Next Article in Journal
Dehydration of Biomass-Derived Butanediols over Rare Earth Zirconate Catalysts
Next Article in Special Issue
The Denitration and Dedusting Behavior of Catalytic Filter and Its Industrial Application in Glass Kilns
Previous Article in Journal
Soot Combustion over Niobium-Doped Cryptomelane (K-OMS-2) Nanorods—Redox State of Manganese and the Lattice Strain Control the Catalysts Performance
Previous Article in Special Issue
Promoting Effect of Mn on In Situ Synthesized Cu-SSZ-13 for NH3-SCR
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Insight into the In Situ SO2 Poisoning Mechanism over Cu-SSZ-13 for the Selective Catalytic Reduction of NOx with NH3

1
School of Environment and Natural Resources, Renmin University of China, Beijing 100872, China
2
State Key Joint Laboratory of Environment Simulation and Pollution Control, School of Environment, Tsinghua University, Beijing 100084, China
3
Department of Chemistry, Renmin University of China, Beijing 100872, China
*
Author to whom correspondence should be addressed.
Y.Q., C.F., C.S. have equal contributions.
Submission received: 31 October 2020 / Revised: 15 November 2020 / Accepted: 25 November 2020 / Published: 29 November 2020

Abstract

:
To reveal the nature of SO2 poisoning over Cu-SSZ-13 catalyst under actual exhaust conditions, the catalyst was pretreated at 200 and 500 °C in a flow containing NH3, NO, O2, SO2, and H2O. Brunner−Emmet−Teller (BET), X-ray diffraction(XRD), thermo gravimetric analyzer (TGA), ultraviolet Raman spectroscopy (UV Raman), temperature-programmed reduction with H2 (H2-TPR), temperature-programmed desorption of NO+O2 (NO+O2-TPD), NH3-TPD, in situ diffuse reflectance infrared Fourier transform spectroscopy (in situ DRIFTS), and an activity test were utilized to monitor the changes of Cu-SSZ-13 before and after in situ SO2 poisoning. According to the characterization results, the types and generated amount of sulfated species were directly related to poisoning temperature. Three sulfate species, including (NH4)2SO4, CuSO4, and Al2(SO4)3, were found to form on CZ-S-200, while only the latter two sulfate species were observed over CZ-S-500. Furthermore, SO2 poisoning had a negative effect on low-temperature selective catalytic reduction (SCR) activity, which was mainly due to the sulfation of active sites, including Z2Cu, ZCuOH, and Si-O(H)-Al. In contrast, SO2 poisoning had a positive effect on high-temperature SCR activity, owing to the inhibition of the NH3 oxidation reaction. The above findings may be a useful guideline to design excellent SO2-resistant Cu-based zeolite catalysts.

1. Introduction

For a decade or so, Cu-SSZ-13 zeolite with a chabazite (CHA) topological structure has been extensively used/investigated for the ammonia selective catalytic reduction of NOx (NH3-SCR) from diesel engine emissions [1,2,3]. In general, Cu-SSZ-13 is prepared by solution ion-exchange method using NH4-SSZ-13 and Cu salts [1,4] and one-pot methods [5,6,7]. According to recent research findings, active Cu species in Cu-SSZ-13 are present as two types [8,9]: (1) Cu2+ ions balanced by two adjacent negative framework charges (Z2Cu) of the 6-membered ring; (2) [Cu(OH)]+ species balanced by one negative framework charge (ZCuOH) of the 8-membered ring. Furthermore, these active Cu species could transform into other active-type species under different conditions, such as [(NH3)2CuII-O2-CuII(NH3)2] species [10].
Despite the above achievements, there are still some questions existing about Cu-SSZ-13 under actual exhaust conditions. Among them, SO2 poisoning deactivation has been recognized as a substantial barrier for its wide practical application. To date, the impact of SO2 poisoning on the selective catalytic reduction (SCR) performance of Cu-SSZ-13 and the corresponding deactivation mechanism have been investigated and can be summarized as follows: (1) low-temperature SCR activity was severely influenced, while high-temperature SCR activity seemed almost unaffected [11]. Nevertheless, low-temperature SCR activity is rather important because of the increased challenge in reducing NOx emission during a cold start period or low-load condition [12]. (2) Three sulfate species, namely sulfuric acid (H2SO4), copper sulfate (CuSO4), and aluminum sulfate (Al2(SO4)3), formed on Cu-SSZ-13 when it was exposed to SO2+O2 atmosphere [13]. Note that two active Cu species, i.e., Z2Cu and ZCuOH, poisoned by SO2 were very different. ZCuOH species more easily interacted with SO2 to form CuSO4 [14]. Furthermore, ammonium sulfate ((NH4)2SO4) would generate in case of the coexistence of SO2 and NH3 [15]. This poisoning behavior exhibited a reversible nature and had a threshold of about 350 °C due to decomposition of (NH4)2SO4 [14]. (3) The presence of SO3 was found to cause more serious deactivation over Cu-SSZ-13 [16]. Moreover, the active sites interacted with SO2 and SO3 via different reaction mechanisms [17]. (4) The SCR activity of poisoned Cu-SSZ-13 could be partially recovered using a high-temperature treatment [18].
Overall, previous studies have systematically investigated the effect of SO2/SO3 poisoning on Cu-SSZ-13. However, little attention has been paid to the impact of SO2 poisoning on Cu-SSZ-13 under actual exhaust conditions. Herein, a deactivation study of Cu-SSZ-13 was conducted using SO2 exposure under SCR conditions at 200 and 500 °C. The effect of in situ SO2 poisoning on SCR performance was investigated. Furthermore, various techniques were employed to reveal the underlying in situ SO2 poisoning mechanism.

