6,7,8]. CHA zeolites are a group of small-pore structures (pore size of ~3.8 Å) that consist of connected TO4 (T represents the framework atom) tetrahedral units by sharing oxygen atoms [6]. Because of the advantages of excellent shape selectivity, hydrothermal aging (HTA) resistance, adjustable acid sites, and a huge specific surface, Cu-SSZ-13 has been widely used in the ATS [6,7].
Rapidly expanding regulations have been enabled by the ultra-low sulfur content in fuel [7]. A drastic reduction in sulfur content has been achieved in diesel engines [7,8]. However, even with the ultra-low sulfur content (10 ppm) under CHINA Ⅵ regulations, sulfur poisoning remains one of the most significant deterrents to the catalyst in real applications [9,10]. Compared with Fe-zeolite, Cu-zeolite is more sensitive to sulfur, which has been proven by many groups [9,10,11]. Meanwhile, accumulated lifetime exposure to sulfur in the diesel ATS always results in a dramatic deterioration of NOx conversion [12,13,14].
Numerous studies have been carried out to explore the damaging effects of sulfur on Cu-CHA structures and the nature of activity recovery during DPF active regeneration [12]. It is generally considered that the deactivation over Cu-CHA is caused by damage to zeolite structures, including the drastic decrease in the specific surface area and the binding of SOx to copper species [15]. In addition, SO2 and NH3 tend to adsorb together to form sulfate species on isolated Cu2+, which further aggravates the pore block and SCR reaction at low temperatures [14,15]. Gao et al. [15] proved that with an increase in SO2 concentration and sulfur exposure time, more sulfate formed and NOx conversion worsened for Cu-SAPO-34. Meanwhile, the active site of Cu(OH)+ can also form copper sulfate species under the impact of SO2. CuO species could further promote the oxidation of SO2 and thus accelerate sulfur poisoning [15,16].
Currently, there are no on-board diagnostics (OBD) specified for sulfur poisoning and regeneration in real applications, resulting in massive parts of failure caused by sulfur poisoning/regeneration. It is essential to explore the nature of sulfur poisoning and regeneration over Cu-CHA, thereby providing guidance for future sulfur poisoning diagnosis, effective regeneration strategies, and a design for an efficient catalyst for the ATS to minimize NOx deterioration in applications.

2. Research on Copper’s Active Sites and NH3-SCR Reaction over Cu-CHA

A major breakthrough was achieved several years ago: researchers discovered that Cu-CHA (e.g., Cu-SAPO-34 and Cu-SSZ-13) could meet the HTA resistance and low/high-temperature activity for commercial utilization [15]. Both catalysts were used in the ATS of diesel vehicles before 2011. However, Cu-SAPO-34 experienced unexpected failure in diesel after-treatment commercial utilization due to a lack of durability at low temperatures upon contact with moisture [16]. Wang et al. [16] studied the deactivation mechanism of Cu-SAPO-34 at an atomic level. It was proposed that Cu-SAPO-34 was deactivated by the formation of Cu-aluminate-like species that were derived from the interaction between Cu(OH)2 and the hydrolysis of framework Al species. Compared with Cu-SAPO-34, Cu-SSZ-13 was reported to have better HTA resistance and predominant NO conversion efficiency [17]. Studies further verified the hydrothermal durability of Cu-SSZ-13 and found nearly no deterioration in comparison with the activity of fresh samples [18].
In principle, there are four most probable ion exchange sites on Cu-SSZ-13, i.e., in the center of double-6-membered rings, inside the CHA cage near the face of the six-membered ring, and in the CHA cage along the three-dimensional axis and near the eight-membered ring [19]. On the basis of previous studies, Gao et al. [20] calculated and analyzed the possible locations of Cu2+ active sites on Cu-SSZ-13 by electron paramagnetic resonance (EPR) and other measurements, as shown in Figure 2. The results revealed that Cu2+ was able to migrate within different locations and even within unit cells. When the ion exchange rate was lower than 23%, which was normally accompanied by a low Cu loading and a Cu-Cu spacing greater than 20 Å, Cu2+ occupied a six-square cell and resided in a six-membered ring. With an increase in the ion exchange rate, Cu loading became relatively higher. Cu2+ resided in the six-membered ring, and the other resided near the eight-membered ring in the cage. When the Cu-Cu spacing narrowed to 4 Å–5 Å, the Cu2+ was supposed to be located in sites A and B (Figure 2) or the hexagonal cell with a distribution of mirror symmetry.

2.1. NH3-SCR Reaction Mechanism over the Cu-CHA Reaction

Many researchers have been devoted to studying the SCR reaction cycle over Cu/SSZ-13, including the oxidation and reduction half-cycle, as shown in Figure 3a [21]. It has been reported that NH3 can react with [CuII(OH)]+ to form a [(NH3)2-CuII-OH]+ complex that migrates into the CHA cage. NO adsorbed on CuII and was oxidized to HONO, resulting in CuII reduction to CuI. In the oxidation half-cycle at low temperatures (below 250 °C), CuI and CuII participated in the catalytic reaction in (the formation of pairing with CuI(NH3)2). It migrated to another CuI(NH3)2 and was activated by O2 to produce a CuI(NH3)2-O2-[CuI(NH3)2 dimer, and it was calculated by density functional theory (DFT) that O2 activation was the rate-determining step of the NH3-SCR reaction on Cu-SSZ-13 [21,22]. Palucci et al. calculated the diffusion radius of Cu(NH3)2+ and found that it could migrate through the 8MR CHA [21]. More recently, Negri et al. discovered the formation of a [Cu2(NH3)4O2]2+ (Figure 3b) complex with a side-on formation of a side-group (μ-η2, η2-peroxodiiamino dicopper) structure by X-ray absorption spectroscopy (EXAFS) and other measurements [22]. Meanwhile, their findings revealed that [Cu2(NH3)4O2]2+ could be completely reduced to [CuI(NH3)2]+ when NO and NH3 appeared in the mixture [22]. These findings indicated a possible low-temperature SCR reaction mechanism of these complexes with NO. When the reaction temperature increased, isolated Cu ions species were anchored at the ion exchange site due to the decomposition of Cu(NH3)n species [23,24,25,26]. Gao and his coworkers calculated that the activation energies of standard SCR at high and low temperatures were about 140 kJ/mol and 70 kJ/mol, respectively [5,23]. Janessens et al. [24] proposed the reaction mechanism of standard SCR and fast SCR, which was more like the reaction cycle at high temperatures, as shown in Figure 3c. According to this mechanism, the mixture of NO and O2 was oxidized on the catalyst surface and subsequently reduced in the atmosphere of NO and NH3. Contact of Cu+ with NO2 or with a mixture of NO and O2 could form the same dicoordinated (Cu-NO3). In the presence of NH3 and NO, Cu2+ was reduced to Cu+ and formed a product of [CuI (NH3)2]+, which was a linear coordination of Cu+ with NH3.

