Next Article in Journal
Synthesis and Photocatalytic Activity of Hierarchical Zn-ZSM-5 Structures
Previous Article in Journal
Mesoporous TiO2 from Metal-Organic Frameworks for Photoluminescence-Based Optical Sensing of Oxygen
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insight into Platinum Poisoning Effect on Cu-SSZ-13 in Selective Catalytic Reduction of NOx with NH3

1
Department of Environmental Science & Engineering, College of Chemical Engineering, Huaqiao University, Xiamen 361021, China
2
School of Chemical Engineering, Tianjin University, Tianjin 300072, China
3
School of Materials and Chemical Engineering, Chuzhou University, Chuzhou 239000, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(7), 796; https://doi.org/10.3390/catal11070796
Submission received: 13 May 2021 / Revised: 22 June 2021 / Accepted: 28 June 2021 / Published: 29 June 2021
(This article belongs to the Section Environmental Catalysis)

Abstract

:
Platinum’s (Pt) poisoning effect on Cu-SSZ-13 and its regeneration were investigated. The Pt enhanced the parallel reactions, such as NH3 oxidation and NO oxidation reactions, which decreased the deNOx activities. In the temperature range below 330 °C, the deactivation of Cu-SSZ-13 by Pt poisoning was primarily caused by the overconsumption of NH3, due to the enhanced NH3-selective oxidation reaction, while the formation of NOx in NH3 oxidation and NO oxidation into NO2 further aggravated the degradation when the temperature was above 460 °C. The non-selective NH3 oxidation and non-selective NOx catalytic reduction reactions resulted in increased N2O formation over Pt-doped samples. The transformation of Pt0 into PtOx after hydrothermal aging recovered the deNOx activities of the Pt-poisoned samples.

1. Introduction

Cu-exchanged SSZ-13 has recently been commercialized as a catalyst for the selective catalytic reduction of NOx with a NH3 (NH3-SCR) section in a diesel engine exhaust after-treatment system [1]. For automotive applications, there is a critical requirement for the application life of the SCR catalyst in order to meet the NOx emission standard [2].
Sulfur poisoning (including SO2 and SO3), hydrothermal aging, and chemical contamination (such as Ca, P, Pt, Pd, etc.) would deactivate the Cu-SSZ-13 [3,4,5,6,7]. Specific to the sulfur poisoning, the deactivation caused by SO2 poisoning is mostly reversible, as NH4HSO4 and CuHSO4 intermediates formed on Cu-SSZ-13 can be decomposed after thermal treatment at high temperatures [8]. However, the deactivation caused by SO3 poisoning due to the formation of copper sulfate and structural damage is irreversible [3,7].
Specific to hydrothermal treatment, the deactivation of Cu-SSZ-13 is typically attributed to the destruction and transformation of cupric sites from SCR-active to -inactive forms, which is irreversible [5,9]. Nevertheless, the effect of chemical poisoning on Cu-SSZ-13 is more complex than that of sulfur poisoning and hydrothermal treatment. In addition to the poisoning effect of the destruction and transformation of isolated Cu2+ ions into inactive forms [6,10,11,12], the additional chemical contaminants may act as active sites that catalyze side reactions in the NH3-SCR reaction system.
In general, the standard SCR reaction is the target pathway for the NOx elimination in the NH3-SCR system, as shown in Reaction 1 (R.1):
4NH3 + 4NO + O2 = 4N2 + 6H2
However, some parallel reactions also occur along with the standard reaction, as shown in R.2–R.7.
4NO + 4NH3 + 3O2 = 4N2O + 6H2O
4NH3 + 3O2 = 2N2 + 6H2O
2NH3 + 2O2 = N2O + 3H2O
4NH3 + 5O2 = 4NO + 6H2O
4NH3 + 7O2 = 4NO2 + 6H2O
2NO + O2 = 2NO2
Without the contaminants, the rate of the standard NH3-SCR reaction is much higher than that of the side reactions when using the Cu-SSZ-13 as catalyst, because the isolated Cu2+ ions primarily catalyze the standard reaction [13,14]. Therefore, the Cu-SSZ-13 exhibited excellent deNOx activity. However, the presence of additional chemical contamination sites might help to catalyze the reactions in R.2–R.7, which increases the parallel reactions rate [6]. As such, the competition between the parallel reactions (R.2–R.7) and the standard reaction (R.1) would reduce the SCR activity of Cu-SSZ-13 (R.1) [6].
Pt is a typical chemical contamination species for Cu-SSZ-13 [6]. In an after-treatment system, the SCR catalyst is located downstream of the diesel oxidation catalyst (DOC) and diesel particulate filter (DPF) sections [15]. The Pt coating in DOC and DPF would be volatilized and deposited on the SCR catalyst due to the high exhaust temperature, which significantly decreases the deNOx activities of the SCR catalysts [16,17,18]. The previous work in the industry investigated the Pt poisoning effect on Cu-SSZ-13 with Pt loading in the weight range of 0–0.016 wt. % [19]. This work found that 0.008 wt. % was the threshold with little influence on the activity of Cu-SSZ-13, above which the activity decreased significantly. Lezcano-Gonzalez et al. [6] investigated the effect of series pollutants (Pt, Zn, Ca and P) on Cu/SSZ-13 catalysts during the NH3-SCR process, and Pt poisoning (with Pt loading of 1–2 wt. %) resulted in the NOx conversion reduction, which was predicted to be due to the Pt species’ oxidation ability that promoted the NH3 oxidation reaction (R.3–R.6). A similar poisoning effect of Pt was also observed on Cu-SAPO-34 [18]. It should be noted that the amount of Pt has a significant impact on the SCR performance of Cu-SSZ-13 [19]. Therefore, the effect of a Pt loading of 0.01–0.1 wt. % on Cu-SSZ-13 should be investigated. In addition, the hydrothermal treatment effect on the Pt poisoning influence on Pt/Cu-SSZ-13 has also not been reported.
In this work, a series of Pt/Cu-SSZ-13 samples with different Pt doping contents (0.01–0.1 wt. %) was prepared to elucidate the poisoning impact of Pt on the NH3-SCR performance of the catalysts. The Pt’s influence on the NH3-SCR reaction network was investigated based on transient reaction experiments. The Pt poisoning effects on the structure and cupric sites of Cu-SSZ-13 were evaluated by XRD and H2-TPR measurements. The effect of hydrothermal treatment on the recovery of a Pt-contaminated sample was analyzed.