2. Results

2.1. Catalyst Characterization

2.1.1. BET Results

The N2 adsorption–desorption isotherms of Cu-SSZ-13 before and after in situ SO2 poisoning are shown in Figure 1. All the isotherms can be ascribed to type I, the characteristic physical adsorption of microporous materials, according to Brunauer-Deming-Deming-Teller (BDDT) classification. Furthermore, the N2 physisorption results of the investigated samples are summarized in Table 1. It can clearly be seen that in situ SO2 poisoning had a significant impact on the surface area and pore volume of CZ-F, possibly due to the blocking effect of sulfate species [19]. Furthermore, the decrease degree of SBET and Vtotal for CZ-S-200 was more serious than that for CZ-S-500, possibly due to the fact that more sulfate species were generated during the in situ SO2 poisoning process.

2.1.2. XRD Results

The XRD patterns of Cu-SSZ-13 before and after in situ SO2 poisoning were characterized and the results are shown in Figure 2. CZ-F showed typical chabazite (CHA) structure (PDF #52-0784). Furthermore, no diffraction peaks of CuO phases (PDF #48-1548) were observed, suggesting that Cu species in this sample were (1) in the form of isolated Cu2+ ions; (2) highly defective and non-crystalline; (3) crystalline but with particle sizes lower than the XRD detection limit. After in situ SO2 poisoning at 200 and 500 °C, no diffraction peaks of sulfate species were observed over CZ-S-200 and CZ-S-500. Meanwhile, the relative crystallinities of these two samples were still high, indicating that SO2 poisoning did not damage the CHA structure.

2.1.3. TGA Results

Figure 3 shows the TGA profiles of Cu-SSZ-13 before and after in situ SO2 poisoning. Sample weight loss was found in one temperature regime over CZ-F, while it was found in three and four temperature regimes over CZ-S-200 and CZ-S-500, respectively. The dramatic weight loss below 200 °C observed over all the samples can be assigned to the evaporation of adsorbed H2O [20]. Weight loss at 300–500 °C was only found over CZ-S-200, which can be attributed to the decomposition of (NH4)2SO4 species [21]. Weight loss at 500–800 °C was observed over CZ-S-200 and CZ-S-500, which can be ascribed to the decomposition of CuSO4 species [14]. Note that the weight loss in this temperature region for CZ-S-500 was obviously higher than that for CZ-S-200, indicating that more copper sulfate species were formed as SO2 poisoning temperature increased, well consistent with the previous study [22]. Furthermore, a slight weight loss above 800 °C was observed over CZ-S-200 and CZ-S-500, which can be due to the decomposition of Al2(SO4)3 species [13]. Furthermore, the weight loss above 300 °C decreased in the order of CZ-S-200 (2.32 wt%) > CZ-S-500 (1.88 wt%), thereby proving that more sulfate species were generated during the in situ SO2 poisoning process.

2.1.4. UV Raman Results

UV Raman spectra were collected for Cu-SSZ-13 before and after in situ SO2 poisoning. As shown in Figure 4, the spectrum of CZ-F mainly contained five bands at 330, 475, 800, 1060, and 1200 cm−1. The band at 330 cm−1 can be assigned to the T-O-T vibration mode of the 6-membered rings of CHA structure, while the band at 475 cm−1 can be attributed to the vs(T-O-T) mode of a combination of 4-, 6-, and 8-member rings of CHA structure [23,24]. Meanwhile, the bands at 800 and 1060/1200 cm−1 were typical symmetric and nonsymmetric Si-O stretching vibration modes of zeolite SSZ-13 [25]. Note that no bands related to copper species were observed over CZ-F. Specially, the absence of a band at 350 cm−1 indicates that this sample does not contain Cu-O-Cu species [26]. After in situ SO2 poisoning, except for the above bands, a new band at 1142 cm−1 due to asymmetric S-O stretching vibrations appeared over CZ-S-200 and CZ-S-500 [27], further confirming that surface sulfate species exist in these two samples.

2.1.5. H2-TPR Results

Temperature-programmed reduction with H2 (H2-TPR) measurements were carried out to characterize the redox property of Cu-SSZ-13 before and after in situ SO2 poisoning. According to previous studies, for Cu-SSZ-13, the reduction of isolated Cu2+ to Cu+ occurs below 600 °C, while the reduction of Cu+ to Cu0 occurs at much higher temperatures [4]. It is important to note that the reduction of CuO to Cu0 occurs at about 300 °C [7]. As shown in Figure 5, the main existence of Cu species in CZ-F was in the form of Cu2+ ions. After in situ SO2 poisoning, two new H2 consumption peaks were observed over CZ-S-200 and CZ-S-500. The peak around 200–300 °C can be assigned to the reduction of Cu2+ in CuSO4, while the peak around 400–600 °C can be attributed to the reduction of SO42- in CuSO4 and (NH4)2SO4 [22]. Note that the peak intensity below 200 °C decreased distinctly after in situ SO2 poisoning, due to the fact that Cu2+ ions, including ZCuOH and Z2Cu species, can interact with SO2 to form CuSO4 species. Furthermore, the reduction of Cu2+ in CuSO4 to Cu0 only required one single step [28], which led to significantly higher H2 consumption over CZ-S-200 and CZ-S-500 compared to CZ-F.