2.2. HTA Resistance over Cu-CHA

Due to the humid and high-temperature atmosphere in the ATS, researchers have been devoted to studying the HTA deactivation mechanism over Cu-CHA. Kwak [27] carried out studies on the HTA resistance over Cu-SSZ-13 and found that the lifetime exposure to exhaust emitted from diesel engines had little impact on its NH3-SCR activity. Moreover, it has been reported that two types of active sites exist in Cu-SSZ-13, namely, the isolated Cu2+ that resides in the six-membered ring and the hydroxylated [CuOH]+ that resides in the eight-membered ring site [23,25,27]. Except for the Cu active sites, Brønsted acid sites functioning as NH3 storage has also been reported as the rate-determining factor in NH3-SCR of Cu-SSZ-13. Gao et al. comprehensively studied the hydrothermal aging mechanism over Cu/SSZ-13 [25]. It was reported that dealumination occurred and Al(OH)3 formed, as shown in Figure 4a [25]. Moisture attacked the Si–OH–Al, resulting in the break of the Al–O bond. A further attack could completely dislodge the framework Al to form Al(OH)3 outside of the framework. The dissociation of Al led to the absence of ion exchange sites and the unstable existence of Cu2+ ions on the exchange sites. Cu aggregated to form CuOx species under high temperatures, as shown in Figure 4b [26]. Song et al. [26] found that [CuOH]+ had lower hydrothermal stability than Cu2+ at high temperatures, which indicated that Cu species residing in the eight-membered ring were more likely to aggregate to form Cu(OH)2. Furthermore, Cu(OH)2 could travel through the channels of Cu/SSZ-13, resulting in further destruction of the pore structure on the catalyst [24,26,27].

3. Sulfur Poisoning and Regeneration over Cu/CHA

The sulfur poisoning over Cu-CHA is closely related to the NH3-SCR mechanism. Meanwhile, the poisoning process is also closely related to the parameters of catalyst status, Cu loading, HTA, sulfurizing atmosphere (SO2, SO3, SOx, H2O, and NH3), sulfur exposure time, exposure temperature, evaluation conditions, etc. [28,29].

3.1. SO2 Poisoning over Cu/CHA

Great efforts have been made to study the impact of sulfur on catalysts. It has been proven by many groups that SO2 severely inhibits the NH3-SCR activity over Cu-CHA at low temperatures, while the high-temperature performance is nearly inhibited [25,26,27,28,29,30]. Wang et al. [30] verified that when Cu-SAPO-34 was exposed in SOx with different gaseous compositions, the NOx conversion at low temperatures was severely inhibited (100~300 °C) in a standard SCR reaction, as shown in Figure 5a. Furthermore, the TOF results showed that all the samples exhibited identical calculation values, which indicated that different SOx composition poisoning on Cu/SAPO-34 were all caused by the reduction of Cu2+ quantity in the kinetic temperature range. In addition, for the remaining Cu2+ that was not poisoned by SOx, the activity remained unchanged [30].
Su [31] further compared the sulfate species on Cu-SAPO-34 and Cu-SSZ-13 and found that three main sulfates formed on zeolite, including H2SO4, CuSO4, and Al2(SO4)3. With stronger oxidation, more sulfates could be loaded on Cu-SSZ-13 compared with Cu-SAPO-34. By contrast, Cu-SSZ-13 and Cu-SAPO-34 tended to generate H2SO4 and Al2(SO4)3, respectively. H2SO4 could be decomposed at a lower temperature than Al2(SO4)3, resulting in the complete regeneration of Cu-SSZ-13 at 450 °C [30,31].
When evaluating the NH3-SCR performance of Cu-SSZ-13 on a synthetic gas bench lower than 350 °C, the catalytic activity of sulfated samples was significantly decreased, indicating that most of the active sites were impacted by sulfur. When the reaction temperature reached beyond 350 °C, the impact of sulfur poisoning gradually decreased, which indicated the removal of sulfur substances and the recovery of active sites [31,32]. Luo et al. [32] revealed that both the Cu2+ active sites coordinating with the six- and eight-membered rings were significantly reduced after sulfur exposure using an analysis of diffuse reflectance infrared Fourier transform spectroscopy (DRIFTs). The isolated Cu2+ was supposed to react with sulfur, which resulted in a decrease in active sites and NOx conversion. Jangjou [33] reported that the decreased activity of Cu-SSZ-13 after sulfur exposure was related to the Cu-S species generated on the active sites. In addition, compared with isolated Cu2+, active Cu (OH)+ on the Cu/SSZ-13 framework was more likely to react with SO2 to form sulfate. The sulfur exposure time, sulfur concentration, and exposure temperature of CHA were investigated by Ren et al. [34]. It was found that the sulfur loadings in the catalyst increased with an extension of the sulfur exposure time, an increase in SO2 content, and a decrease in exposure temperature. The relationship between sulfur content with a specific surface and the activity was investigated. Combined with thermo-gravimetric analysis (TGA) results, it was pointed out that pore blocks by ammonium sulfate and a decrease in the specific surface area on the framework directly led to deteriorated NH3-SCR activity. Previous studies [33,34,35] observed SO2 had a higher selective reactivity with CuII species when it coordinated with NH3 and extra-framework oxygen, in particular for [CuII2(NH3)4O2]2+, by X-ray adsorption spectroscopy (XAS), as shown in Figure 6. By contrast, SO2 was less reactive with Cu species in the absence of NH3 or [CuII2(NH3)4O2]2+, resulting in less NH3-SCR deactivation under low temperatures [35].
In a typical NH3-SCR reaction, urea injection can be identified in the downstream sulfur atmosphere, and the co-adsorption of ammonia sulfate and Cu-S substance must be considered. Jangjou [33] found that when feeding gas containing a mixture of SO2, NH3, and NO, NH3 facilitated SO2 adsorption on Cu-SAPO-34 at 200 °C. Sulfur was stored on zeolite and co-adsorbed with NH3, which resulted in the formation of ammonium sulfate. On the one hand, the DRIFT results showed that in the absence of NH3, Cu2+ resided at the six-membered ring and was completely poisoned by sulfur and partially poisoned at the eight-membered ring in a typical NO adsorption reaction [35,36]. On the other hand, the co-adsorbed ammonium sulfate could be consumed by NO in the SCR reaction [36]. Shen [37] proved that no matter which sulfate was generated at the active sites, the decrease in the active sites of isolated Cu2+ mainly accounted for the decrease in activity. Wijayanti et al. [38] studied the sulfate species on Cu-CHA and found that (NH4)2SO4 and CuSO4 formed at 200 and 400 °C, respectively, on the isolated Cu2+ active sites. Both led to deactivation over Cu-CHA [39,40]. It was proven that two kinds of active Cu species exist on Cu/CHA, namely Cu2+-2Z and [Cu (OH)+]-Z [40,41,42,43]. As shown in Figure 7, Jangjou [33] pointed out that Cu2+-2Z had a different low-temperature SO2 poisoning mechanism compared with [Cu (OH)+]-Z. Cu2+-2Z active sites were less reactive with SO2. However, when SO2 and NH3 co-existed in a typical SCR reaction, NH3 and SO2 could react at a low temperature and decompose below 550 °C. On the contrary, SO2 reactively adsorbed with [Cu (OH)+]-Z and formed CuHSO3, which could be composed below 580 °C. Above that, CuHSO3 was oxidized to CuHSO4 with gaseous oxygen feeding. Meanwhile, CuHSO4 required a relatively high decomposition temperature, i.e., 750 °C, according to DFT calculation and FTIR measurement. In addition, this mechanism was further proven by Hannershoi [44] through quantitative identification of EPR. Olsson et al. proposed a kinetic model of SO2 poisoning and regeneration in which the copper resided in the six-membered rings (S1Cu), and copper in the larger CHA cages (S2 and S3) adsorbed SO2 and formed S1Cu-SO2 and S2-SO2. In addition, NH3 was adsorbed on those sites and formed S1Cu-SO2-(NH3)2 and S2-SO2-(NH3)2 complexes, which worked as precursors of (NH4)2SO4 [45].