2. Results and Discussion

2.1. Influence of Pt Poisoning on the Catalytic Performance of Cu-SSZ-13

Figure 1 shows the SCR performance of the Cu-SSZ-13 and Pt poisoning samples with and without hydrothermal treatment. The Cu-SSZ-13 presented above 75% NOx conversion at 150–550 °C, and the N2O concentration was below 20 ppm, indicating that the NO and NH3 were primarily converted via the standard NH3-SCR reaction (R.1) pathway over Cu-SSZ-13. For Cu-SSZ-13 doping with Pt (PtxCu), the NOx conversion decreased with increasing Pt loading, particularly in the temperature range above 300 °C. Notably, larger concentrations of N2O were formed in PtxCu samples than in Cu-SSZ-13 (Figure 2). This indicates that the parallel reactions (e.g., R.2–R.7) should be accelerated by Pt sites [18,20]. In addition, the N2O formation shows a volcanic curve, with the peaks at around 330 °C in the Pt-doped samples. This demonstrates that the parallel reactions (such as R.2 and R.4) contributing to the N2O formation mainly affect the NOx conversion below 330 °C. In other words, some other parallel reactions resulted in a decrease in deNOx activity above 330 °C. This part will be further analyzed in Section 3.2, accompanied by the transient reaction results. In addition, in our work, the NOx conversion decreased with the increasing Pt loading from 0.01 to 0.1 wt. % (Figure 1). However, in the work by Lezcano-Gonzalez et al. [6], the Pt-poisoned Cu-SSZ-13 showed a similar loss of NOx conversion in the Cu-SSZ-13 with Pt loadings of 1 and 2 wt. %. This indicates that the Pt poisoning effect might not change significantly when the Pt loading reaches a high content (around 1 wt.%).
The NOx conversion of Cu-SSZ-13 decreased after hydrothermal treatment. However, the hydrothermally aged PtxCu samples showed only slightly inferior NOx conversions as compared to the Cu-H sample, which was much higher than its fresh counterparts, as shown in Figure 1 and Figure 2. In addition, the N2O formation over the PtxCu samples decreased after hydrothermal treatment. This demonstrates that the hydrothermal treatment partially recovered the activity of Pt-poisoned Cu-SSZ-13.