2.1.6. NO+O2-TPD Results

Figure 6 shows the temperature-programmed desorption of NO+O2 (NO+O2-TPD) profiles of Cu-SSZ-13 before and after in situ SO2 poisoning. The desorbed NOx species (NO and NO2) over CZ-F, CZ-S-200, and CZ-S-500 were mainly from weakly adsorbed NOx, surface nitrite, and nitrate species. Furthermore, the NOx storage capacities of the catalysts were calculated from Figure 6 and the results are shown in Table 1. Compared to CZ-F, the NOx storage capacities of CZ-S-200 and CZ-S-500 decreased significantly. Based on the above characterization results, it is reasonable to presume that the formation of sulfate species suppressed the adsorption of NOx species on the catalyst surface. Note that except for weakly adsorbed NO, the formation of other adsorbed NOx species starts with NO oxidation to produce gaseous NO2, which then adsorbs on the catalyst surface in the form of weakly adsorbed NO2, nitrite, and nitrate species [29]. Therefore, for CZ-S-200, the degradation of NOx storage capacities could be owing to (1) the coverage of adsorption sites by the generated (NH4)2SO4 species; and (2) the inhibition of oxidation of NO to NO2 by Cu species interacting with SO2 to form CuSO4 species. Nevertheless, the degradation of NOx storage capacities for CZ-S-500 could be caused by point (2). Furthermore, from Table 1, it could also be concluded that the influence of point (1) on NOx storage capacity was obviously stronger than point (2).

2.1.7. NH3-TPD Results

NH3-TPD experiments were performed to investigate the Brønsted and Lewis acid sites of Cu-SSZ-13 before and after in situ SO2 poisoning. As shown in Figure 7, three main NH3 desorption states were observed over CZ-F, CZ-S-200, and CZ-S-500: a low-temperature feature below 250 °C assigned to physisorbed NH3, a medium-temperature feature at about 250–425 °C attributed to NH3 adsorbed on Lewis acid sites (i.e., Cu2+ ions), and a high-temperature feature above 425 °C due to NH3 adsorbed on Brønsted acid sites (i.e., Si-O(H)-Al sites) [30,31]. In addition, NH3 storage capacities of the catalysts were calculated from Figure 8 and the results are shown in Table 1. Obviously, NH3 storage capacity was improved after in situ SO2 poisoning, which might be due to (1) newly generated acidic sites in the form of S-OH; and (2) decomposition of formed (NH4)2SO4 species (only for CZ-S-200). Note that the Brønsted acid sites decreased significantly after in situ SO2 poisoning, suggesting the bond breakage of Si-O(H)-Al sites and the corresponding Al atoms transforming into Al2(SO4)3 [32]. However, the amount of reduction in Brønsted acid sites was almost unchanged as SO2 poisoning temperature increased, indicating that the amount of Al2(SO4)3 species generated at different SO2 poisoning temperature remain constant. Note that as a result of the coaction of the abovementioned reasons, CZ-S-200 had much stronger NH3 storage capacity in comparison with CZ-S-500.

2.1.8. In Situ DRIFTS Results

To further investigate the changes of Cu species in SO2 poisoning samples, in situ DRIFTS measurements were carried out using NH3 as a probe molecule. It is well known that the symmetric and asymmetric T-O-T vibrations of zeolite SSZ-13 disturbed by Cu species are in the band range of 850–1000 cm−1 [33]. As shown in Figure 8, two negative bands were obviously observed over CZ-F. Based on the previous research, the band at about 898 cm−1 can be assigned to Z2Cu ions, i.e., Cu2+, located in the 6-member rings [34], while the band at about 950 cm−1 can be ascribed to ZCuOH ions, i.e., [Cu(OH)]+, located in the 8-member rings of CHA cages [8]. After SO2 poisoning at 200 °C, the band intensity of both ZCu and ZCuOH decreased dramatically over CZ-S-200. This may be due to (1) the two Cu species being able to interact with SO2 and ZCuOH being more reactive with SO2 [14]; and (2) the two Cu species being partially covered by the generated (NH4)2SO4. As a result, the above possible reasons led to reducing the disturbing influence of the Cu species on T-O-T vibrations. Note that the band assigned to ZCu species almost disappeared, indicating that the coverage of (NH4)2SO4 had a far greater impact on ZCu sites. By contrast, Cu species interacting with SO2 were mainly a reason that accounted for the decrease in band intensity over CZ-S-500, because (NH4)2SO4 had decomposed completely at 200 °C (Figure 3). Note that the degree of decrease in ZCuOH species was obviously higher than that for Z2Cu, further confirming that ZCuOH was more reactive with SO2.

