Next Article in Journal
Preparation and Electrocatalysis Application of Pure Metallic Aerogel: A Review
Next Article in Special Issue
New Insight into the In Situ SO2 Poisoning Mechanism over Cu-SSZ-13 for the Selective Catalytic Reduction of NOx with NH3
Previous Article in Journal
Analysis and Model-Based Description of the Total Process of Periodic Deactivation and Regeneration of a VOx Catalyst for Selective Dehydrogenation of Propane
Previous Article in Special Issue
A Dynamic Model Incorporated with Delay Estimation and Variable Selection for Control of SCR De-NOx Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Promoting Effect of Mn on In Situ Synthesized Cu-SSZ-13 for NH3-SCR

1
State Key Joint Laboratory of Environment Simulation and Pollution Control, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
Center for Excellence in Regional Atmospheric Environment, Institute of Urban Environment, Chinese Academy of Sciences, Xiamen 361021, China
*
Author to whom correspondence should be addressed.
Jinpeng Du and Jingyi Wang contributed equally to this work.
Catalysts 2020, 10(12), 1375; https://doi.org/10.3390/catal10121375
Submission received: 23 October 2020 / Revised: 15 November 2020 / Accepted: 23 November 2020 / Published: 25 November 2020

Abstract

:
The effect of Mn impregnation on the NH3-SCR (selective catalytic reduction of NOx by NH3) activity of in situ synthesized Cu-SSZ-13 was investigated in this work. It was found that Mn addition could efficiently improve the low-temperature activity of Cu-SSZ-13. The optimal amount of Mn was 5 wt.%, and NOx conversion was improved by more than 20% over a temperature range of 120 °C to 150 °C. SEM (scanning electron microscopy), XRD (X-ray diffraction), N2 adsorption-desorption, H2-TPR (temperature programmed reduction of H2), NH3-TPD (temperature programmed desorption of NH3) and in situ DRIFTS (diffuse reflectance infrared Fourier transform spectroscopy) experiments were conducted to investigate the changes in the zeolite structure, active sites, acid sites and reaction mechanism. The impregnated MnOx species caused a decline in the crystallinity of Cu-SSZ-13 but markedly improved the redox ability. Nitrate and nitrite species were observed in the Mn-modified Cu-SSZ-13, and the formation of these species was thought to cause the observed increase in low-temperature NH3-SCR activity. The results show that the addition of Mn is a promising method for promoting the low-temperature catalytic activity of Cu-SSZ-13.

1. Introduction

The emission of nitrogen oxides (NOx) from heavy-duty diesel vehicles makes a large contribution to atmospheric pollution [1,2,3]. The selective catalytic reduction of NOx by NH3 (NH3-SCR) has proven to be one of the most efficient technologies for NOx abatement from heavy-duty diesel vehicles [4,5,6,7,8]. As an NH3-SCR catalyst, Cu-SSZ-13 zeolite has, with its small-pore structure, attracted much attention in recent years [9,10,11]. NH3-solvated Cu ions, which are weakly tethered to framework Al and regarded as mobile active sites, make the NH3-SCR reaction occur in a near-homogeneous fashion in a limited region [12]. The pore size of Cu-SSZ-13 is around 3.8 Å, which screens large molecules such as HCs from entering the zeolite cavities and also restricts the out-diffusion of Al(OH)3 species formed under hydrothermal conditions [13]. Therefore, the special pore structure endows Cu-SSZ-13 with excellent NH3-SCR activity, N2 selectivity, HCs resistance and hydrothermal stability.
Ren et al. invented an in situ synthesis method for Cu-SSZ-13 that uses Cu-tetraethylenepentamine (Cu-TEPA) as a template to directly introduce Cu into the precursor gel [14,15]. Compared with the conventional ion-exchange method, this in situ synthesis method avoids the complicated ion-exchange process and reduces the production of wastewater, which is beneficial for practical application. Furthermore, Xie et al. and Shan et al. optimized the structure of and copper species in in situ synthesized Cu-SSZ-13 by HNO3 and NH4NO3 after treatment, leading to marked increases in the catalytic activity and hydrothermal stability [16,17]. Nevertheless, the low-temperature activity (<180 °C) of in situ synthesized Cu-SSZ-13 is still low and needs to be improved [18,19,20].
Numerous studies showed that, at temperatures below 180 °C, Mn-based catalysts were very active in the NH3-SCR reaction [21,22,23,24,25,26,27]. For instance, over 90% NOx conversion could be obtained over MnOx-CeO2 complex oxide catalysts at 150 °C [21,28]. Tang et al. doped a series of metals on Mn-based catalysts, and they found that the Mn-Fe-Ce catalyst performed the best. A 95% NOx conversion could be obtained, and it also had an excellent SO2 and H2O tolerance [27]. Ye et al. synthesized a series of Mn/SSZ-13 catalysts with various Mn loadings, and over 95% NOx conversion was obtained at 150 °C for the best sample [29]. Meanwhile, Mn-based bimetallic zeolites also showed excellent low-temperature NH3-SCR activity. Kim et al. added Mn into the Fe-ZSM-5 catalyst, and the catalytic performance was improved [30]. The NOx conversion of Cu-SSZ-39 was also raised over the whole temperature range after the addition of Mn [31]. Song et al. synthesized Cu-Mn-SSZ-13 catalysts by a coexchange method, and both the catalytic activity and hydrothermal stability were improved with the appropriate loading amount [32]. It seems that the addition of Mn has positive effects on the NH3-SCR activity of Cu-SSZ-13. To meet the demands of a practical application, the low-temperature catalytic activity of in situ synthesized Cu-SSZ-13, which is one of the most promising catalysts utilized in heavy-duty diesel vehicles, should also be of concern.
In this study, a series of Mn/Cu-SSZ-13 catalysts with various Mn loadings were prepared by impregnating manganese acetates into in situ synthesized Cu-SSZ-13. The results showed that the addition of Mn could improve the catalytic performance of Cu-SSZ-13 catalysts. The effects of Mn impregnation on the zeolite framework and morphology of Cu-SSZ-13 were characterized by XRD and SEM. H2-TPR and NH3-TPD were utilized to assess the redox ability and acidity of Cu-SSZ-13 and Mn/Cu-SSZ-13. In addition, a series of in situ DRIFTS experiments were carried out to investigate the NH3-SCR mechanism over the prepared catalysts.