2.2. Transient Reactions

To investigate the effect of Pt on the standard NH3-SCR reaction and parallel reactions with Cu-SSZ-13 below and above 330 °C, the transient reactions were conducted, and the results are given in Table 1. The procedure of the transient reactions is shown in Figure 3. The NH3 oxidation results in Figure S1 reveal that the Pt poisoning enhanced the NH3 oxidation capacity of PtxCu samples. However, small amounts of NO and NO2 were formed on the Cu-SSZ-13 and Pt-poisoned samples during the NH3 oxidation period at 330 °C, as shown in Table 1. This indicates that the Pt sites mainly enhanced the NH3 selectivity of the oxidation reaction to N2 (R.3), which led to insufficient NH3 taking part in the standard SCR reaction, and decreased the NOx conversion of PtxCu samples (Figure 1) below 330 °C as a result.
Both the non-selective NH3 oxidation reaction (NSNO, R.4) and the non-selective NOx catalytic reduction reaction (NSCR, R.2) contributed to the N2O formation [21]. The larger content of N2O in the PtxCu sample than Cu-SSZ-13 (Figure 2) must be attributed to the enhancement of the NSNO and NSCR reaction rates. However, the N2O formation cannot be attributed to the enhancement of the non-selective NH3 oxidation reaction (NSNO, R.4), because only a low concentration of N2O was formed in the NH3 oxidation period at 330 °C, as shown in Table 1. After the introduction of NO, the N2O concentration was significantly increased (Table 1). This indicates that the N2O formation was primarily attributed to the non-selective NOx catalytic reduction reaction (NSCR, R.2), which was catalyzed by Pt sites.
More NO was formed during the NH3 oxidation at 460 °C (R.6) than at 330 °C. This indicates that, in addition to the decrease in reductant involved in the standard SCR reaction due to the enhancement of the NH3 oxidation reaction by Pt, the oxidation of NH3 to NO (R.6) also contributed to the decrease in NOx conversion at high temperatures (460 °C).
Besides this, at 460 °C, a large amount of NO2 was also produced during the NH3 oxidation of Pt doping samples, while none was observed in Cu-SSZ-13 (Table 1), indicating that a part of the NH3 was oxidized to NO2 by Pt at 460 °C. Meanwhile, NO oxidation into NO2 (R.7) was also enhanced in the PtxCu samples, as confirmed by the higher concentration of NO2 formed over PtxCu than Cu-SSZ-13 during the NO oxidation period (Table 1). Furthermore, the total concentration of NO2 formed in the NH3 oxidation and NO oxidation reactions was similar to that formed in the NH3-SCR period. This demonstrates that Pt loading accelerated the reactions in R.6 and R.7, resulting in the increased NO2 formation in the NH3-SCR process. Those observations above imply that, beside the overconsumption of NH3 following R.1, the additional NO and NO2 production via NH3 and NO oxidation (R.5-R.7) also led to the decline in NOx conversion in the PtxCu samples at high temperatures (around 460 °C) [22,23].

2.3. Pt Poisoning Effect on the Texture and Structure

The ICP results of the samples are given in Table 2. The fresh and aged samples showed similar Cu loading contents, and the Pt amount increased with the increment in the Pt loading. The hydrothermal treatment did not change the Pt content.
The surface area and pore volume decreased with the increasing Pt loading (Table 2). No significant framework dealumination took place as a result of Pt impregnation, as only a slight decline of the peak at 58 ppm, attributed to the tetrahedral Al species, was observed in the Pt-doped samples compared to Cu-SSZ-13 (Figure S4) [14,24]. In addition, no significant formation of octahedrally coordinated Al species (showing a resonance peak at 0 ppm) could be detected [14,24]. These observations above indicate that the dealumination should not be the primary reason for the decline in BET surface area. Actually, SSZ-13 possesses small windows of about 3.8 Å, so mass transfer limitations might be induced due to the deposition of impurities, even at low quantities. Moreover, taking into account the diameter of Pt0 and PtOx (around 1 nm, shown in the TEM results in Figure S6), it is reasonable to assume that some of the Pt might be located in the zeolite micropores, causing a decrease in the micropore volume and the BET surface area. However, those values were largely maintained after hydrothermal treatment, suggesting that the structure of the hydrothermally aged sample remains intact, consistent with the XRD results in Figure S2.