2.2. NH3-SCR and NH3 Oxdidation Performances

The standard NH3-SCR light-off curves of NOx for Cu-SSZ-13 before and after in situ SO2 poisoning are depicted in Figure 9a, in which their corresponding N2 selectivity is also shown. CZ-F exhibited over 80% NOx conversion at about 200–550 °C. After in situ SO2 poisoning, CZ-S-200 and CZ-S-200 had obviously lower NOx conversions below 300 °C, while they showed higher NOx conversions above 350 °C. Note that the impact of in situ SO2 poisoning on low-temperature SCR activity was lower than the simple sulfation at same temperature [17]. This may be due to the competitive adsorption between SO2 and NH3/NO on the active sites, which could slow the formation of sulfate species on the catalyst surface. As for the improved high-temperature SCR activity, we will discuss it in detail below. In addition, Cu-SSZ-13 before and after in situ SO2 poisoning showed high N2 selectivity, over 98%, in the whole temperature range, illustrating that SO2 poisoning had little effect on its N2 selectivity.
To further investigate the impact of in situ SO2 poisoning on the active Cu species, NH3 oxidation tests were performed. As shown in Figure 9b, for CZ-F, NH3 conversions started at 200 °C and increased as reaction temperature rose. In contrast, NH3 conversions significantly declined over CZ-S-200 and CZ-S-200. According to the previous studies, only ZCuOH species, among all active species for NH3-SCR reaction, were mainly active sites for NH3 oxidation reaction [35]. Therefore, the above results indicate that partial ZCuOH sites on CZ-S-200 and CZ-S-200 were poisoned, well consistent with the in situ DRIFTS results.

3. Discussion

Based on the above findings, a straightforward in situ SO2 poisoning mechanism could be proposed and is illustrated in Figure 10. For in situ SO2 poisoning at 200 °C, three reaction routes coexisted: (1) SO2 interacted with NH3 to form (NH4)2SO4, which could cover the active sites and cause pore blocking (Table 1). Furthermore, this species affected more Z2Cu than ZCuOH (Figure 8). (2) SO2 interacted with both Z2Cu and ZCuOH to form CuSO4 (Figure 3 and Figure 5), only the interaction with Z2Cu species was slight (Figure 8). (3) SO2 broke Si-O(H)-Al bonds and interacted with Al atoms to form Al2(SO4)3 (Figure 3), leading to the decrease in Brønsted acid sites (Figure 7). For in situ SO2 poisoning at 500 °C, only reaction routes (2) and (3) coexisted. Note that as SO2 poisoning temperature increased, more CuSO4 species were formed (Figure 3), while the amount of Al2(SO4)3 species generated was almost unchanged (Figure 7). In particular, the crystal structure of Cu-SSZ-13 was not damaged during in situ SO2 poisoning process, whether at 200 or 500 °C.
Combined with the reactive results shown in Figure 1, the effect of in situ SO2 poisoning on the SCR performance of Cu-SSZ-13 could fall into two categories. At low-temperature regimes, there was a certain loss in SCR performance in comparison with that of fresh sample, and this loss increased with the increase in SO2 poisoning temperature. The main reasons may be as follows: (1) the formation of CuSO4 species resulted in the reduction in active Cu sites (i.e., Z2Cu and ZCuOH), which were more susceptible to be sulfur-poisoned with the increasing temperatures; (2) for the NH3-SCR reaction, Brønsted acid sites play an important and beneficial role at both low- and high-temperature regimes [31]. Therefore, the decrease in Brønsted acid sites after SO2 poisoning inevitably led to the deterioration of SCR performance. Note that the decrease degrees of SBET and Vtotal for CZ-S-200 were more serious than for CZ-S-500 (Table 1), which appeared contradictory to their low-temperature SCR performance, implying that the decrease in surface area and pore volume was not a dominant deactivation factor for Cu-SSZ-13 after SO2 poisoning. At high-temperature regimes, the NH3 oxidation reaction would compete with the NH3-SCR reaction for the consumption of -NH2, an intermediate active NH3 species [36]. That is the major reason why high-temperature activities decrease over SCR catalysts. From the in situ DRIFTS (Figure 8) and NH3 oxidation test (Figure 9b) results, for Cu-SSZ-13, ZCuOH species were active sites for the NH3 oxidation reaction, which was also very easily reactive with SO2. Therefore, the sulfation of ZCuOH species could inhibit the occurrence of the NH3 oxidation reaction to some extent. From Figure 9a, it can be seen that although partial active sites, including Cu species and Brønsted acid sites, were lost during the SO2 poisoning process, the residual active sites can suffice to sustain the NH3-SCR reaction proceeding. Even improved SCR performances were observed over CZ-S-200 and CZ-S-500 at high temperatures. Among the three catalysts, CZ-S-500 exhibited the best SCR activity at high temperatures as most ZCuOH species in it were sulfated during the in situ SO2 poisoning process at 500 °C.

4. Materials and Methods

4.1. Catalyst Preparation

The Cu-SSZ-13 sample (labeled as CZ-F) used in this study was supplied by China Catalyst Holding Co., LTD (Beijing, China). The Si/Al ratio was 13.5 and the Cu content was 2.67 wt%. In situ poisoning samples were prepared by being treated in a gas flow of 500 ppm NO, 500 ppm NH3, 5% O2, 50 ppm SO2, 5% H2O, and N2 balance for 24 h at 200 and 500 °C. They were labeled as CZ-S-200 and CZ-S-500, respectively.