3.2. SO3 Poisoning over Cu/CHA

In a typical ATS, a substantial percentage of SO2 in the diesel exhausts can be oxidized by DOC. The impact of an upstream DOC catalyst on Cu-SSZ-13 was investigated by Kumar [46]. DOC was used for oxidizing CO/HC/soot and generating heat for the downstream DPF for the passive/active soot oxidization and fast reaction on SCR. During DOC oxidation, SO2 tended to be oxidized into SOx, especially SO3. Different sulfur species were added to the synthetic gas bench to study their influence on the NH3-SCR activity of Cu-SAPO-34. The results showed that, compared with SO2, SO3 severely damaged the NH3-SCR activity. Kumar also found that more copper sulfate could be generated under SO3. Sulfated samples could not recover to the initial status after regeneration, inferring the formation of sulfate on Cu-SAPO-34. Wang et al. [30] pointed out that Cu-SSZ-13 can also oxidize SO2 into SO3 due to its strong oxidation, which can lead to the formation of H2SO4, CuSO4, and Al2 (SO4)3 in a typical NH3-SCR reaction. Kumar et al. [46] quantified the gaseous sulfur species on Cu-CHA zeolite and found that the NH3-SCR performance was relevant to the sulfur loading and quantification. The FTIR measurement reported SO3 transformation into H2SO4 in the H2O gaseous feeding. Cheng [47] evaluated the influence of SO2 and SO3 on Cu-zeolite and found that SO3 had a more severe deterioration effect on NH3-SCR performance. With identical sulfur content adsorption, it was easier for SO3 to generate more sulfate. X-ray near edge spectroscopy (XANES) and X-ray photoelectron spectroscopy (XPS) techniques revealed that more CuSO4 was generated under SO3 exposure at the active sites compared with SO2. SO3 could result in a decrease in activity and could be recovered by the transformation of CuSO4 into isolated Cu2+ [48,49]. H2O also impacted sulfate formation on Cu-SSZ-13. Massive CuSO4 was formed on Cu-SSZ-13 compared with moisture-free conditions in the feeding gas. Therefore, the SOx could react with single Al-coordination Z-Cu-OH without H2O. By contrast, SOx could also react with dual Al coordination Z2-Cu with moisture that produced more copper sulfate, as shown in Equations (1) and (2) [41,49,50]. Shan’s study also proved that SO2 could dislodge the extra-framework Al that resulted from the dealumination during the hydrothermal aging process. More Cu2+ active sites formed as CuOx after high-temperature sulfurization when compared with samples with only hydrothermal aging [50].
Z Cu OH + SO X     Z Cu HSO X + 1
Z 2 Cu + SO X + H 2 O     Z OH HSO X + 1 + Z H
It is worth noting that NO2/NOx also has a significant impact on NH3-SCR activity after exposure to sulfur. Compared with FSCR and slow SCR reactions, sulfated Cu/CHA experienced a drastic drop in NOx conversion in the standard SCR reaction but no obvious decrease in fast SCR and slow SCR reactions [28,29,51,52]. Wijayanti et al. [29] found that (NH4)2SO4 and CuSO4 formed under 200 and 400 °C, respectively, in a typical fast SCR reaction. In addition, sulfur deactivation was also closely related to Cu loading, Cu/Al, the aging degree, H2O gaseous feeding, etc.
On the basis of these studies, we found that the sulfate species and loadings generated under different sulfur exposure conditions were different and that the deactivation of NH3-SCR activity over Cu/CHA was closely related to the decrease in active sites. Furthermore, hydrothermal aging typically took place when the catalyst was exposed in the humid atmosphere during DPF regeneration. It was pointed out that SO3 can accelerate the dealumination on Cu-SSZ-13 and result in the irreversible collapse of the framework at high temperatures [50,51,52,53,54].