2. Results and Discussion

2.1. NH3-SCR Performance

The standard NH3-SCR catalytic activity profiles of all the prepared catalysts are presented in Figure 1. As can be seen, all the catalysts showed a broad window of over 90% NOx conversion from around 160 °C to 330 °C. As can be seen, after impregnation with Mn, all catalysts presented a better catalytic activity below 160 °C when compared with Cu-SSZ-13. Among all the catalysts, 5% Mn/Cu-SSZ-13 showed the best catalytic activity in the low temperature range, and nearly 95% NOx conversion was obtained around 140 °C. With a further increase in the Mn content, however, the low-temperature NOx conversion began to decrease when compared with 5% Mn/Cu-SSZ-13. To analyze the selectivity of the prepared catalysts, the N2O concentrations in the effluents from the catalysts are summarized in Figure S1. More N2O was produced with an increase of the Mn content, indicating that the addition of Mn would decrease the N2 selectivity of the catalysts. Nevertheless, the N2O formation was under 20 ppm for Mn/Cu-SSZ-13 with an Mn content below 7%. The accurate control of the Mn content was of vital importance to the catalytic performance, and 5% Mn/Cu-SSZ-13 performed the best among all the prepared catalysts. The following sections will compare the characterization results between 5% Mn/Cu-SSZ-13 and Cu-SSZ-13; the promotion effect of Mn on Cu-SSZ-13 was also investigated.

2.2. Characterization of Catalysts

The SEM images illustrate the morphology of Cu-SSZ-13 (Figure 2a) and 5% Mn/Cu-SSZ-13 (Figure 2b). As can be seen, in both cases the catalyst grains had a cubic shape with a crystallite size of around 300–500 nm. No discernable destruction of the grains could be observed after 5% Mn was impregnated into Cu-SSZ-13, indicating that the morphology of Cu-SSZ-13 was not affected by doping with Mn. Furthermore, the N2 adsorption-desorption experimental results showed that the surface area and pore volume were 613 m2/g and 0.212 cm3/g for Cu-SSZ-13, and 596 m2/g and 0.206 cm3/g for 5% Mn/Cu-SSZ-13. Compared with Cu-SSZ-13, there was only a slight decrease in the surface area and pore volume for 5% Mn/Cu-SSZ-13, indicating that the channel structure was not affected by Mn impregnation. In addition, the SEM images of catalysts with other Mn contents are presented in Figure S2. The morphologies of these catalysts showed little change after Mn impregnation, except for 14% Mn/Cu-SSZ-13, which showed some degradation of the structure. In summary, when the Mn content was below 10%, the MnOx species over Cu-SSZ-13 were highly dispersed and did not affect the channel structure of Cu-SSZ-13.
Furthermore, EDX was utilized to detect the composition of atoms in Cu-SSZ-13 and 5% Cu-SSZ-13, and the results are presented in Figure S3. Ten spots were scanned for each catalyst, and the atomic composition was calculated as the average value of the 10 spots. As can be seen, there were fewer O, Al, Si and Cu atoms in 5% Mn/Cu-SSZ-13 than in Cu-SSZ-13, which was due to the addition of Mn. Meanwhile, EDS could only detect the atoms on the surface of the catalysts. 5% Mn was added into the catalysts; however, only 1.8% Mn was detected. These results indicated that more Mn atoms were located in the pores of Cu-SSZ-13, which was in accordance with the results of the SEM images.
The XRD patterns of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13 are presented in Figure 3. Both catalysts showed the typical diffraction peaks of the CHA structure, indicating that the addition of Mn did not change the zeolite structure of CHA. Meanwhile, no peaks resulting from CuOx species could be observed for either Cu-SSZ-13 or 5% Mn/Cu-SSZ-13. Compared with Cu-SSZ-13, no new diffraction peaks emerged in the pattern of 5% Mn/Cu-SSZ-13, which meant that the MnOx species were highly dispersed. This result was in accordance with the SEM images, which showed no evidence of MnOx species accumulation. The XRD patterns of all the prepared catalysts are presented in Figure S4. With the increase in the Mn content, the crystallinity of the catalysts decreased, which meant that Mn impregnation could lead to the deterioration of the framework of Cu-SSZ-13.
H2-TPR experiments were carried out to evaluate the redox ability of the catalysts, and the results are presented in Figure 4. Three peaks can be observed under 500 °C for Cu-SSZ-13. Peak A was assigned to the reduction of Cu(OH)+ species in CHA cages; peak B was attributed to the reduction of CuOx species; and peak C represented the consumption of H2 by Cu2+ species near d6r [33,34]. It is worth mentioning that no peak corresponding to CuOx species was observed in the XRD patterns, indicating that these CuOx species were highly dispersed. In the H2-TPR curve of 5% Mn/Cu-SSZ-13, five peaks can be observed. The reduction of Cu(OH)+ species in CHA cages, CuOx species and Cu2+ species near d6r were also included in Peaks A, B and C [16,33,35,36,37]. However, the peaks for the Cu species shifted to a lower temperature, indicating that the redox ability was improved. Compared to Cu-SSZ-13, there was an increase in the areas of peak A and B, and a decrease in the area of peak C. The decline of peak C can be attributed to a decrement in the amount of Cu2+ species near d6r, because the impregnated Mn ions might occupy some ion-exchange sites. Meanwhile, the increase in peak B was partially due to the formation of new CuO species, which emerged because the ion-exchange sites of Cu ions were occupied by Mn ions. During the addition of Mn, Cu(OH)+ species in CHA cages must have been lost, since these species are less stable than Cu2+ species near d6r. However, the area of peak A increased, which meant that MnOx species also contributed to peak A. Furthermore, peak D and peak E also emerged as new peaks after the Mn addition, and they were also assigned to the reduction of MnOx species. In the previous literature, the reduction peaks of pure MnOx oxides by H2 were frequently observed in the range between 300 °C and 600 °C, and the reduction of these species occurred by a three-step process: MnO2→Mn2O3→Mn3O4→MnO [38,39]. However, with the addition of other elements, the H2-TPR peaks of MnOx species could move to much lower temperatures [40,41]. Therefore, in this study, the reduction of MnO2 to Mn2O3 occurred in the range of 150 °C to 350 °C; peak D represented the reduction of Mn2O3 to Mn3O4, and peak E was assigned to the reduction of Mn3O4 to MnO. Impregnation with Mn efficiently improved the redox ability of Cu-SSZ-13. The reaction process of NH3-SCR over Cu-SSZ-13 was accompanied with an electron transfer between Cu ions and the reactants [9,12,42]. A better redox ability for catalysts might facilitate the electron transfer process, resulting in an improvement in the catalytic activity. As a result, the better redox ability of 5% Mn/Cu-SSZ-13 could be one of the reasons why it exhibited a higher low-temperature activity than Cu-SSZ-13.
XPS was utilized to detect the oxidation state of Cu in Cu-SSZ-13 and Cu & Mn in 5% Mn/Cu-SSZ-13. As presented in Figure S5a, the binding energy values of Cu 2p 1/2 and Cu 2p 3/2 were 953.9 eV and 934.0 eV in both catalysts, respectively. Meanwhile, there were satellite peaks around 965~962 eV and 941~949 eV in 5% Mn/Cu-SSZ-13; however, no satellite peaks were observed in Cu-SSZ-13 [26,27]. Through the deconvolution of the peak with a binding energy of around 940~930 eV, the composition of Cu ions with different valence states was calculated. 75.2% of Cu ions in Cu-SSZ-13 consisted of Cu2+, compared with 86.8% Cu2+ in 5% Mn/Cu-SSZ-13. A higher amount of Cu2+ was beneficial for a low-temperature catalytic activity, so that 5% Mn/Cu-SSZ-13 performed better [26]. Furthermore, the Mn 2p spectra of 5% Mn/Cu-SSZ-13 are presented in Figure S5b, and the binding energy values at 655.3 eV and 643.2 eV were assigned to Mn 2p 1/2 and Mn 2p 3/2, respectively [26,27,39]. To calculate the composition of Mn ions, the peak around 650~640 eV was deconvoluted; the contents of Mn2+, Mn3+ and Mn4+ ions were 17.9%, 62.3% and 19.8%, respectively. As a result, in addition to there being more Cu2+ ions in 5% Mn/Cu-SSZ-13, another reason why Mn-impregnated catalysts performed better in the low-temperature range was the presence of a large amount of Mn3+ and Mn4+ ions [26,27,39].
To compare the acidity of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13, NH3-TPD experiments were carried out, and the results are shown in Figure 5. After the deconvolution of the NH3 desorption profiles, three peaks could be observed for both catalysts. According to the desorption temperature of NH3, Peak A, Peak B and Peak C were assigned to the desorption of NH3 from weak acid sites, medium acid sites and strong acid sites, respectively. To further analyze the amount of each kind of acid site, the NH3 desorption amount per 1 g of catalyst was calculated, and the results are presented in Figure S6. Overall, 5% Mn/Cu-SSZ-13 possessed more acid sites than Cu-SSZ-13, which was attributed to the contribution of the MnOx species. For Cu-SSZ-13, the weakly adsorbed NH3 could be attributed to surface hydroxyl species; the medium acid sites were assigned to Lewis acid sites, which were mainly from Cu2+ species; and the strong acid sites originated from Brønsted acid sites [43,44,45]. Meanwhile, the desorption of NH3 from MnOx species could be observed in the range of 100 °C to 500 °C, so that MnOx species in 5% Mn/Cu-SSZ-13 contributed to weak, medium and strong acid sites simultaneously [46,47,48]. The amount of weakly absorbed NH3 that was desorbed was 0.82 mmol/g for 5% Mn/Cu-SSZ-13, while it was 0.55 mmol/g for Cu-SSZ-13. Meanwhile, the amount of NH3 that was desorbed from strong acid sites was 0.59 mmol/g and 0.44 mmol/g for 5% Mn/Cu-SSZ-13 and Cu-SSZ-13, respectively. However, it is worth mentioning that Cu-SSZ-13 possessed more medium acid sites than 5% Mn/Cu-SSZ-13. This could be due to Cu-SSZ-13 possessing more Cu2+ species than 5% Mn/Cu-SSZ-13, which was in accordance with the H2-TPR results. In conclusion, the increase in the numbers of weak and strong acid sites after the addition of Mn could also be one of the reasons why catalytic activity was promoted.