4.2. Catalytic Performance

The NH3-SCR and NH3 oxidation performance measurements were carried out in a fixed quartz microreactor setup (Beijing, China). Prior to testing, 0.04 g catalysts was sieved to 40–60 mesh. The standard SCR feed gas was composed of 500 ppm NH3, 500 ppm NO, 5% O2, 5% H2O, and N2 balance. Nevertheless, the NH3 oxidation feed gas was composed of 500 ppm NH3, 5% O2, 5% H2O, and N2 balance. The total flow rate was 200 mL/min, and the gas hourly space velocity (GHSV) was 200,000 h−1. For ensuring the stability of the testing results, each temperature test point was kept for 30 min before data were recorded. The concentrations of NH3, NO, NO2, and N2O were continuously monitored by an MKS instrument (Shanghai, China). NOx and NH3 conversions were calculated based on the inlet and outlet gas concentrations at steady state, as shown in Equations (1) and (2), respectively. N2 selectivity was calculated from Equation (3).
NO x   Conversion = [ NO x ] inlet [ NO x ] outlet [ NO x ] inlet × 100 %
NH 3   Conversion = [ NH 3 ] inlet [ NH 3 ] outlet [ NH 3 ] inlet × 100 %
N 2   Selectivity = ( 1 2 [ N 2 O ] outlet [ N O x ] inlet + [ N H 3 ] inlet [ N O x ] outlet [ N H 3 ] outlet ) × 100 %

4.3. Physical and Chemical Characterization

Powder X-ray diffraction (XRD) patterns were taken in the 2θ range of 5–50° at steps of 10° min−1, using an X-ray diffractometer (SmartLab 3 KW, Tokyo, Japan) with Cu Kα (λ = 0.15405 nm) radiation.
The N2 sorption isotherms were measured at 77 K using a Micromeritics 3020 instrument (Waltham, MA, USA) in static mode. Prior to the measurements, the samples were degassed at 300 °C for 4 h.
TGA patterns were obtained within the temperature range from room temperature (RT) to 1000 °C by a Mettler Toledo TGA/DSC 3 instrument (Zurich, Switzerland).
UV Raman spectra were obtained on a Jobin-Yvon T64000 (Paris, France) spectrometer with a triple-stage spectrograph at a resolution of 2 cm−1. The excitation laser line at 325 nm from a Kimmon He-Cd laser was used as the exciting source. The power of the laser on the sample was approximately 4 mW.
Temperature-programmed reduction with H2 (H2-TPR) experiments were carried out on a chemisorption analyzer (Micromeritics Chemisorb, 2920, Waltham, MA, USA) in order to obtain the reducibility of samples and the valence of active sites. The samples in a quartz reactor were tested in 10% H2/Ar gas flow of 100 mL/min at a heating rate of 10 °C min−1. The temperature was raised from 100 to 1000 °C, and the signal of hydrogen consumption was detected by the thermal conductivity detector (TCD).
Temperature-programmed desorption (TPD) of NH3 or NO+O2 was carried out using an MKS instrument to detect different nitrogen-containing species, including NH3, NO, and NO2. Adsorption was carried out directly at room temperature. After adsorption saturation, the NH3 concentration was purged to below 10 ppm, and then the data of temperature-rising desorption were recorded. The temperature-programmed desorption experiments were conducted at a heating rate of 10 °C min−1 from RT to 700 °C in N2 (100 mL min−1).
In situ DRIFTS experiments were performed on an FT-IR spectrometer (Nicolet 6700, Waltham, MA, USA) equipped with a Smart Collector and an MCT/A detector. Prior to each experiment, the catalyst was pretreated at 500 °C for 30 min in N2 at a flow of 100 mL min−1. The DRIFTS spectra were collected by accumulating 32 scans with a resolution of 4 cm−1.

5. Conclusions

In this study, the in situ SO2 poisoning mechanism on Cu-SSZ-13 at 200 and 500 °C was investigated. Three sulfate species, including (NH4)2SO4, CuSO4, and Al2(SO4)3, were found to form on CZ-S-200, while only the latter two sulfate species were observed over CZ-S-500. The deterioration of low-temperature SCR activity was mainly due to the sulfation of active Cu species (i.e., Z2Cu and ZCuOH) and the decrease in Brønsted acid sites. As such, the sulfation degree of active Cu species increased with the increasing in situ SO2 poisoning temperature, whereas the amount of Brønsted acid sites was almost unchanged. Furthermore, the blocking effect of (NH4)2SO4 was relatively small. In contrast, the improvement of high-temperature SCR activity was owing to the inhibition of the NH3 oxidation reaction, from which active Cu species (i.e., ZCuOH) were very easily sulfated under the in situ SO2 poisoning condition.