2.4. Pt Poisoning Effect on the Catalytic Active Sites

2.4.1. H2-TPR

The isolated Cu2+ ions are identified as the active sites that catalyze the standard NH3-SCR reaction [25,26,27]. H2-TPR was carried out on the catalysts to detect the state of cupric species, and the results are shown in Figure 4a. For the fresh samples, two peaks in 220–380 °C and in 600–750 °C were observed, representing the reduction of isolated Cu2+ to Cu+ (R.8) and the reduction of Cu+ to Cu0 (R.9), respectively [28,29].
Cu 2 + + 1 2 H 2 = Cu + + H +
Cu + + 1 2 H 2 = Cu 0 + H +
After deconvolution, the peak in the 220–380 °C range can be divided into two peaks with tops at 250 °C and 325 °C, assigned to the reduction of isolated Cu2+ located in eight- and six-member rings, respectively [30]. After the hydrothermal treatment, the reduction peak below 500 °C was widened, and an additional peak at 190 °C appeared. This can be attributed to the transformation of a part of the isolated Cu2+ ions to CuO during the hydrothermal treatment [31].
The isolated Cu2+ ions content of each sample was normalized using the area of the reduction peaks of isolated Cu2+ ions in Cu-SSZ-13 as a reference, shown in Figure 4b. The Cu-SSZ-13 and PtxCu samples contained similar isolated Cu2+ ions contents. This suggests that the Pt poisoning has no influence on the cupric sites, which is consistent with previous work [6]. Therefore, the parallel reactions should exclusively be catalyzed by the Pt species, which reduced the SCR activities. Furthermore, the SCR active sites, i.e., isolated Cu2+ ions, decreased, and the Pt content remained constant (Figure 4b), while the SCR activity was increased for the PtxCu samples after hydrothermal treatment. This indicates that the hydrothermal conditions might change the state of Pt, which plays a critical role in determining the SCR’s performance.

2.4.2. X-ray Photoelectron Spectroscopy

To detect the changes of state in Pt, the XPS technique was utilized, and the Pt 4f spectra of the fresh and aged samples are given in Figure 5a,b, respectively. The Cu-SSZ-13 sample displayed no Pt signals, and a broad peak at 70–80 eV was observed in the PtxCu samples. The signals for the Pt 4f7/2 peaks with a binding energy value at 71.32 eV confirmed the presence of platinum in the zero oxidation state (Pt0) [32]. The Pt 4f7/2 signals at around 72.96 and 74.01 are attributed to PtO and PtO2 species, respectively, which were formed due to the Pt0 oxidation during the catalyst’s calcination in preparation [33]. In comparison, only reduced Pt was observed in the previous work on Pt-poisoned Cu-SSZ-13 [6]. This might be attributed to the different Pt precursors used in the work by Lezcano-Gonzalez et al. ([Pt(NH3)4](NO3)2) and in our work (Pt(NO3)2), as the different precursors have an influence on the state of the Pt [6,34]. After the hydrothermal treatment, the XPS spectra of the hydrothermally aged samples changed. Only the PtO and PtO2 species remained in the aged samples. This indicates that all the Pt0 was oxidized during the aging process, which is consistent with the pervious works that reported that the Pt0 would be easily oxidized at high temperatures [16,18].
The Pt0 content of each sample was calculated, as shown in Figure 5c (details shown in Figure S3). The Pt0 content increased in the order of Pt3Cu > Pt2Cu > Pt1Cu > Cu-SSZ-13, consistent with the extent of the deactivation of the Pt-loaded samples (Figure 1). Combined with the transient reaction results in Table 1, it is suggested that the Pt0 catalyzed the R.2–7, decreasing the deNOx activities of the PtxCu samples (Figure 1). In addition, the aged Cu-SSZ-13 and PtxCu exhibited similar SCR activities after hydrothermal treatment, although a large content of PtOx still existed in the PtxCu-H samples (Figure 1 and Figure 5b). This demonstrates that the PtOx is ineffective in catalyzing the parallel reactions. In other words, the transformation of Pt0 into PtOx results in the regeneration of PtxCu samples after hydrothermal treatment.