2.3. Investigation of Reaction Process

To further investigate the NH3-SCR mechanism over Cu-SSZ-13 and 5% Mn/Cu-SSZ-13, a series of in situ DRIFTS experiments were carried out. The reaction between preabsorbed NH3 and gaseous NO + O2 over the catalysts was investigated first, and the results are presented in Figure 6. The peaks around 3400–3200 cm−1 and 1456 cm−1 were assigned to the stretching and bending vibrations of N-H bonds in NH4+ absorbed on Brønsted acid sites [16,49,50,51]. As the reaction progressed, the peak around 1456 cm−1 first increased and then decreased. The formation of nitrate species resulted in an increase in the peak area around 1456 cm−1, and these species were then consumed at 60 min [49]. The peaks at 3183 cm−1 were assigned to the stretching vibration of NH3 absorbed on Cu2+. Meanwhile, the peaks at 1619 cm−1 and 1255 cm−1 were attributed to the asymmetric and symmetric vibrations of NH3 absorbed on Lewis acid sites [49,51]. The peak at 1255 cm−1 kept decreasing after the injection of NO and O2; however, the peak around 1619 cm−1 first decreased and then increased. This phenomenon indicated that the peak around 1619 cm−1 was overlapped by peaks from other species. In other reports, the peaks around 1619 cm−1 were assigned to nitrate species on Cu2+ ions and NH3 simultaneously absorbed on Lewis acid sites [52,53]. Therefore, in this study, the peak around 1619 cm−1 at 60 min was attributed to nitrate species adsorbed on Cu2+. Meanwhile, the peak at 1575 cm−1 was also assigned to nitrate species, and nitrate species dominated after NO and O2 were injected after 60 min [52,53]. The negative peaks at 944 cm−1 and 906 cm−1 were assigned to the T-O-T vibration disturbed by Cu(OH)+ and Cu2+ species, respectively [16,33,54]. These two peaks’ tendency can be observed for the peak at 1255 cm−1, and almost all NH3 species absorbed on Cu2+ species were consumed at 60 min.
The in situ DRIFTS spectra of 5% Mn/Cu-SSZ-13 are presented in Figure 6b. The characteristic peaks of NH4+ species on the Brønsted acid sites of 5% Mn/Cu-SSZ-13 around 3400–3200 cm−1 and 1457 cm−1 followed the same tendency as observed in Cu-SSZ-13. This indicates that nitrate species are also critical intermediates in the NH3-SCR process over 5% Mn/Cu-SSZ-13. Apart from Cu2+ species, Mnn+ species also acted as Lewis acid sites in 5% Mn/Cu-SSZ-13. Unlike the spectra of Cu-SSZ-13, two separate peaks at 1636 cm−1 and 1605 cm−1 assigned to the NH3 absorbed on Cu2+ and Mn ions were observed for 5% Mn/Cu-SSZ-13, respectively [25,26,39,49,51]. The presence of Mn ions influenced the Cu2+ species; therefore, the peak for NH3 absorbed on Cu2+ moved to higher wavenumbers when compared with the spectrum of Cu-SSZ-13. Meanwhile, the intensity of negative peaks related to Cu species (around 951 cm−1 and 906 cm−1) decreased when compared with Cu-SSZ-13, which also indicated that Cu2+ species were influenced by the addition of Mn. Furthermore, the peak at 1243 cm−1 was assigned to the symmetric vibration of NH3 absorbed on Lewis acid sites, as indicated in Figure 6a [51]. As the reaction progressed, there was a decrease in the peak areas at 1243 cm−1, 951 cm−1 and 906 cm−1, indicating that NH3 absorbed on Lewis acid sites was continuously consumed after NO and O2 were introduced. However, the peak area at 1636 cm−1 increased with the passage of time. This indicated that the peak corresponding to nitrate species overlapped the peak for NH3 absorbed on Lewis acid sites at 1636 cm−1, and abundant nitrate species accumulated after NO and O2 were added for 60 min.
To sum up, similar NH3-SCR mechanisms were observed for Cu-SSZ-13 and 5% Mn/Cu-SSZ-13. With the injection of NO and O2, nitrate species formed and acted as an important intermediate, indicating that the L-H mechanism took place for both catalysts. Meanwhile, NH3 absorbed on Lewis acid sites was continuously consumed after NO and O2 were introduced, which meant that the E-R mechanism also played a role. As a result, the L-H and E-R mechanisms coexisted in the NH3-SCR process over Cu-SSZ-13 and 5% Mn/Cu-SSZ-13.
The reaction between nitrate species and NH3 was studied. First, catalysts were exposed to NO and O2 for 1 h until saturated, and NH3 was introduced from 0 min after purging by N2 for 1 h. As can be seen in Figure 7a, the peaks between 1650–1500 cm−1 were assigned to Cu-NO3 species [49,55,56]. After NH3 was introduced, the intensity of these peaks decreased, indicating that nitrate species were consumed by NH3, which further proved that the L-H mechanism existed during NH3-SCR for the catalysts. The peaks around 3400–3200 cm−1 and 1457 cm−1 were attributed to NH4+ species on Brønsted acid sites and began to emerge at 6 min [16,49,50,51]. Meanwhile, the negative peaks at 951 cm−1 and 906 cm−1, which were respectively assigned to Cu(OH)+ and Cu2+ species, emerged at 4 min [16,33,54]. Furthermore, new peaks at 1622 cm−1 and 1257 cm−1 emerged at 4 min, and they were assigned to the asymmetric and symmetric vibrations of NH3 absorbed on Lewis acid sites [49,51]. A new peak emerging at 1622 cm−1 overlapped the peak resulting from nitrate species, but the reaction process could be investigated by observing the peak at 1575 cm−1. After around 16 min, the peak at 1575 cm−1 disappeared, indicating that the absorbed nitrate species were consumed. This process was faster than the reaction between absorbed NH3 and NO + O2, as presented in Figure 7. This indicated that, compared with the E-R mechanism, the L-H mechanism proceeded more rapidly for the Cu-SSZ-13 used in this study.