Author Contributions

Conceptualization, C.F. and C.S.; methodology, J.C. (Jiazhe Chen), L.G. and X.N.; formal analysis, J.C. (Jianjun Chen), Y.P., J.L. and W.Y.; investigation, H.Z.; writing—original draft preparation, Y.Q.; writing—review and editing, T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (21806184), the National Engineering Laboratory for Flue Gas Pollutants Control Technology and Equipment (NEL-KF-201904), and a special fund of the State Key Joint Laboratory of Environment Simulation and Pollution Control (20K07ESPCT).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kwak, J.H.; Tonkyn, R.G.; Kim, D.H.; Szanyi, J.; Peden, C.H.F. Excellent activity and selectivity of Cu-SSZ-13 in the selective catalytic reduction of NOx with NH3. J. Catal. 2010, 275, 187–190. [Google Scholar] [CrossRef]
  2. Fickel, D.W.; Lobo, R.F.; Brookhaven, N.L.B.N. Copper Coordination in Cu-SSZ-13 and Cu-SSZ-16 Investigated by Variable-Temperature XRD. J. Phys. Chem. C 2010, 114, 1633–1640. [Google Scholar] [CrossRef]
  3. Zhang, T.; Li, J.; Liu, J.; Wang, D.; Zhao, Z.; Cheng, K.; Li, J. High activity and wide temperature window of Fe-Cu-SSZ-13 in the selective catalytic reduction of NO with ammonia. AIChE J. 2015, 61, 3825–3837. [Google Scholar] [CrossRef]
  4. Gao, F.; Walter, E.D.; Karp, E.M.; Luo, J.; Tonkyn, R.G.; Kwak, J.H.; Szanyi, J.; Peden, C.H.F. Structure–activity relationships in NH3-SCR over Cu-SSZ-13 as probed by reaction kinetics and EPR studies. J. Catal. 2013, 300, 20–29. [Google Scholar] [CrossRef]
  5. Ren, L.; Zhu, L.; Yang, C.; Chen, Y.; Sun, Q.; Zhang, H.; Li, C.; Nawaz, F.; Meng, X.; Xiao, F. 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. [Google Scholar] [CrossRef] [PubMed]
  6. Xie, L.; Liu, F.; Ren, L.; Shi, X.; Xiao, F.; He, H. Excellent Performance of One-Pot Synthesized Cu-SSZ-13 Catalyst for the Selective Catalytic Reduction of NOx with NH3. Environ. Sci. Technol. 2013, 48, 566–572. [Google Scholar] [CrossRef]
  7. Zhang, T.; Qiu, F.; Chang, H.; Li, X.; Li, J. Identification of active sites and reaction mechanism on low-temperature SCR activity over Cu-SSZ-13 catalysts prepared by different methods. Catal. Sci. Technol. 2016, 6, 6294–6304. [Google Scholar] [CrossRef]
  8. Andersen, C.W.; Bremholm, M.; Vennestrøm, P.N.R.; Blichfeld, A.B.; Lundegaard, L.F.; Iversen, B.B. Location of Cu2+ in CHA zeolite investigated by X-ray diffraction using the Rietveld/maximum entropy method. IUCrJ 2014, 1, 382–386. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, R.; McEwen, J.; Kollár, M.; Gao, F.; Wang, Y.; Szanyi, J.; Peden, C.H.F. NO Chemisorption on Cu/SSZ-13: A Comparative Study from Infrared Spectroscopy and DFT Calculations. ACS Catal. 2014, 4, 4093–4105. [Google Scholar] [CrossRef]
  10. Gao, F.; Mei, D.; Wang, Y.; Szanyi, J.; Peden, C.H.F. Selective Catalytic Reduction over Cu/SSZ-13: Linking Homo- and Heterogeneous Catalysis. J. Am. Chem. Soc. 2017, 139, 4935–4942. [Google Scholar] [CrossRef]
  11. 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]
  12. Paolucci, C.; Di Iorio, J.R.; Ribeiro, F.H.; Gounder, R.; Schneider, W.F. Catalysis Science of NOx Selective Catalytic Reduction with Ammonia Over Cu-SSZ-13 and Cu-SAPO-34; Elsevier: New York, NY, USA, 2016. [Google Scholar]
  13. 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]
  14. Jangjou, Y.; Do, Q.; Gu, Y.; Lim, L.; Sun, H.; Wang, D.; Kumar, A.; Li, J.; Grabow, L.C.; Epling, W.S. Nature of Cu Active Centers in Cu-SSZ-13 and Their Responses to SO2 Exposure. ACS Catal. 2018, 8, 1325–1337. [Google Scholar] [CrossRef]
  15. Wijayanti, K.; Xie, K.; Kumar, A.; Kamasamudram, 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]
  16. 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]
  17. 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]
  18. Jiang, H.; Guan, B.; Peng, X.; Wei, Y.; Zhan, R.; Lin, H.; Huang, Z. Effect of sulfur poisoning on the performance and active sites of Cu/SSZ-13 catalyst. Chem. Eng. Sci. 2020, 226, 115855. [Google Scholar] [CrossRef]