3. Experimental

3.1. Catalyst Preparation

The Cu-SSZ-13 was synthesized via one-pot hydrothermal synthesis using copper-tetraethylenepentamine (Cu-TEPA, Guangfu, Tanjin) as a structure-directing agent. In total, 180 mL of gel was prepared with a molar ratio of 2.5 Na2O:1 Al2O3:10 SiO2:150 H2O:2 Cu-TEPA, which was then transferred into a 250 mL autoclave with a Teflon liner and crystallized at 140 °C for 4 days. The solid product was filtrated, washed with water, and then dried at 120 °C for 6 h. Then, 15 g of as-synthesized sample was ion-exchanged with 300 mL NH4NO3 solution (1 mol/L) at 80 °C and kept for 9 h to remove excess Cu. After filtration, washing, and drying at 100 °C, the sample was calcinated in synthetic air at 600 °C for 6 h to obtain Cu-SSZ-13. The Si/Al ratio of Cu-SSZ-13 is 4, as determined by ICP-OES. The details of the preparation can be found in our previous work [14].
The Pt-doped Cu-SSZ-13 catalysts (Pt/Cu-SSZ-13 = 0.01, 0.05, 0.1 wt. %) were prepared by incipient wetness impregnation. Typically, 0.6 mL Pt(NO3)2 solution (Guangfu) at different concentrations (0.088, 0.44 and 0.88 mol/L, respectively) was dropped onto 1 g of Cu-SSZ-13. After calcination at 550 °C for 4 h, the catalysts with different Pt loadings were obtained. The samples were denoted as Pt1Cu, Pt2Cu and Pt3Cu for short, respectively.
All catalysts were treated at 750 °C in air with 10% H2O for 12 h. The flow rate was kept at 500 mL/min. The hydrothermally aged samples with Pt loading were distinguished with the suffix of (-H), and Cu-H was used as the abbreviation for the aged Cu-SSZ-13.

3.2. Catalytic Performance Test

The NH3-SCR activities of the catalysts were tested in a fixed-bed quartz reactor (10 mm in diameter). Typically, 100 mg of the catalyst (40–60 mesh) was used, which was pre-activated in 5% O2/N2 flow at 550 °C for 1 h. The reactant gases contain 500 ppm NO, 500 ppm NH3, 5 vol.% O2, and N2 as balance. The gas flow was kept at 500 mL/min, and the corresponding GHSV was 150,000 h−1. The concentrations of the products (NO, N2O and NO2) were recorded on an FTIR gases analyzer (Thermo Nicolet iS10, Madison, WI, USA) with a 2 m gas cell when the reaction reached a steady state (less than 1 ppm change in ten minutes). The NOx (NOx = NO + NO2) conversion was calculated according to R.10:
N O x = 1 N O x o u t N O x i n × 100 %

3.3. Transient Reaction

Transient reactions were carried out at 330 and 460 °C. First, a similar reaction condition as in the standard SCR measurement mentioned above was adopted for the NH3 oxidation reaction, in which NO was not introduced into the gas mixture. After the reaction reached a steady state, 500 ppm NO was switched into the reaction system. The concentrations of the gases (NO, NO2, N2O and NH3) were measured with the FTIR gas analyzer (Thermo Nicolet iS10).
The surface area and pore volume were determined by N2 physisorption measurement at −196 °C using a Quantachrome Autosorb-1, with pre-degassing at 250 °C for 6 h. The chemical composition of the sample was determined by ICP-OES (VISTA-MPX, Varian, Palo Alto, CA, USA). The H2 temperature-programmed reduction (H2-TPR) was conducted on an AutoChem II 2920 (Micromeritics, Norcross, GA, USA) instrument. The samples were treated at 550 °C in 10%O2/N2 for 1 h before testing. X-ray photoelectron spectroscopy (XPS) was performed with a PHI-1600 ESCA system spectrometer(Perkin- Elmer Co., USA) using Mg Kα as the X-ray source (1253.6 eV) under a residual pressure of 5 × 10−6 Pa. The error of the binding energy was calibrated using C1s at 284.6 eV as the standard.

4. Conclusions

The Cu-SSZ-13 suffered severe deactivation after Pt impregnation, resulting in a decrease in NOx conversion. The Pt0 species accelerated the NH3 selective oxidation reaction to N2 in the temperature range below 330 °C, reducing the deNOx activity of PtxCu samples due to the insufficient amount of reductant taking part in the standard NH3-SCR reaction. In addition to the over-consumption of NH3 due to NH3 oxidation, the formation of NO and NO2 via NH3 and NO oxidation (R.3–R.7) further increased the loss of deNOx activity in Pt-poisoned samples at high temperatures (around 460 °C). The NSCR (R.2) reaction contributed to the increased N2O formation in Pt-poisoned samples at around 330 °C. After hydrothermal treatment, the activity was mostly recovered due to the transformation of Pt0, which catalyzed the parallel reactions, into PtOx species.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11070796/s1, Figure S1: NH3 conversion as a function of temperature during NH3 oxidation experiment for the Cu-SSZ-13 and PtxCu samples. Reactant feed contains 500 ppm NH3, 5% O2 5% H2O, balanced with N2; flow rate at 500 mL/min, GHSV: 150,000 h−1, Figure S2: NO2 formation as a function of temperature during NH3-SCR reaction. The reactant feed contains 500 ppm NH3, 5% O2, 5% H2O, balanced with N2; flow rate at 500 mL/min, GHSV: 150,000 h−1, Figure S3: XRD patterns of the (a) fresh and (b) aged catalysts, Figure S4: Solid state 27Al MAS NMR spectra of the Cu-SSZ-13 and Pt-poisoned samples, Figure S5: Ratio of different Pt species in the Pt loading samples quantified by the XPS spectra in Figure 5, Figure S6: TEM images of (a) Pt3Cu and (b) Pt3Cu-H samples.