A similar process was observed for 5% Mn/Cu-SSZ-13, as presented in Figure 7b. The peaks around 1650–1500 cm−1 were assigned to Cu-NO3 species [49,55,56]. Compared with Cu-SSZ-13, more Cu-NO3 species emerged in the spectrum of 5% Mn/Cu-SSZ-13. Meanwhile, the peak at 1574 cm−1 could hardly be seen after 8 min when compared to 16 min in Cu-SSZ-13, which meant that Cu-NO3 species were consumed more rapidly with 5% Mn/Cu-SSZ-13. Furthermore, the peaks around 1300 cm−1 and 1200 cm−1 were assigned to bidentate nitrate species in 5% Mn/Cu-SSZ-13, and these species can hardly be seen in Cu-SSZ-13 [55,57]. New peaks at 1632 cm−1, 1604 cm−1 and 1246 cm−1 assigned to NH3 absorbed on Lewis acid sites emerged starting at 2 min [49,51]. Negative peaks at 946 and 906 representing Cu(OH)+ and Cu2+ species also appeared at 2 min [16,33,54]. Furthermore, peaks around 3400–3200 cm−1 and 1476 cm−1, which represented NH4+ species absorbed on Brønsted acid sites, also emerged at 4 min [16,49,50,51].
In conclusion, after being exposed to NO and O2, 5% Mn/Cu-SSZ-13 possessed more nitrate species than Cu-SSZ-13, and these species were consumed more rapidly in 5% Mn/Cu-SSZ-13. As we discussed above, the L-H mechanism played an important role in the NH3-SCR process of our catalysts, and nitrate species were critical intermediates in the L-H mechanism process. The NH3-SCR process over Cu-SSZ-13 can be regarded as the redox cycle of Cu+/Cu2+. Cu2+ ions were reduced by NO and NH3, forming N2 and H2O, which consisted of the reduction half cycle. In the oxidation half cycle, two Cu+ ions migrated into one cage to activate O2 and NO, forming nitrate species, and this was the rate-determining step of the whole process [9,12,41,54]. However, with the addition of Mn, the interaction between Mn and Cu facilitated the redox ability of catalysts; meanwhile, MnOx species are also excellent catalysts for NO oxidation [39], so that nitrates formed more easily in 5% Mn/Cu-SSZ-13 and the former rate-determining step was avoided. As a result, 5% Mn/Cu-SSZ-13 exhibited a better NH3-SCR performance than Cu-SSZ-13.

2.4. The Stability of 5% Mn/Cu-SSZ-13

To investigate the stability of the catalysts, an NH3-SCR catalytic activity durability test was carried out, and the results are presented in Figure 8. 5% Mn/Cu-SSZ-13 was chosen as the testing sample due to it having performed the best among all catalysts; meanwhile, the temperature was set at 135 °C when a high NOx conversion could be obtained and it was not over 100%. After being tested for 24 h, about 90% of NOx conversion could still be obtained, and no obvious decrease of the catalytic activity was observed. Furthermore, an XRD experiment was also conducted on the fresh 5% Mn/Cu-SSZ-13 catalyst, 5% Mn/Cu-SSZ-13 catalyst tested for NH3-SCR at 135 °C for 24 h and 5% Mn/Cu-SSZ-13 catalyst tested for NH3-SCR in the whole temperature range. As can be seen in Figure S7, no obvious decrease of the crystallinity was observed in the catalysts that were either tested at 135 °C for 24 h or tested across the whole temperature range. Durability is of great importance in order for NH3-SCR catalysts to be utilized in diesel vehicles, and the results indicated that the 5% Mn/Cu-SSZ-13 catalyst possessed excellent stability as well as a bright future in practical applications.

3. Materials and Methods

3.1. Catalyst Preparation

Cu-SSZ-13 catalysts with 4.6 wt% Cu and an Si/Al ratio of 4.8 were prepared by an in situ synthesis method described previously [14,15,58]. To obtain Mn/Cu-SSZ-13 catalysts with a varying Mn content, Cu-SSZ-13 was impregnated with manganese acetate (C4H6MnO4·4H2O) (SINOPHARM, Beijing, China) solutions of different concentrations. First, 5 g of Cu-SSZ-13 catalyst powder was dispersed in 500 mL deionized water and stirred for 1 h. Then, different amounts of manganese acetate with the equivalent Mn weight of 0.15 g, 0.25 g, 0.35 g, 0.5 g and 0.7 g were added into the solution and stirred for 2 h. The obtained solution was evaporated to dryness by a rotary evaporator at 50 °C, and the powder was placed overnight in an oven at 80 °C. Finally, all the Cu-SSZ-13 powders with and without Mn addition were calcined at 600 °C for 6 h with a ramp rate of 10 °C/min. The obtained catalysts were denoted as Cu-SSZ-13, 3% Mn/Cu-SSZ-13, 5% Mn/Cu-SSZ-13, 7% Mn/Cu-SSZ-13, 10% Mn/Cu-SSZ-13 and 14% Mn/Cu-SSZ-13, respectively.