3.3. Sulfur Regeneration over Cu-CHA

The reversible and irreversible deactivation over Cu-CHA after sulfur poisoning has been investigated in recent years. In real applications, Cu-CHA regenerated along with the active regeneration over the DPF catalyst. Sulfates formed on Cu-CHA catalysts can be partially or completely restored during DPF active regeneration.
The regeneration mechanism over Cu-CHA varied with the feeding gas. Zhang et al. [40] reported that when both SO2 and NH3 were co-added into the NH3-TPD, Cu-CHA was found with more ammonia stored. Ammonia sulfate was desorbed below 450 °C, indicating a relatively lower regeneration temperature. Su et al. [31] studied the sulfate regeneration of Cu/SSZ-13 catalyst and found that ammonium sulfide species could be removed at 450 °C, while copper sulfate needed a higher temperature before being completely removed.
Hammershøi et al. [44] studied the regeneration activity for the SO2 poisoning of Cu-SAPO-34 and concluded that its NH3-SCR activity could recover nearly 80% below 550 °C. When heated to 700 °C, it gained complete recovery. Wang et al. [30] reported the recovery mechanism of Cu-SSZ-13 in a wide temperature range. The phenomenon of reversible and irreversible deactivation was similar to that of Cu-SAPO-34. Wang explained the correlation between the regeneration temperature and SCR activity. It was reported that when the generation temperature was below 600 °C, there was a clear trend of irreversible deactivation over Cu-CHA, in a particular activity in a low temperature range [46]. Since CuSO4 was decomposed to abundant CuO, which inhibited the SCR performance, the NH3 oxidation increased at high temperatures [55,56,57]. When the regeneration temperature was increased to 650 °C, the deactivation at low temperatures declined significantly, which accounted for the CuSO4 decomposition and migration to isolated Cu2+ and Cu+ [41,42,43]. These results were in agreement with studies of Shen et al. [55] and Jangjou et al. [33] regarding the recovery mechanism of stable copper sulfate decomposition on the Cu/SSZ-13 catalyst.
All in all, the relatively low SO2 resistance limits the broader application of Cu-CHA in NH3-SCR. Except for the regeneration method reviewed above, it was reported that some reductants, such as NH3, C3H6, and n-C12H26, can achieve the removal of sulfate without heat treatment by changing the oxidation state of Cu [58,59,60]. Kumar et al. [46] reported that 1000 × 10−6 C3H6 can achieve the complete removal of sulfate in the oxidizing atmosphere. It was proposed that recovery is beneficial for the reduction in binding energy between sulfate and Cu. Yu and his coworkers found that the hybrid catalysts incorporated ZnTiOx, which was proposed to work as a sacrificial reaction with SO2, thus preventing the Cu2+ from poisoning by SO2 [61]. Other alkaline elements, such as Ce and Fe modification, can also improve the SO2 resistance over Cu-CHA [59,60,61]. These findings enhance the SO2 tolerance of Cu-CHA to some extent and give guidance to the further design of Cu-CHA in NH3-SCR applications. However, the effect is still limited, and more studies are required in the future.

4. Conclusions

The deactivation and regeneration behavior over Cu-CHA in the NH3-SCR reactions were reviewed in this paper. According to the findings reviewed above, it is obvious that different sulfate species and loading generated under different sulfur exposure conditions can be identified. Sulfur poisoning and regeneration can proceed through different mechanisms depending on the conditions.
The sulfur poisoning that occurs at low temperatures over Cu-CHA by feeding SO2 has an impact on SCR activity. Notably, NH3 and SO2 tend to form ammonium sulfate and coordinate with the isolated Cu2+, which results in pore blocks and the deactivation of activity. Both copper sulfate and ammonia sulfate can impact NH3-SCR activity by reducing the content of Cu2+. In addition, the thermal decomposition temperatures of ammonium sulfate species and copper sulfate are around 550 °C and 700 °C, respectively. Meanwhile, the regeneration process is always accompanied by the redispersion of copper. Therefore, the activity caused by SO2 or ammonium sulfate species could be partially or completely recovered.
SO3 is reported to have more severe deactivation within 200~400 °C compared with SO2, which is mainly attributed to the massive CuSO4 formation. The presence of SO3 could cause dealumination on the framework, resulting in the fracture of the Si–O(H)–Al bond. In addition, SO3 can poison more isolated Cu2+; therefore, the decrease in Cu2+ is the main reason for the decrease in NH3-SCR activity. Furthermore, the decrease in Cu2+ and the fracture of the Si–O(H)–Al bond lead to a drop in high-temperature activity. The formation of copper sulfate requires nearly 700 °C for regeneration. It is always accompanied by the formation of CuO and Cu migration in the cage. However, the active sites cannot completely recover because of the influence of dealumination.
It can be seen that many trials have been carried out to improve the SO2 tolerance over Cu-CHA. The addition of other metals as a sacrificial component is proposed as an efficient way to improve its SO2 tolerance in real applications. Moreover, it is essential to clarify the nature of co-poisoning by alkaline, PO43−, etc. with sulfur in real applications. This could provide a specified regeneration strategy for calibration over the ATS, thus enhancing the durability and NOx abatement of Cu-CHA in industrial applications.

Author Contributions

Conceptualization, J.M. and H.L. investigation, J.M. and F.Y.; writing—review and editing, J.M. and S.C. Chang; supervision, Y.Z.; project administration, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Provincial Key S&T Special Projects of Yunnan (grant number 202102AB080007) and the National Key Research and Development Program of China (grant number 2021YFB35033200).