Author Contributions

H.Z.: literature search, figures, study design, data collection, writing; L.H.: literature search, figures, study design; Y.W.: figures, study design; J.Z.: data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Scientific Research Funds of Huaqiao University (605-50Y200270001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Beale, A.M.; Gao, F.; Lezcano-Gonzalez, I.; Peden, C.H.F.; Szanyi, J. Recent advances in automotive catalysis for NOx emission control by small-pore microporous materials. Chem. Soc. Rev. 2015, 44, 7371–7405. [Google Scholar] [CrossRef]
  2. Nova, I.; Tronconi, E. Urea-SCR Technology for deNOx After Treatment of Diesel Exhausts; Springer: New York, NY, USA, 2014. [Google Scholar]
  3. Shen, M.; Zhang, Y.; Wang, J.; Wang, C.; Wang, J. Nature of SO3 poisoning on Cu/SAPO-34 SCR catalysts. Catalysts 2018, 358, 277–286. [Google Scholar] [CrossRef]
  4. Kumar, A.; Smith, M.A.; Kamasamudram, K.; Currier, N.W.; An, H.; Yezerets, A. Identification of two types of Cu sites in Cu/SSZ-13 and their unique responses to hydrothermal aging and sulfur poisoning. Catal. Today 2014, 231, 75–82. [Google Scholar] [CrossRef]
  5. Gao, F.; Szanyi, J. On the hydrothermal stability of Cu/SSZ-13 SCR catalysts. Appl. Catal. A Gen. 2018, 560, 185–194. [Google Scholar] [CrossRef]
  6. Lezcano-Gonzalez, I.; Deka, U.; van der Bij, H.E.; Paalanen, P.; Arstad, B.; Weckhuysen, B.M.; Beale, A.M. Chemical deactivation of Cu-SSZ-13 ammonia selective catalytic reduction (NH3-SCR) systems. Appl. Catal. B Environ. 2014, 339–349, 154–155. [Google Scholar] [CrossRef]
  7. 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]
  8. Wang, C.; Wang, J.; Wang, J.; Yu, T.; Shen, M.; Wang, W.; Li, W. 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]
  9. 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]
  10. Ma, J.; Si, Z.C.; Weng, D.; Wu, X.D.; Ma, Y. Potassium poisoning on Cu-SAPO-34 catalyst for selective catalytic reduction of NOx with ammonia. Chem. Eng. J. 2015, 267, 191–200. [Google Scholar] [CrossRef]
  11. Fan, C.; Chen, Z.; Pang, L.; Ming, S.; Dong, C.; Albert, K.B.; Liu, P.; Wang, J.; Zhu, D.; Chen, H.; et al. Steam and alkali resistant Cu-SSZ-13 catalyst for the selective catalytic reduction of NOx in diesel exhaust. Chem. Eng. J. 2018, 334, 344–354. [Google Scholar] [CrossRef]
  12. Shwan, S.; Jansson, J.; Olsson, L.; Skoglundh, M. Chemical deactivation of Fe-BEA as NH3-SCR catalyst—Effect of phosphorous. Appl. Catal. B Environ. 2014, 147, 111–123. [Google Scholar] [CrossRef]
  13. Wang, J.; Zhao, H.; Haller, G.; Li, Y. Recent advances in the selective catalytic reduction of NOx with NH3 on Cu-Chabazite catalysts. Appl. Catal. B Environ. 2017, 202, 346–354. [Google Scholar] [CrossRef]
  14. Zhao, H.; Zhao, Y.; Liu, M.; Li, X.; Ma, Y.; Yong, X.; Chen, H.; Li, Y. 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]
  15. Gu, Y.; Marino, S.; Cortés-Reyes, M.; Pieta, I.S.; Pihl, J.A.; Epling, W.S. Integration of an Oxidation Catalyst with Pd/Zeolite-Based Passive NOx Adsorbers: Impacts on Degradation Resistance and Desorption Characteristics. Ind. Eng. Chem. Res. 2020, 60, 6455–6464. [Google Scholar] [CrossRef]
  16. Chen, X.; Currier, N.; Yezerets, A.; Kamasamudram, K. Axially Resolved Performance of Cu-Zeolite SCR Catalysts. SAE Int. J. Engines 2013, 6, 856–861. [Google Scholar] [CrossRef]