3.2. Catalyst Characterization

SEM (scanning electron microscopy) (HITACHI, Beijing, China) was utilized to observe the morphology of the catalysts. EDX (Energy Dispersive X-Ray Spectroscopy) (HITACHI, Beijing, China) was used to detect the atoms on the surface of the catalysts. XRD (X-ray diffraction) was carried out to analyze the crystal structure of the catalysts, using a Brucker D8 Advance diffractometer (Brucker, Beijing, China) with Cu Kα (λ = 0.15406 nm). The scan range was from 5°–90°, with step size of 0.02°. N2 adsorption-desorption was carried out to compare the surface areas and pore volumes of the catalysts before and after Mn addition. The surface area of the catalysts was calculated by the multipore BET method, and the micropore volume was calculated by the t-plot method.
H2-TPR was carried out to evaluate the redox ability of the catalysts, and the experiment was conducted on a chemisorption analyzer (Micromeritics AutoChem 2920, Beijing, China). First, the catalysts were pretreated under 20% O2/N2 at 500 °C for 1 h. After cooling down to room temperature, H2/Ar was introduced. After the baseline was stable, the temperature was raised to 800 °C with a ramp rate of 10 °C/min. XPS (X-ray photoelectron spectroscopy) (Thermo, Beijing, China) was utilized to detect the valance state of atoms in the catalysts. NH3-TPD (temperature programmed desorption of NH3) (Thermo, Beijing, China) was conducted to characterize the acidity of the catalysts. 100 mg of catalyst was placed into a quartz tube sealed at both ends with quartz wool. Catalysts were pretreated at 500 °C under 20% O2/N2 for 1 h. After cooling to 50 °C, 500 ppm NH3/N2 was introduced until the catalysts were saturated with NH3, after which the catalysts were purged by N2 until all the unabsorbed NH3 was removed. Finally, with a total flow of N2 of 250 mL/min, the catalysts were heated from 50 °C to 800 °C with a ramp rate of 10 °C/min. The composition of the outlet gas was analyzed by a Nicolet IS 10 infrared spectrometer (Thermo, Beijing, China).
In situ DRIFTS (diffuse reflectance infrared Fourier transform spectroscopy) was used to investigate the reaction mechanism over different catalysts, and the experiments were carried out on a Nicolet IS50 spectrometer (Thermo, Beijing, China) with an MCT/A detector. All catalysts were pretreated at 500 °C for 30 min with 20% O2/N2 before the reaction. The whole reaction process was conducted at 150 °C with a total gas flow of 100 mL/min, and the concentrations of NO, NH3 and O2 were 1000 ppm, 1000 ppm and 5%, respectively. The first series of experiments were carried out to investigate the reaction between adsorbed NH3 and gaseous NO + O2. Catalysts were exposed to NH3 until saturated and then purged by N2 to eliminate surplus NH3. Then, NO and O2 were introduced, and the spectra were collected. In the second series of experiments, the reaction between adsorbed NO + O2 and gaseous NH3 was investigated. NO and O2 were introduced to the catalysts until saturated, after which the unabsorbed gas was purged by N2. Finally, NH3 was introduced, and the spectra were recorded.

3.3. Catalytic Tests

NH3-SCR tests were carried out in a fixed bed reactor (WeiYe, Beijing, China), with the catalysts loaded in quartz tubes sealed by quartz wool. All the catalysts were sieved to a 40–60 mesh before the NH3-SCR tests. The total gas flow was 250 mL/min with the gas composition: [NO] = [NH3] = 500 ppm, [O2] = 5%, N2 balance, and GHSV = 100,000 h−1. The outlet gas was detected by a Nicolet IS 10 infrared spectrometer, and the NOx conversion was calculated by the following equation:
NO x   c o n v e r s i o n   =   ( 1 [ NO x ] o u t [ NO x ] i n ) × 100 %   ( NO x   =   NO   +   NO 2 )

4. Conclusions

In this study, it was found that impregnation with an appropriate amount of Mn could efficiently improve the low-temperature catalytic activity of in situ synthesized Cu-SSZ-13. The promotion effect of Mn on Cu-SSZ-13 was thoroughly investigated from the aspects of the structural framework, redox ability, acid sites and reaction mechanism. It was observed that MnOx species tended to be located in the pores of Cu-SSZ-13 in a highly dispersed form. Meanwhile, the addition of Mn markedly facilitated the redox ability of Cu-SSZ-13. The improvement of the redox ability promoted the formation of nitrate species, which were critical intermediates in the NH3-SCR process. As a result, the low-temperature catalytic performance was greatly improved in Mn-impregnated Cu-SSZ-13 catalysts.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/12/1375/s1, Figure S1: Concentration of N2O produced by the prepared catalysts under NH3-SCR conditions. Figure S2: SEM images of prepared catalysts. (a) Cu-SSZ-13, (b) 3% Mn/Cu-SSZ-13, (c) 5% Mn/Cu-SSZ-13, (d) 7% Mn/Cu-SSZ-13, (e) 10% Mn/Cu-SSZ-13, (f) 14% Mn/Cu-SSZ-13. Figure S3: EDX results of catalysts. (a) Cu-SSZ-13, (b) 5% Mn/Cu-SSZ-13. Figure S4: XRD patterns of Cu-SSZ-13 catalysts with and without Mn impregnation. Figure S5: XPS spectra of catalysts. (a) Cu 2p spectra of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13, (b) Mn 2p spectra of 5% Mn/Cu-SSZ-13. Figure S6: Integrated NH3 desorption amounts per 1 g of catalyst, calculated from NH3-TPD results. Figure S7. XRD patterns of 5% Mn/Cu-SSZ-13 catalysts before and after being used.