Conflicts of Interest

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

References

  1. Lee, K.J.; Kumar, P.A.; Maqbool, M.S.; Rao, K.N.; Song, K.H.; Ha, H.P. Ceria added Sb-V2O5/TiO2 catalysts for low temperature NH3-SCR: Physical-chemical properties and catalytic activity. Appl. Catal. B Environ. 2013, 142–143, 705–717. [Google Scholar] [CrossRef]
  2. Zhao, H.W.; Zhao, Y.N.; Liu, M.K.; Li, X.H.; Ma, Y.H.; Yong, X.; Chen, H.; Li, Y.D. Phosphorus modification to improve the hydrothermal stability of a Cu-SSZ-13 catalyst for selective reduction of NOx with NH3. Appl. Catal. B Environ. 2019, 252, 230–239. [Google Scholar] [CrossRef]
  3. Ryu, T.; Kim, H.; Hong, S.B. Nature of active sites in Cu-LTA NH3-SCR catalysts: A comparative study with Cu-SSZ-13. Appl. Catal. B Environ. 2019, 245, 513–521. [Google Scholar] [CrossRef]
  4. Zhao, W.Y.; Li, Z.Q.; Wang, Y.; Fan, R.R.; Zhang, C.; Wang, Y.; Guo, X.; Wang, R.; Zhang, S.L. Ce and Zr Modified WO3-TiO2 Catalysts for Selective Catalytic Reduction of NOx by NH3. Catalysts 2018, 8, 375–381. [Google Scholar] [CrossRef] [Green Version]
  5. Shan, Y.L.; Du, J.P.; Zhang, Y.; Shan, W.P.; Shi, X.Y.; Yu, Y.B.; Zhang, R.D.; Meng, X.J.; Xiao, F.S.; He, H. Selective catalytic reduction of NOx with NH3: Opportunities and challenges of Cu-based small-pore zeolites. NSR 2021, 8, nwab010. [Google Scholar] [CrossRef] [PubMed]
  6. Yahiro, H.; Iwamoto, M. Copper ion-exchanged zeolite catalysts in de-NOx reaction. Appl. Catal. A Gen. 2001, 222, 163–181. [Google Scholar] [CrossRef]
  7. Zones, S.I. Zeolite SSZ-13 and Its Method of Preparation. US Patent 4,544,538, 1985. [Google Scholar]
  8. Ma, L.; Cheng, Y.; Cavataio, G.; McCabe, R.W.; Fu, L.; Li, J. Characterization of commercial Cu-SSZ-13 and Cu-SAPO-34 catalysts with hydrothermal treatment for NH3-SCR of NOx in diesel exhaust. Chem. Eng. J. 2013, 225, 323–330. [Google Scholar] [CrossRef]
  9. Wang, J.C.; Peng, Z.L.; Chen, Y. In-situ hydrothermal synthesis of Cu-SSZ-13/cordierite for the catalytic removal of NOx from diesel vehicles by NH3. Chem. Eng. J. 2015, 263, 9–19. [Google Scholar] [CrossRef]
  10. Xue, J.J.; Wang, X.Q.; Qi, G.S.; Wang, J.; Shen, M.Q.; Li, W. Characterization of copper species over Cu/SAPO-34 in selective catalytic reduction of NOx with ammonia: Relationships between active Cu sites and de-NOx performance at low temperature. J. Catal. 2013, 297, 56–64. [Google Scholar] [CrossRef]
  11. Metkar, P.S.; Harold, M.P.; Balakotaiah, V. Experimental and kinetic modeling study of NH3-SCR of NOx on Fe-ZSM-5, Cu-chabazite and combined Fe- and Cu-zeolite monolithic catalysts. Chem. Eng. Sci. 2013, 87, 51–66. [Google Scholar] [CrossRef]
  12. Cejka, J.; Avelino, C.; Stacey, Z. Zeolites and Catalysis: Synthesis, Reactions and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2010. [Google Scholar]
  13. Wang, J.C.; Peng, Z.L.; Qiao, H.; Han, L.; Bao, W.R.; Chang, L.; Gang, F.; Liu, W. Influence of aging on in situ hydrothermally synthesized Cu-SSZ-13 catalyst for NH3-SCR reaction. R. Soc. Chem. 2014, 4, 42403–42411. [Google Scholar] [CrossRef]
  14. Ciardelli, C.; Nova, I.; Tronconi, E. Reactivity of NO/NO2-NH3 SCR system for diesel exhaust aftertreatment: Identification of the reaction network as a function of temperature and NO2 feed content. Appl. Catal. B Environ. 2007, 70, 80–90. [Google Scholar] [CrossRef]
  15. Gao, F.; Kwak, J.H.; Szanyi, J.; Peden, C.H.F. Current understanding of Cu-exchanged Chabazite molecular sieves for use as commercial diesel engine DeNOx catalysts. Top. Catal. 2013, 56, 1441–1459. [Google Scholar] [CrossRef]
  16. Wang, A.; Chen, Y.; Walter, E.D.; Washton, N.M.; Mei, D.; Varga, T.; Wang, Y.; Szanyi, J.; Wang, Y.; Peden, C.H.F.; et al. Unraveling the mysterious failure of Cu/SAPO-34 selective catalytic reduction catalysts. Nat. Commun. 2019, 10, 1137. [Google Scholar] [CrossRef] [Green Version]
  17. Sappok, A.G.; Wong, V.W. Physical Characterization of Ash Species in Diesel Exhaust Entering After Treatment Systems; SAE International 2007-01-0318; Massachusetts Institute of Technology, Sloan Automotive Laboratory: Cambridge, MA, USA, 2007. [Google Scholar]
  18. Korhonen, S.T.; Fickel, D.W.; Lobo, R.F.; Weckhuysen, B.M.; Beale, A.M. Isolated Cu2+ ions: Active sites for selective catalytic reduction of NO. Chem. Comm. 2011, 47, 800–802. [Google Scholar] [CrossRef] [PubMed]
  19. Fickel, D.W.; D’Addio, E.; Lauterbach, J.A. The ammonia selective catalytic reduction activity of copper-exchanged small-pore zeolites. Appl. Catal. B Environ. 2011, 102, 441–448. [Google Scholar] [CrossRef]