  19. Wijayanti, K.; Andonova, S.; Kumar, A.; Li, J.; Kamasamudram, K.; Currier, N.W.; Yezerets, A.; Olsson, L. Impact of sulfur oxide on NH3-SCR over Cu-SAPO-34. Appl. Catal. B Environ. 2015, 166, 568–579. [Google Scholar] [CrossRef]
  20. Wang, L.; Gaudet, J.R.; Li, W.; Weng, D. Migration of Cu species in Cu/SAPO-34 during hydrothermal aging. J. Catal. 2013, 306, 68–77. [Google Scholar] [CrossRef]
  21. Zhang, L.; Wang, D.; Liu, Y.; Kamasamudram, K.; Li, J.; Epling, W. 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]
  22. Du, J.; Shi, X.; Shan, Y.; Xu, G.; Sun, Y.; Wang, Y.; Yu, Y.; Shan, W.; He, H. Effects of SO2 on Cu-SSZ-39 catalyst for the selective catalytic reduction of NOx with NH3. Catal. Sci. Technol. 2020, 10, 1256–1263. [Google Scholar] [CrossRef]
  23. Dutta, P.K.; Shieh, D.C.; Puri, M. Correlation of framework Raman bands of zeolites with structure. Zeolites 1988, 8, 306–309. [Google Scholar] [CrossRef]
  24. Dutta, P.K.; Rao, K.M.; Park, J.Y. Correlation of Raman spectra of zeolites with framework architecture. J. Phys. Chem. 1991, 95, 6654–6656. [Google Scholar] [CrossRef]
  25. McMillan, P. Structural studies of silicate glasses and melts-applications and limitations of Raman spectroscopy. Am. Mineral. 1984, 69, 622–644. [Google Scholar]
  26. Guo, Q.; Fan, F.; Ligthart, D.A.J.M.; Li, G.; Feng, Z.; Hensen, E.J.M.; Li, C. Effect of the Nature and Location of Copper Species on the Catalytic Nitric Oxide Selective Catalytic Reduction Performance of the Copper/SSZ-13 Zeolite. ChemCatChem 2014, 6, 634–639. [Google Scholar] [CrossRef]
  27. Sato, S.; Higuchi, S.; Tanaka, S. Identification and Determination of Oxygen-Containing Inorganosulfur Compounds by Laser Raman Spectrometry. Appl. Spectrosc. 1985, 39, 822–827. [Google Scholar] [CrossRef]
  28. Shan, Y.; Shi, X.; Yan, Z.; Liu, J.; Yu, Y.; He, H. Deactivation of Cu-SSZ-13 in the presence of SO2 during hydrothermal aging. Catal. Today 2019, 320, 84–90. [Google Scholar] [CrossRef]
  29. Wang, D.; Zhang, L.; Kamasamudram, K.; Epling, W.S. In Situ-DRIFTS Study of Selective Catalytic Reduction of NOx by NH3 over Cu-Exchanged SAPO-34. ACS Catal. 2013, 3, 871–881. [Google Scholar] [CrossRef]
  30. Zhang, T.; Chang, H.; You, Y.; Shi, C.; Li, J. Excellent Activity and Selectivity of One-Pot Synthesized Cu–SSZ-13 Catalyst in the Selective Catalytic Oxidation of Ammonia to Nitrogen. Environ. Sci. Technol. 2018, 52, 4802–4808. [Google Scholar] [CrossRef]
  31. Gao, F.; Washton, N.M.; Wang, Y.; Kollár, M.; Szanyi, J.; Peden, C.H.F. Effects of Si/Al ratio on Cu/SSZ-13 NH3-SCR catalysts: Implications for the active Cu species and the roles of Brønsted acidity. J. Catal. 2015, 331, 25–38. [Google Scholar] [CrossRef] [Green Version]
  32. Shen, M.; Zhang, Y.; Wang, J.; Wang, C.; Wang, J. Nature of SO3 poisoning on Cu/SAPO-34 SCR catalysts. J. Catal. 2018, 358, 277–286. [Google Scholar] [CrossRef]
  33. Kwak, J.H.; Zhu, H.; Lee, J.H.; Peden, C.H.F.; Szanyi, J. Two different cationic positions in Cu-SSZ-13? Chem. Commun. 2012, 48, 4758–4760. [Google Scholar] [CrossRef] [PubMed]
  34. Verma, A.A.; Bates, S.A.; Anggara, T.; Paolucci, C.; Parekh, A.A.; Kamasamudram, K.; Yezerets, A.; Miller, J.T.; Delgass, W.N.; Schneider, W.F.; et al. NO oxidation: A probe reaction on Cu-SSZ-13. J. Catal. 2014, 312, 179–190. [Google Scholar] [CrossRef]
  35. 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.; et al. 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]
  36. Yang, S.; Guo, Y.; Chang, H.; Ma, L.; Peng, Y.; Qu, Z.; Yan, N.; Wang, C.; Li, J. Novel effect of SO2 on the SCR reaction over CeO2: Mechanism and significance. Appl. Catal. B Environ. 2013, 136, 19–28. [Google Scholar] [CrossRef]
Figure 1. N2 adsorption–desorption isotherms of Cu-SSZ-13 before and after in situ SO2 poisoning.
Figure 1. N2 adsorption–desorption isotherms of Cu-SSZ-13 before and after in situ SO2 poisoning.
Catalysts 10 01391 g001
Figure 2. XRD patterns of Cu-SSZ-13 before and after in situ SO2 poisoning.
Figure 2. XRD patterns of Cu-SSZ-13 before and after in situ SO2 poisoning.
Catalysts 10 01391 g002
Figure 3. TGA profiles of Cu-SSZ-13 before and after in situ SO2 poisoning. All the samples were tested in a flow of air.