Author Contributions

H.H., Y.S. and Y.Z. designed the experiments; J.D. and J.W. performed the experiments, analyzed the data and wrote the manuscript; X.S. and H.H. checked and corrected the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

We wish to express our appreciation for funding from the National Natural Science Foundation of China (grant number: 21637005, 21906172). We also appreciate the funding and support received from Sinopec: Technology Development Contract (contract number: 319002).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tan, Y.; Henderick, P.; Yoon, S.; Herner, J.; Montes, T.; Boriboonsomsin, K.; Johnson, K.; Scora, G.; Sandez, D.; Durbin, T.D. On-Board Sensor-Based NOx Emissions from Heavy-Duty Diesel Vehicles. Environ. Sci. Technol. 2019, 53, 5504–5511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Walters, W.W.; Goodwin, S.R.; Michalski, G. Nitrogen stable isotope composition (delta15N) of vehicle-emitted NOx. Environ. Sci. Technol. 2015, 49, 2278–2285. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, L.; Franco, V.; Mock, P.; Kolke, R.; Zhang, S.; Wu, Y.; German, J. Experimental Assessment of NOx Emissions from 73 Euro 6 Diesel Passenger Cars. Environ. Sci. Technol. 2015, 49, 14409–14415. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, P.; Chen, S.; Gao, S.; Zhang, J.; Wang, H.; Wu, Z. Niobium oxide confined by ceria nanotubes as a novel SCR catalyst with excellent resistance to potassium, phosphorus, and lead. Appl. Catal. B Environ. 2018, 231, 299–309. [Google Scholar] [CrossRef]
  5. Cai, S.; Zhang, D.; Zhang, L.; Huang, L.; Li, H.; Gao, R.; Shi, L.; Zhang, J. Comparative study of 3D ordered macroporous Ce0.75Zr0.2M0.05O2−δ(M = Fe, Cu, Mn, Co) for selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 2014, 4, 93–101. [Google Scholar] [CrossRef]
  6. Li, P.; Xin, Y.; Li, Q.; Wang, Z.; Zhang, Z.; Zheng, L. Ce-Ti amorphous oxides for selective catalytic reduction of NO with NH3: Confirmation of Ce-O-Ti active sites. Environ. Sci. Technol. 2012, 46, 9600–9605. [Google Scholar] [CrossRef]
  7. Chen, Y.; Li, C.; Chen, J.; Tang, X. Self-Prevention of Well-Defined-Facet Fe2O3/MoO3 against Deposition of Ammonium Bisulfate in Low-Temperature NH3-SCR. Environ. Sci. Technol. 2018, 52, 11796–11802. [Google Scholar] [CrossRef]
  8. Han, S.; Ye, Q.; Cheng, S.; Kang, T.; Dai, H. Effect of the hydrothermal aging temperature and Cu/Al ratio on the hydrothermal stability of CuSSZ-13 catalysts for NH3-SCR. Catal. Sci. Technol. 2017, 7, 703–717. [Google Scholar] [CrossRef]
  9. Gao, F.; Peden, C.H. Recent Progress in Atomic-Level Understanding of Cu/SSZ-13 Selective Catalytic Reduction Catalysts. Catalysts 2018, 8, 140. [Google Scholar] [CrossRef] [Green Version]
  10. 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]
  11. Kwak, J.H.; Tran, D.; Burton, S.D.; Szanyi, J.; Lee, J.H.; Peden, C.H.F. Effects of hydrothermal aging on NH3-SCR reaction over Cu/zeolites. J. Catal. 2012, 287, 203–209. [Google Scholar] [CrossRef]
  12. Paolucci, C.; Khurana, I.; Parekh, A.A.; Li, S.; Shih, A.J.; Li, H.; Di Iorio, J.R.; Albarracin-Caballero, J.D.; 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] [PubMed] [Green Version]
  13. Gao, F.; Kwak, J.; Szanyi, J.; Peden, C.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]
  14. Ren, L.; Zhu, L.; Yang, C.; Chen, Y.; Sun, Q.; Zhang, H.; Li, C.; Nawaz, F.; Meng, X.; 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]
  15. Ren, L.; Zhang, Y.; Zeng, S.; Zhu, L.; Sun, Q.; Zhang, H.; Yang, C.; Meng, X.; Yang, X.; Xiao, F.-S. Design and Synthesis of a Catalytically Active Cu-SSZ-13 Zeolite from a Copper-Amine Complex Template. Chin. J. Catal. 2012, 33, 92–105. [Google Scholar] [CrossRef]
  16. Shan, Y.; Du, J.; Yu, Y.; Shan, W.; Shi, X.; He, H. Precise control of post-treatment significantly increases hydrothermal stability of in-situ synthesized cu-zeolites for NH3-SCR reaction. Appl. Catal. B Environ. 2020, 266. [Google Scholar] [CrossRef]
  17. Xie, L.; Liu, F.; Shi, X.; Xiao, F.-S.; He, H. Effects of post-treatment method and Na co-cation on the hydrothermal stability of Cu–SSZ-13 catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2015, 179, 206–212. [Google Scholar] [CrossRef]
  18. Wang, D.; Gao, F.; Peden, C.H.F.; Li, J.; Kamasamudram, K.; Epling, W.S. Selective Catalytic Reduction of NOx with NH3 over a Cu-SSZ-13 Catalyst Prepared by a Solid-State Ion-Exchange Method. ChemCatChem 2014, 6, 1579–1583. [Google Scholar] [CrossRef]
  19. Kim, J.; Cho, S.J.; Kim, D.H. 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]
  20. Hammershøi, P.S.; Negri, C.; Berlier, G.; Bordiga, S.; Beato, P.; Janssens, T.V.W. Temperature-programmed reduction with NO as a characterization of active Cu in Cu-CHA catalysts for NH3-SCR. Catal. Sci. Technol. 2019, 9, 2608–2619. [Google Scholar] [CrossRef]
  21. Liu, Z.; Yi, Y.; Zhang, S.; Zhu, T.; Zhu, J.; Wang, J. Selective catalytic reduction of NOx with NH3 over Mn-Ce mixed oxide catalyst at low temperatures. Catal. Today 2013, 216, 76–81. [Google Scholar] [CrossRef]
  22. Meng, D.; Zhan, W.; Guo, Y.; Guo, Y.; Wang, L.; Lu, G. A Highly Effective Catalyst of Sm-MnOx for the NH3-SCR of NOx at Low Temperature: Promotional Role of Sm and Its Catalytic Performance. ACS Catal. 2015, 5, 5973–5983. [Google Scholar] [CrossRef]