  20. Gao, F.; Walter, E.D.; Karp, E.M. Structure–activity relationships in NH3-SCR over Cu-SSZ-13 as probed by reaction kinetics and EPF studies. J. Catal. 2013, 300, 20–29. [Google Scholar] [CrossRef]
  21. Paolucci, C.; Khurana, I.; Parekh, A.A.; Li, S.; Shih, A.J.; Li, H.; Iorio, J.R.D.; Calallerg, A.; Yezerets, A.; Miller, J.T.; et al. Dynamic multinuclear sites formed by mobilized copper ions in NOx selective catalytic reduction. Science 2017, 357, 898–903. [Google Scholar] [CrossRef] [Green Version]
  22. Negri, C.; Selleri, T.; Borfecchia, E.; Martini, A.; Lomachenko, K.A.; Janssens, T.V.W.; Cutini, M.; Bordiga, S.; Berlier, G. Structure and reactivity of oxygenbridged diamino dicopper (II) complexes n Cu-CHA catalyst for NH3-SCR. J. Am. Chem. Soc. 2020, 142, 15884–15896. [Google Scholar] [CrossRef] [PubMed]
  23. Gao, F.; Walter, E.D.; Kollar, M.; Wang, Y.; Szanyi, J.; Peden, C.H.F. Understanding ammonia selective catalytic reduction kinetics over Cu/SSZ-13 from motion of the Cu ions. J. Catal. 2014, 319, 1–14. [Google Scholar] [CrossRef]
  24. Janssens, T.V.W.; Falsig, H.; Lundegaard, L.F.; Vennestrøm, P.N.R.; Rasmussen, S.B.; Moses, P.G.; Giordanino, F.; Borfecchia, E.; Lomachenko, K.A.; Lamberti, C.; et al. A consistent reaction scheme for the selective catalytic reduction of nitrogen oxides with ammonia. ACS Catal. 2015, 5, 2832–2845. [Google Scholar] [CrossRef] [Green Version]
  25. Gao, F.; Wang, Y.; Washton, N.M.; Kollar, M.; Szanyi, J.; Peden, C.F.F. Effects of alkali and alkaline earth cocations on the activity and hydrothermal stability of Cu/SSZ-13 NH3-SCR Catalysts. ACS Catal. 2015, 5, 6780–6791. [Google Scholar] [CrossRef]
  26. Song, J.; Wang, Y.; Walter, E.D.; Washton, N.M.; Mei, D.; Kovarik, L.; Engelhard, M.H.; Prodinger, S.; Wang, Y.; Peden, C.H.F. Toward rational design of Cu/SSZ-13 selective catalytic reduction catalysts: Implications from atomic-level understanding of hydrothermal stability. ACS Catal. 2017, 7, 8214–8227. [Google Scholar] [CrossRef]
  27. 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]
  28. Dahlin, S.; Lantto, C.; Englund, J.; Bjrn, W.; Skoglundh, M.; Lars, J.P. Chemical aging of Cu-SSZ-13 SCR catalysts for heavy-duty vehicles-Influence of sulfur dioxide. Catal. Today 2019, 320, 72–83. [Google Scholar] [CrossRef]
  29. Wijayanti, K.; Leistner, K.; Chand, S.; Kumar, A.; Kamasamudram, K.; Currier, N.W.; Yezerets, A.; Olsson, L. Deactivation of Cu-SSZ-13 by SO2 exposure under SCR conditions. Catal. Sci. Technol. 2016, 6, 2565–2579. [Google Scholar] [CrossRef]
  30. Wang, C.; Wang, J.; Wang, J.; Yu, T.; Shen, M. The effect of sulfate species on the activity of NH3-SCR over Cu/SAPO-34. Appl. Catal. B Environ. 2017, 204, 239–249. [Google Scholar] [CrossRef]
  31. Su, W.; Li, Z.; Zhang, Y.; Meng, C.; Li, J. Identification of sulfate species and their influence on SCR performance of Cu/CHA catalyst. Catal. Sci. Technol. 2017, 7, 1523–1528. [Google Scholar] [CrossRef]
  32. Luo, J.; Gao, F.; Kamasamudram, K.; Currier, N.; Peden, C.H.F.; Yezerets, A. New insights into Cu/SSZ-13 SCR catalyst acidity. Part I: Nature of acidic sites probed by NH3 titration. J. Catal. 2017, 348, 291–310. [Google Scholar] [CrossRef] [Green Version]
  33. Jangjou, Y.; Do, Q.; Gu, Y.; Lim, L.G.; Sun, H.; Wang, D.; Kumar, A.; Li, J.; Grabow, L.C.; Epling, W. Nature of Cu active centers in Cu-SSZ-13 and their responses to SO2 exposure. ACS Catal. 2018, 8, 1325–1337. [Google Scholar] [CrossRef]
  34. Ren, L.; Zhu, L.F.; Yang, C.G.; Chen, Y.M.; Su, Q.; Zhang, H.Y.; Li, C.; Nawaz, F.; Meng, X.J.; Xiao, F.S. Designed copper-amine complex as an efficient template for one-pot synthesis of Cu-SSZ-13 zeolite with excellent activity for selective catalytic reduction of NOx by NH3. Chem. Commun. 2011, 47, 9789–9791. [Google Scholar] [CrossRef] [PubMed]
  35. Molokova, A.Y.; Borfecchia, E.; Martini, A.; Pankin, I.A.; Atzori, C.; Mathon, O.; Bordiga, S.; Wen, F.; Vennestrøm, P.N.R.; Berlier, G.; et al. SO2 Poisoning of Cu-CHA deNOx Catalyst: The Most Vulnerable Cu Species Identified by X-ray Absorption Spectroscopy. JACS Au 2022, 2, 787–792. [Google Scholar] [CrossRef]
  36. Kirill, A.L.; Borfecchia, E.; Chiara, N.; Gloria, B.; Carlo, L.; Pablo, B.; Hanne, F.; Silvia, B. The Cu-CHA deNOx catalyst in action: Temperature-dependent NH3-SCR monitored by operando X-ray absorption and emission spectroscopy. J. Am. Chem. Soc. 2016, 138, 12025–12028. [Google Scholar]