  17. Girard, J.W.; Montreuil, C.; Kim, J.; Cavataio, G.; Lambert, C. Origin and Effect of Ultra-Low Platinum Contamination on Diesel-SCR Catalysts. SAE Int. J. Fuels Lubr. 2008, 1, 1553–1559. [Google Scholar]
  18. 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]
  19. Cavataio, G.; Jen, H.-W.; Girard, J.W.; Dobson, D.; Warner, J.R.; Lambert, C.K. Laboratory Study of Soot, Propylene, and Diesel Fuel Impact on Zeolite-Based SCR Filter Catalysts. SAE Int. J. Fuels Lubr. 2009, 2, 204–216. [Google Scholar] [CrossRef]
  20. Lezcano-Gonzalez, I.; Wragg, D.S.; Slawinski, W.A.; Hemelsoet, K.; Deyne, A.; Waroquier, M.; van Speybroeck, V.; Beale, A.M. Determination of the Nature of the Cu Coordination Complexes Formed in the Presence of NO and NH3 within SSZ-13. J. Phys. Chem. C 2015, 119, 24393–24403. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, J.; Cui, H.; Dong, X.; Zhao, H.; Wang, Y.; Chen, H.; Yao, M.; Li, Y. N2O formation in the selective catalytic reduction of NOx with NH3ona CeMoOx catalyst. Appl. Catal. A Gen. 2015, 505, 8–15. [Google Scholar] [CrossRef]
  22. Dong, X.; Wang, J.; Zhao, H.; Li, Y. The promotion effect of CeOx on Cu-SAPO-34 catalyst for selective catalytic reduction of NOx with ammonia. Catal. Today 2015, 258, 28–34. [Google Scholar] [CrossRef]
  23. Zhao, H.; Li, H.; Li, X.; Liu, M.; Li, Y. The promotion effect of Fe to Cu-SAPO-34 for selective catalytic reduction of NOx with NH3. Catal. Today 2017, 297, 84–91. [Google Scholar] [CrossRef]
  24. Zhao, H.; Zhao, Y.; Ma, Y.; Yong, X.; Wei, M.; Chen, H.; Zhang, C.; Li, Y. Enhanced hydrothermal stability of a Cu-SSZ-13 catalyst for the selective reduction of NOx by NH3 synthesized with SAPO-34 micro-crystallite as seed. J. Catal. 2019, 377, 218–223. [Google Scholar] [CrossRef]
  25. 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]
  26. 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, 118655. [Google Scholar] [CrossRef]
  27. Shan, Y.; Sun, Y.; Du, J.; Zhang, Y.; Shi, X.; Yu, Y.; Shan, W.; He, H.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]
  28. 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]
  29. 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]
  30. 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]
  31. Li, X.; Zhao, Y.; Zhao, H.; Liu, M.; Ma, Y.; Yong, X.; Chen, H.; Li, Y. The Cu migration of Cu-SAPO-34 catalyst for ammonia selective catalytic reduction of NOx during high temperature hydrothermal aging treatment. Catal. Today 2019, 237, 126–133. [Google Scholar] [CrossRef]
  32. Cordatos, H.; Gorte, R.J. CO, NO, and H2 Adsorption on Ceria-Supported Pd. J. Catal. 1996, 159, 112–118. [Google Scholar] [CrossRef]
  33. Şen, F.; Gökağaç, G.; Şen, S. High performance Pt nanoparticles prepared by new surfactants for C1 to C3 alcohol oxidation reactions. J. Nanopart. Res. 2013, 15, 1979. [Google Scholar] [CrossRef]
  34. Borges, L.R.; Lopez-Castillo, A.; Meira, D.M.; Gallo, J.M.R.; Zanchet, D.; Bueno, J.M.C. Effect of the Pt Precursor and Loading on the Structural Parameters and Catalytic Properties of Pt/Al2O3. ChemCatChem 2019, 11, 3064–3074. [Google Scholar] [CrossRef]