  23. Kang, M.; Park, E.D.; Kim, J.M.; Yie, J.E. Cu–Mn mixed oxides for low temperature NO reduction with NH3. Catal. Today 2006, 111, 236–241. [Google Scholar] [CrossRef]
  24. Zuo, J.; Chen, Z.; Wang, F.; Yu, Y.; Wang, L.; Li, X. Low-Temperature Selective Catalytic Reduction of NOx with NH3 over Novel Mn–Zr Mixed Oxide Catalysts. Ind. Eng. Chem. Res. 2014, 53, 2647–2655. [Google Scholar] [CrossRef]
  25. Wu, S.; Zhang, L.; Wang, X.; Zou, W.; Cao, Y.; Sun, J.; Tang, C.; Gao, F.; Deng, Y.; Dong, L. Synthesis, characterization and catalytic performance of FeMnTiOx mixed oxides catalyst prepared by a CTAB-assisted process for mid-low temperature NH3-SCR. Appl. Catal. A Gen. 2015, 505, 235–242. [Google Scholar] [CrossRef]
  26. Yan, Q.; Chen, S.; Zhang, C.; Wang, Q.; Louis, B. Synthesis and catalytic performance of Cu1Mn0.5Ti0.5Ox mixed oxide as low-temperature NH3 -SCR catalyst with enhanced SO2 resistance. Appl. Catal. B Environ. 2018, 238, 236–247. [Google Scholar] [CrossRef]
  27. Tang, X.; Wang, C.; Gao, F.; Ma, Y.; Yi, H.; Zhao, S.; Zhou, Y. Effect of hierarchical element doping on the low-temperature activity of manganese-based catalysts for NH3-SCR. J. Environ. Chem. Eng. 2020, 8, 104399–104408. [Google Scholar] [CrossRef]
  28. Qi, G.; Yang, R.T.; Chang, R. MnOx-CeO2 mixed oxides prepared by co-precipitation for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2004, 51, 93–106. [Google Scholar] [CrossRef]
  29. Ye, Y.; Shen, F.; Wang, H.; Chen, R. SSZ-13-supported manganese oxide catalysts for low temperature selective catalytic reduction of NOx by NH3. J. Chem. Sci. 2017, 129, 765–774. [Google Scholar] [CrossRef]
  30. Kim, Y.J.; Kwon, H.J.; Heo, I.; Nam, I.-S.; Cho, B.K.; Choung, J.W.; Cha, M.-S.; Yeo, G.K. Mn–Fe/ZSM5 as a low-temperature SCR catalyst to remove NOx from diesel engine exhaust. Appl. Catal. B Environ. 2012, 126, 9–21. [Google Scholar] [CrossRef]
  31. Yao, W.; Gang-gang, L.; Shao-qing, Z.; Xin-yan, Z.; Xin, Z.; Zheng-ping, H. 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. [Google Scholar] [CrossRef]
  32. Song, C.; Zhang, L.; Li, Z.; Luac, Y.; Li, K. Co-Exchange of Mn: A Simple Method to Improve Both the Hydrothermal Stability and Activity of Cu–SSZ-13 NH3–SCR Catalysts. Catalysts 2019, 9, 455. [Google Scholar] [CrossRef] [Green Version]
  33. 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]
  34. Shan, Y.; Sun, Y.; Du, J.; Zhang, Y.; He, H.; Yu, Y.; Shan, W.; He, H. Hydrothermal aging alleviates the inhibition effects of NO2 on Cu-SSZ-13 for NH3-SCR. Appl. Catal. B Environ. 2020, 275, 119105. [Google Scholar] [CrossRef]
  35. 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]
  36. Kim, Y.J.; Lee, J.K.; Min, K.M.; Hong, S.B.; Nam, I.-S.; Cho, B.K. Hydrothermal stability of CuSSZ13 for reducing NOx by NH3. J. Catal. 2014, 311, 447–457. [Google Scholar] [CrossRef]
  37. Wang, Y.; Shi, X.; Shan, Y.; Du, J.; Liu, K.; He, H. Hydrothermal Stability Enhancement of Al-Rich Cu-SSZ-13 for NH3 Selective Catalytic Reduction Reaction by Ion Exchange with Cerium and Samarium. Ind. Eng. Chem. Res. 2020, 59, 6416–6423. [Google Scholar] [CrossRef]
  38. Du, J.; Qu, Z.; Dong, C.; Song, L.; Qin, Y.; Huang, N. Low-temperature abatement of toluene over Mn-Ce oxides catalysts synthesized by a modified hydrothermal approach. Appl. Surf. Sci. 2018, 433, 1025–1035. [Google Scholar] [CrossRef]
  39. Sun, P.; Guo, R.-T.; Liu, S.-M.; Wang, S.-X.; Pan, W.-G.; Li, M.-Y. The enhanced performance of MnOx catalyst for NH3-SCR reaction by the modification with Eu. Appl. Catal. A Gen. 2017, 531, 129–138. [Google Scholar] [CrossRef]
  40. Deng, H.; Kang, S.; Ma, J.; Wang, L.; Zhang, C.; He, H. Role of Structural Defects in MnOx Promoted by Ag Doping in the Catalytic Combustion of Volatile Organic Compounds and Ambient Decomposition of O3. Environ. Sci. Technol. 2019, 53, 10871–10879. [Google Scholar] [CrossRef]
  41. Li, X.; Ma, J.; He, H. Tuning the Chemical State of Silver on Ag–Mn Catalysts to Enhance the Ozone Decomposition Performance. Environ. Sci. Technol. 2020, 54, 11566–11575. [Google Scholar] [CrossRef] [PubMed]
  42. 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] [PubMed]
  43. Yu, T.; Xu, M.; Huang, Y.; Wang, J.; Wang, J.; Lv, L.; Qi, G.; Li, W.; Shen, M. Insight of platinum poisoning Cu/SAPO-34 during NH3-SCR and its promotion on catalysts regeneration after hydrothermal treatment. Appl. Catal. B Environ. 2017, 204, 525–536. [Google Scholar] [CrossRef]
  44. Wang, J.; Yu, T.; Wang, X.; Qi, G.; Xue, J.; Shen, M.; Li, W. The influence of silicon on the catalytic properties of Cu/SAPO-34 for NOx reduction by ammonia-SCR. Appl. Catal. B Environ. 2012, 127, 137–147. [Google Scholar] [CrossRef]
  45. Liu, X.; Xiaodong, W.; Weng, D.; Si, Z.; Ran, R. Evolution of copper species on Cu/SAPO-34 SCR catalysts upon hydrothermal aging. Catal. Today 2017, 281, 596–604. [Google Scholar] [CrossRef]
  46. Chen, X.; Wang, P.; Fang, P.; Ren, T.; Liu, Y.; Cen, C.; Wang, H.; Wu, Z. Tuning the property of Mn-Ce composite oxides by titanate nanotubes to improve the activity, selectivity and SO2/H2O tolerance in middle temperature NH3-SCR reaction. Fuel Process. Technol. 2017, 167, 221–228. [Google Scholar] [CrossRef]
  47. Tian, W.; Yang, H.; Fan, X.; Zhang, X. Catalytic reduction of NOx with NH3 over different-shaped MnO2 at low temperature. J. Hazard. Mater. 2011, 188, 105–109. [Google Scholar] [CrossRef] [PubMed]