  37. Shen, M.; Zhang, Y.; Wang, J. Nature of SO3 poisoning on Cu/SAPO-34 SCR catalysts. J. Catal. 2018, 358, 277–286. [Google Scholar] [CrossRef]
  38. Wijayanti, K.; Xie, K.; Kumar, A.; Krishna, K.; Olsson, L. Effect of gas compositions on SO2 poisoning over Cu/SSZ-13 used for NH3-SCR. Appl. Catal. B Environ. 2017, 219, 142–154. [Google Scholar] [CrossRef]
  39. Ming, S.; Pang, L.; Chen, Z.; Guo, Y.; Li, T. Insight into SO2 poisoning over Cu-SAPO-18 used for NH3-SCR. Microporous Mesoporous Mater. 2020, 303, 110294. [Google Scholar] [CrossRef]
  40. Zhang, L.; Wang, D.; Liu, Y.; Krishna. SO2 poisoning impact on the NH3-SCR reaction over a commercial Cu-SAPO-34 SCR catalyst. Appl. Catal. B Environ. 2014, 156, 371–377. [Google Scholar] [CrossRef]
  41. Cheng, Y.; Lambert, C.; Kim, D.H.; Kwak, J.H.; Cho, S.J.; Peden, C.H.F. The different impacts of SO2 and SO3 on Cu/zeolite SCR catalysts. Catal. Today 2010, 151, 266–270. [Google Scholar] [CrossRef]
  42. Liu, X.J.; Li, Y.H.; Zhang, R.R. Ammonia selective catalytic reduction of NO over Ce-Fe/Cu-SSZ-13 catalysts. RSC Adv. 2015, 5, 85453–85459. [Google Scholar] [CrossRef]
  43. Kumar, A.; Smith, M.A.; Kamasamudram, K.; Currier, N.W.; An, H.; Yezerets, A. Impact of different forms of feed sulfur on small-pore Cu-zeolite SCR catalyst. Catal. Today 2014, 231, 75–82. [Google Scholar] [CrossRef]
  44. Hammershøi, P.S.; Godiksen, A.L.; Mossin, S.; Vennestrøm, P.N.R.; Jensen, A.D.; Janssens, T.V.W. Site selective adsorption and relocation of SOx in deactivation of Cu-CHA catalysts for NH3-SCR. React. Chem. Eng. 2019, 4, 1081–1089. [Google Scholar] [CrossRef]
  45. Olsson, L.; Wijayanti, K.; Leistner, K.; Kumar, A.; Joshi, S.Y.; Kamasamudram, K.; Currier, N.W.; Yezerets, A. A kinetic model for sulfur poisoning and regeneration of Cu/SSZ-13 used for NH3-SCR. Appl. Catal. B Environ. 2016, 183, 394–406. [Google Scholar] [CrossRef]
  46. Kumar, M.; Luo, H.; RomáN-Leshkov, Y.; Rimer, J.D. SSZ-13 crystallization by particle attachment and deterministic pathways to crystal size control. J. Am. Chem. Soc. 2015, 137, 13007–13017. [Google Scholar] [CrossRef] [PubMed]
  47. Cheng, Y.; Montreuil, C.; Cavataio, G.; Lambert, C. Sulfur Tolerance and DeSOx Studies on Diesel SCR Catalysts. SAE Int. J. Fuels Lubr. 2008, 1, 471–476. [Google Scholar] [CrossRef]
  48. Hammershøi, P.S.; Jangjou, Y.; Epling, W.S.; Jensen, A.D.; Janssens, T.V.W. Reversible and irreversible deactivation of Cu-CHA NH3-SCR catalysts by SO2 and SO3. Appl. Catal. B Environ. 2018, 226, 38–45. [Google Scholar] [CrossRef] [Green Version]
  49. Li, P.; Yu, F.; Zhu, M.; Tang, C.; Dai, B.; Dong, L. Selective catalytic reduction De-NOx catalysts. Prog. Chem. 2016, 28, 1578–1590. [Google Scholar]
  50. Shan, Y.; Shi, X.; Yan, Z.D.; He, H. Deactivation of Cu-SSZ-13 in the presence of SO2 during hydrothermal aging. Catal. Today 2019, 320, 84–90. [Google Scholar] [CrossRef]
  51. Shan, Y.; Shan, W.; Shi, X.; Du, J.; He, H. A comparative study of the activity and hydrothermal stability of Al-rich Cu-SSZ-39 and Cu-SSZ-13. Appl Catal B 2020, 264, 118511–118520. [Google Scholar] [CrossRef]
  52. Wang, A.; Olsson, L. Insight into the SO2 poisoning mechanism for NOx removal by NH3 SCR over Cu/LTA and Cu/SSZ-13. Chem. Eng. J. 2020, 395, 125048–125059. [Google Scholar] [CrossRef]
  53. Jo, D.; Park, G.T.; Ryu, T.; Hong, S.B. Economical synthesis of high-silica LTA zeolites: A step forward in developing a new commercial NH3-SCR catalyst. Appl. Catal. B 2019, 243, 212–219. [Google Scholar] [CrossRef]
  54. Wang, Y.; Li, Z.; Fan, R.; Guo, X.; Liu, W. Deactivation and Regeneration for the SO2-Poisoning of a Cu-SSZ-13 Catalyst in the NH3-SCR Reaction. Catalysts 2019, 9, 797. [Google Scholar] [CrossRef] [Green Version]
  55. Shen, M.; Wang, Z.; Li, X.; Wang, J.; Wang, J.; Wang, C.; Wang, J. Effects of regeneration conditions on sulfated CuSSZ-13 catalyst for NH3-SCR. Korean J. Chem. Eng. 2019, 36, 1249–1257. [Google Scholar] [CrossRef]
  56. Ando, R.; Banno, Y.; Nagata, M. Detailed Mechanism of S Poisoning and De-Sulfation Treatment of Cu-SCR Catalyst; SAE Technical Paper Series; SAE: Warrendale, PA, USA, 2017. [Google Scholar]
  57. Kumar, A.; Smith, M.; Kamasamudram, K.; Curier, N.W.; Yezerets, A. Chemical deSOx: An effective way to recover Cu-zeolite SCR catalysts from sulfur poisoning. Catal. Today 2016, 267, 10–16. [Google Scholar] [CrossRef]
  58. Mesilov, V.V.; Dahlin, S.; Bergman, S.L.; Xi, S.; Han, J.; Olsson, L.; Pettersson, L.J.; Bernasek, S.L. Regeneration of sulfurpoisonedCu-SSZ-13 catalysts: Copper speciation and catalytic performance evaluation. Appl. Catal. B Environ. 2021, 299, 120626. [Google Scholar] [CrossRef]