Figure 1. NOx conversion during standard NH3−SCR with the fresh and aged samples. The reaction conditions: 500 ppm NH3, 500 ppm NO, 5% O2, 5% H2O, N2 balance, flow rate at 500 mL/min, 100 mg of catalyst.
Figure 1. NOx conversion during standard NH3−SCR with the fresh and aged samples. The reaction conditions: 500 ppm NH3, 500 ppm NO, 5% O2, 5% H2O, N2 balance, flow rate at 500 mL/min, 100 mg of catalyst.
Catalysts 11 00796 g001
Figure 2. N2O concentration during standard NH3−SCR of fresh and aged catalyst samples. The reaction conditions: 500 ppm NH3, 500 ppm NO, 5% O2, 5% H2O, N2 balance, flow rate at 500 mL/min, 100 mg of catalyst.
Figure 2. N2O concentration during standard NH3−SCR of fresh and aged catalyst samples. The reaction conditions: 500 ppm NH3, 500 ppm NO, 5% O2, 5% H2O, N2 balance, flow rate at 500 mL/min, 100 mg of catalyst.
Catalysts 11 00796 g002
Figure 3. The procedure of the transient reactions.
Figure 3. The procedure of the transient reactions.
Catalysts 11 00796 g003
Figure 4. (a) H2-TPR spectra of fresh and aged Cu-SSZ-13 and PtxCu samples; (b) the relative Cu2+ ion content in the catalyst samples, using Cu-SSZ-13 as a reference.
Figure 4. (a) H2-TPR spectra of fresh and aged Cu-SSZ-13 and PtxCu samples; (b) the relative Cu2+ ion content in the catalyst samples, using Cu-SSZ-13 as a reference.
Catalysts 11 00796 g004
Figure 5. Pt 4f XPS results of the catalyst samples. (a) Fresh samples, (b) aged samples, (c) Pt0 content in PtxCu samples.
Figure 5. Pt 4f XPS results of the catalyst samples. (a) Fresh samples, (b) aged samples, (c) Pt0 content in PtxCu samples.
Catalysts 11 00796 g005
Table 1. Transient reaction results for the catalyst samples at 330 °C and 460 °C.
Table 1. Transient reaction results for the catalyst samples at 330 °C and 460 °C.
Sample T (°C)Reaction
NH3 OxidationSCR ReactionNO Oxidation
NONH3N2ONO2NONH3N2ONO2NON2ONO2
(ppm)(ppm)(ppm)
Cu-SSZ-1333005940004048515
46000100050479113
Pt1Cu33000708501811465112
460480433060844437156
Pt2Cu3300012015203914435223
4609804233560892397181
Pt3Cu330170241023205929424135
4601600535488012118404190
Table 2. The element content and textural properties of the catalysts.
Table 2. The element content and textural properties of the catalysts.
SampleCu Content a
(wt. %)
Pt Content a
wt. %
Surface Area b
(m2/g)
Pore Volume b
(cm3/g)
Cu-SSZ-134.11NDe4980.21
Pt1Cu4.130.0094120.2
Pt2Cu4.080.0513900.19
Pt3Cu4.140.13820.19
Cu-H4.06ND4060.17
Pt1Cu-H4.100.013550.17
Pt2Cu-H4.100.0483420.16
Pt3Cu-H4.090.13350.15
a Determined with ICP. b Determined with N2 physisorption.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, H.; Han, L.; Wang, Y.; Zheng, J. Insight into Platinum Poisoning Effect on Cu-SSZ-13 in Selective Catalytic Reduction of NOx with NH3. Catalysts 2021, 11, 796. https://doi.org/10.3390/catal11070796

AMA Style

Zhao H, Han L, Wang Y, Zheng J. Insight into Platinum Poisoning Effect on Cu-SSZ-13 in Selective Catalytic Reduction of NOx with NH3. Catalysts. 2021; 11(7):796. https://doi.org/10.3390/catal11070796

Chicago/Turabian Style

Zhao, Huawang, Lei Han, Yujie Wang, and Jiandong Zheng. 2021. "Insight into Platinum Poisoning Effect on Cu-SSZ-13 in Selective Catalytic Reduction of NOx with NH3" Catalysts 11, no. 7: 796. https://doi.org/10.3390/catal11070796

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