  48. Yang, G.; Zhao, H.; Luo, X.; Shi, K.; Wang, W.; Chen, Q.; Fan, H.; Wu, T. Promotion effect and mechanism of the addition of Mo on the enhanced low temperature SCR of NOx by NH3 over MnOx/γ-Al2O3 catalysts. Appl. Catal. B Environ. 2019, 245, 743–752. [Google Scholar] [CrossRef] [Green Version]
  49. Xie, L.; Liu, F.; Liu, K.; Shi, X.; He, H. Inhibitory effect of NO2 on the selective catalytic reduction of NOx with NH3 over one-pot-synthesized Cu–SSZ-13 catalyst. Catal. Sci. Technol. 2014, 4, 1104–1110. [Google Scholar] [CrossRef]
  50. Shi, X.; Liu, F.; Xie, L.; Shan, W.; He, H. NH3-SCR Performance of Fresh and Hydrothermally Aged Fe-ZSM-5 in Standard and Fast Selective Catalytic Reduction Reactions. Environ. Sci. Technol. 2013, 47, 3293–3298. [Google Scholar] [CrossRef] [PubMed]
  51. Lezcano-Gonzalez, I.; Deka, U.; Arstad, B.; Deyne, A.V.Y.-D.; Hemelsoet, K.; Waroquier, M.; Van Speybroeck, V.; Weckhuysen, B.M.; Beale, A.M. Determining the storage, availability and reactivity of NH3within Cu-Chabazite-based Ammonia Selective Catalytic Reduction systems. Phys. Chem. Chem. Phys. 2014, 16, 1639–1650. [Google Scholar] [CrossRef] [PubMed]
  52. Kubota, H.; Liu, C.; Toyao, T.; Maeno, Z.; Ogura, M.; Nakazawa, N.; Inagaki, S.; Kubota, Y.; Shimizu, K.-I. Formation and Reactions of NH4NO3 during Transient and Steady-State NH3-SCR of NOx over H-AFX Zeolites: Spectroscopic and Theoretical Studies. ACS Catal. 2020, 10, 2334–2344. [Google Scholar] [CrossRef]
  53. Ruggeri, M.P.; Nova, I.; Tronconi, E.; Pihl, J.A.; Toops, T.J.; Partridge, W.P. In-situ DRIFTS measurements for the mechanistic study of NO oxidation over a commercial Cu-CHA catalyst. Appl. Catal. B Environ. 2015, 181–192. [Google Scholar] [CrossRef] [Green Version]
  54. 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]
  55. Zhang, Y.; Peng, Y.; Li, K.; Liu, S.; Chen, J.; Li, J.; Gao, F.; Peden, C.H.F. Using Transient FTIR Spectroscopy to Probe Active Sites and Reaction Intermediates for Selective Catalytic Reduction of NO on Cu/SSZ-13 Catalysts. ACS Catal. 2019, 9, 6137–6145. [Google Scholar] [CrossRef]
  56. Chen, H.-Y.; Wei, Z.; Kollar, M.; Gao, F.; Wang, Y.; Szanyi, J.; Peden, C.H.F. A comparative study of N2O formation during the selective catalytic reduction of NOx with NH3 on zeolite supported Cu catalysts. J. Catal. 2015, 329, 490–498. [Google Scholar] [CrossRef] [Green Version]
  57. Hadjiivanov, K.I. Identification of Neutral and Charged NxOySurface Species by IR Spectroscopy. Catal. Rev. 2000, 42, 71–144. [Google Scholar] [CrossRef]
  58. Xie, L.; Liu, F.; Ren, L.; Shi, X.; Xiao, F.-S.; 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]
Figure 1. NH3-SCR performance of the prepared catalysts. Test conditions: [NO] = [NH3] = 500 ppm, [O2] = 5%, N2 balance, GHSV = 100,000 h−1.
Figure 1. NH3-SCR performance of the prepared catalysts. Test conditions: [NO] = [NH3] = 500 ppm, [O2] = 5%, N2 balance, GHSV = 100,000 h−1.
Catalysts 10 01375 g001
Figure 2. SEM images of (a) Cu-SSZ-13 and (b) 5% Mn/Cu-SSZ-13.
Figure 2. SEM images of (a) Cu-SSZ-13 and (b) 5% Mn/Cu-SSZ-13.
Catalysts 10 01375 g002
Figure 3. XRD patterns of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13.
Figure 3. XRD patterns of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13.
Catalysts 10 01375 g003
Figure 4. H2-TPR curves of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13.
Figure 4. H2-TPR curves of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13.
Catalysts 10 01375 g004
Figure 5. NH3-TPD profiles of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13.
Figure 5. NH3-TPD profiles of Cu-SSZ-13 and 5% Mn/Cu-SSZ-13.
Catalysts 10 01375 g005
Figure 6. In situ DRIFTS spectra of (a) Cu-SSZ-13 and (b) 5% Mn/Cu-SSZ-13. Catalysts were exposed to 1000 ppm NH3/N2 until saturated and then purged by N2 to remove weakly absorbed NH3. 1000 ppm NO/N2 and 5% O2 were injected starting at 0 min.
Figure 6. In situ DRIFTS spectra of (a) Cu-SSZ-13 and (b) 5% Mn/Cu-SSZ-13. Catalysts were exposed to 1000 ppm NH3/N2 until saturated and then purged by N2 to remove weakly absorbed NH3. 1000 ppm NO/N2 and 5% O2 were injected starting at 0 min.
Catalysts 10 01375 g006
Figure 7. In situ DRIFTS spectra of (a) Cu-SSZ-13 and (b) 5% Mn/Cu-SSZ-13. Catalysts were exposed to 1000 ppm NO/N2 and 5% O2 until saturated and then purged by N2 to remove weakly absorbed NO and O2. 1000 ppm NH3/N2 was introduced at 0 min.
Figure 7. In situ DRIFTS spectra of (a) Cu-SSZ-13 and (b) 5% Mn/Cu-SSZ-13. Catalysts were exposed to 1000 ppm NO/N2 and 5% O2 until saturated and then purged by N2 to remove weakly absorbed NO and O2. 1000 ppm NH3/N2 was introduced at 0 min.
Catalysts 10 01375 g007
Figure 8. NH3-SCR activity test of 5% Mn/Cu-SSZ-13 at 135 °C for 24 h.
Figure 8. NH3-SCR activity test of 5% Mn/Cu-SSZ-13 at 135 °C for 24 h.
Catalysts 10 01375 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Du, 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. https://doi.org/10.3390/catal10121375

AMA Style

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(12):1375. https://doi.org/10.3390/catal10121375

Chicago/Turabian Style

Du, Jinpeng, Jingyi Wang, Xiaoyan Shi, Yulong Shan, Yan Zhang, and Hong He. 2020. "Promoting Effect of Mn on In Situ Synthesized Cu-SSZ-13 for NH3-SCR" Catalysts 10, no. 12: 1375. https://doi.org/10.3390/catal10121375

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