Figure 3. TGA profiles of Cu-SSZ-13 before and after in situ SO2 poisoning. All the samples were tested in a flow of air.
Catalysts 10 01391 g003
Figure 4. UV Raman spectra of Cu-SSZ-13 before and after in situ SO2 poisoning: (a) CZ-F, (b) CZ-S-200, and (c) CZ-S-500.
Figure 4. UV Raman spectra of Cu-SSZ-13 before and after in situ SO2 poisoning: (a) CZ-F, (b) CZ-S-200, and (c) CZ-S-500.
Catalysts 10 01391 g004
Figure 5. Temperature-programmed reduction with H2 (H2-TPR) profiles of Cu-SSZ-13 before and after in situ SO2 poisoning.
Figure 5. Temperature-programmed reduction with H2 (H2-TPR) profiles of Cu-SSZ-13 before and after in situ SO2 poisoning.
Catalysts 10 01391 g005
Figure 6. Temperature-programmed desorption of NO+O2 (NO+O2-TPD) profiles of Cu-SSZ-13 before and after in situ SO2 poisoning.
Figure 6. Temperature-programmed desorption of NO+O2 (NO+O2-TPD) profiles of Cu-SSZ-13 before and after in situ SO2 poisoning.
Catalysts 10 01391 g006
Figure 7. NH3-TPD profiles of Cu-SSZ-13 before and after in situ SO2 poisoning.
Figure 7. NH3-TPD profiles of Cu-SSZ-13 before and after in situ SO2 poisoning.
Catalysts 10 01391 g007
Figure 8. In situ DRIFTS spectra of Cu-SSZ-13 before and after in situ SO2 poisoning using NH3 as probe molecule.
Figure 8. In situ DRIFTS spectra of Cu-SSZ-13 before and after in situ SO2 poisoning using NH3 as probe molecule.
Catalysts 10 01391 g008
Figure 9. (a) NOx conversion and N2 selectivity of Cu-SSZ-13 before and after in situ SO2 poisoning. Reaction conditions: [NO]inlet = 500 ppm, [NH3]inlet = 500 ppm, [O2]inlet = 5%, [H2O]inlet = 5%, N2 as balance gas, total gas flow rate—200 mL/min, and gas hourly space velocity (GHSV) = 200,000 h−1. (b) NH3 conversion of Cu-SSZ-13 before and after in situ SO2 poisoning. Reaction conditions: [NH3]inlet = 500 ppm, [O2]inlet = 5%, [H2O]inlet = 5%, N2 as balance gas, total gas flow rate—200 mL/min, and GHSV = 200,000 h−1.2
Figure 9. (a) NOx conversion and N2 selectivity of Cu-SSZ-13 before and after in situ SO2 poisoning. Reaction conditions: [NO]inlet = 500 ppm, [NH3]inlet = 500 ppm, [O2]inlet = 5%, [H2O]inlet = 5%, N2 as balance gas, total gas flow rate—200 mL/min, and gas hourly space velocity (GHSV) = 200,000 h−1. (b) NH3 conversion of Cu-SSZ-13 before and after in situ SO2 poisoning. Reaction conditions: [NH3]inlet = 500 ppm, [O2]inlet = 5%, [H2O]inlet = 5%, N2 as balance gas, total gas flow rate—200 mL/min, and GHSV = 200,000 h−1.2
Catalysts 10 01391 g009
Figure 10. Schematic diagram of the in situ SO2 poisoning mechanism over Cu-SSZ-13 at 200 and 500 °C.
Figure 10. Schematic diagram of the in situ SO2 poisoning mechanism over Cu-SSZ-13 at 200 and 500 °C.
Catalysts 10 01391 g010
Table 1. Textural properties and NH3 and NOx capacities of Cu-SSZ-13 before and after in situ SO2 poisoning.
Table 1. Textural properties and NH3 and NOx capacities of Cu-SSZ-13 before and after in situ SO2 poisoning.
SampleSBET (m2 g−1)Vtotal (cm3 g−1)NH3 Storage Capacity (μmol g−1)NOx Storage Capacity (μmol g−1)
CZ-F5430.31250.836.6
CZ-S-2004390.26325.56.1
CZ-S-5005090.30308.522.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qiu, Y.; Fan, C.; Sun, C.; Zhu, H.; Yi, W.; Chen, J.; Guo, L.; Niu, X.; Chen, J.; Peng, Y.; et al. New Insight into the In Situ SO2 Poisoning Mechanism over Cu-SSZ-13 for the Selective Catalytic Reduction of NOx with NH3. Catalysts 2020, 10, 1391. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10121391

AMA Style

Qiu Y, Fan C, Sun C, Zhu H, Yi W, Chen J, Guo L, Niu X, Chen J, Peng Y, et al. New Insight into the In Situ SO2 Poisoning Mechanism over Cu-SSZ-13 for the Selective Catalytic Reduction of NOx with NH3. Catalysts. 2020; 10(12):1391. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10121391

Chicago/Turabian Style

Qiu, Yu, Chi Fan, Changcheng Sun, Hongchang Zhu, Wentian Yi, Jiazhe Chen, Luyao Guo, Xiaoxue Niu, Jianjun Chen, Yue Peng, and et al. 2020. "New Insight into the In Situ SO2 Poisoning Mechanism over Cu-SSZ-13 for the Selective Catalytic Reduction of NOx with NH3" Catalysts 10, no. 12: 1391. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10121391

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