  59. Borfecchia, E.; Negri, C.; Lomachenko, K.A.; Lamberti, C.; Janssens, T.V.W.; Berlier, G. Temperature-dependent dynamics of NH3-derived Cu species in the Cu-CHA SCR catalyst. React. Chem. Eng. 2019, 4, 1067–1080. [Google Scholar] [CrossRef]
  60. Sushkevich, V.L.; Safonova, O.V.; Palagin, D.; Newton, M.A.; van Bokhoven, J.A. Structure of copper sites in zeolites examined by Fourier and wavelet transform analysis of EXAFS. Chem. Sci. 2020, 11, 5299–5312. [Google Scholar] [CrossRef]
  61. Yu, R.; Zhao, Z.; Huang, S.; Zhang, W. Cu-SSZ-13 zeolite–metal oxide hybrid catalysts with enhanced SO2-tolerance in the NH3-SCR of NOx. Appl. Catal. B 2020, 269, 118825. [Google Scholar] [CrossRef]
Figure 1. A typical advanced emission control system in diesel vehicles.
Figure 1. A typical advanced emission control system in diesel vehicles.
Catalysts 12 01499 g001
Figure 2. Schematic drawing of the SSZ-13 hexagonal unit cell structure and possible Cu2+ locations (A, A’, B, C, D, E represent the possible Cu2+ locations).
Figure 2. Schematic drawing of the SSZ-13 hexagonal unit cell structure and possible Cu2+ locations (A, A’, B, C, D, E represent the possible Cu2+ locations).
Catalysts 12 01499 g002
Figure 3. (a) Possible low-temperature SCR reaction cycle. Adapted with permission from Ref. [21]. Copyright 2017 Science. (b) μ-η22-peroxo diamino dicopper (II) (side-on). Adapted with permission from Ref. [22]. Copyright 2020 American Chemical Society. (c) Possible high-temperature SCR reaction cycle. Adapted with permission from Ref. [24]. Copyright 2015 ACS Catalysts.
Figure 3. (a) Possible low-temperature SCR reaction cycle. Adapted with permission from Ref. [21]. Copyright 2017 Science. (b) μ-η22-peroxo diamino dicopper (II) (side-on). Adapted with permission from Ref. [22]. Copyright 2020 American Chemical Society. (c) Possible high-temperature SCR reaction cycle. Adapted with permission from Ref. [24]. Copyright 2015 ACS Catalysts.
Catalysts 12 01499 g003
Figure 4. (a) Mechanism of dealumination over SSZ-13. Adapted with permission from Ref. [25]. Copyright 2015 ACS Catalysis. (b) Cu species reaction with water. Adapted with permission from Ref. [26]. Copyright 2017 ACS Catalysis.
Figure 4. (a) Mechanism of dealumination over SSZ-13. Adapted with permission from Ref. [25]. Copyright 2015 ACS Catalysis. (b) Cu species reaction with water. Adapted with permission from Ref. [26]. Copyright 2017 ACS Catalysis.
Catalysts 12 01499 g004
Figure 5. (a) NOx conversion of fresh and sulfated samples over Cu-SSZ-13; gas composition: 500 ppm NO, 500 ppm NH3, 5% O2, 7% CO2, N2. Adapted with permission from Ref. [30]. Copyright 2017 Applied Catalysis B: Environmental. (b) TOFs results over fresh and sulfated catalysts. Adapted with permission from Ref. [30]. Copyright 2017 Applied Catalysis B: Environmental.
Figure 5. (a) NOx conversion of fresh and sulfated samples over Cu-SSZ-13; gas composition: 500 ppm NO, 500 ppm NH3, 5% O2, 7% CO2, N2. Adapted with permission from Ref. [30]. Copyright 2017 Applied Catalysis B: Environmental. (b) TOFs results over fresh and sulfated catalysts. Adapted with permission from Ref. [30]. Copyright 2017 Applied Catalysis B: Environmental.
Catalysts 12 01499 g005
Figure 6. Cu K-edge XAS collected in situ during the exposure of the Cu species obtained to SO2.
Figure 6. Cu K-edge XAS collected in situ during the exposure of the Cu species obtained to SO2.
Catalysts 12 01499 g006
Figure 7. Low-temperature SO2 poisoning mechanism. Adapted with permission from Ref. [33]. Copyright 2018 ACS Catalysis.
Figure 7. Low-temperature SO2 poisoning mechanism. Adapted with permission from Ref. [33]. Copyright 2018 ACS Catalysis.
Catalysts 12 01499 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, J.; Chang, S.; Yu, F.; Lai, H.; Zhao, Y. Research Progress on Sulfur Deactivation and Regeneration over Cu-CHA Zeolite Catalyst. Catalysts 2022, 12, 1499. https://doi.org/10.3390/catal12121499

AMA Style

Ma J, Chang S, Yu F, Lai H, Zhao Y. Research Progress on Sulfur Deactivation and Regeneration over Cu-CHA Zeolite Catalyst. Catalysts. 2022; 12(12):1499. https://doi.org/10.3390/catal12121499

Chicago/Turabian Style

Ma, Jiangli, Shiying Chang, Fei Yu, Huilong Lai, and Yunkun Zhao. 2022. "Research Progress on Sulfur Deactivation and Regeneration over Cu-CHA Zeolite Catalyst" Catalysts 12, no. 12: 1499. https://doi.org/10.3390/catal12121499

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