Next Article in Journal
Zeolites as Acid/Basic Solid Catalysts: Recent Synthetic Developments
Next Article in Special Issue
Optimization of Ammonia Oxidation Using Response Surface Methodology
Previous Article in Journal
The Use of Power Ultrasound for the Production of PEMFC and PEMWE Catalysts and Low-Pt Loading and High-Performing Electrodes
Previous Article in Special Issue
Effect of Preparation Method of Co-Ce Catalysts on CH4 Combustion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Various Pd Species in Pd/BEA for Cold Start Application

1
Key Laboratory for Green Chemical Technology of State Education Ministry, School of Chemical Engineering & Technology, Tianjin University, Tianjin 300072, China
2
State Key Laboratory of Engines, Tianjin University, Tianjin 300072, China
3
Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 300072, China
4
Wuxi Weifu Lida Catalytic Converter Co., Ltd., Wuxi 214028, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(3), 247; https://doi.org/10.3390/catal9030247
Submission received: 16 February 2019 / Revised: 2 March 2019 / Accepted: 4 March 2019 / Published: 7 March 2019
(This article belongs to the Special Issue Emissions Control Catalysis)

Abstract

:
A series of Pd/BEA catalysts with various Pd loadings were synthesized. Two active Pd2+ species, Z-Pd2+-Z and Z-Pd(OH)+, on exchanged sites of zeolites, were identified by in situ FTIR using CO and NH3 respectively. Higher NOx storage capacity of Z-Pd2+-Z was demonstrated compared with that of Z-Pd(OH)+, which was caused by the different resistance to H2O. Besides, lower Pd loading led to a sharp decline of Z-Pd(OH)+, which was attributed to the ‘exchange preference’ for Z-Pd2+-Z in BEA. Based on this research, the atom utilization of Pd can be improved by decreasing Pd loading.

Graphical Abstract

1. Introduction

The exhaust regulation on NOx emissions is getting more stringent [1]. Presently, NH3 selective catalytic reduction (SCR) [2] and NOx storage reduction (NSR) [3] are widely used for NOx removal. However, standard NOx aftertreatment technologies fail to function efficiently at low temperatures, which results in a large proportion of the tailpipe NOx emission [4]. Meanwhile, high efficiency internal combustion engines require new and/or improved technologies which specifically address their low exhaust temperatures. In response to difficulties of low temperature emissions control, numerous efforts are underway to develop catalysts that light-off at temperatures below 150 °C. Passive NOx adsorbers (PNAs) could play a critical role in enabling high efficiency advanced combustion systems.
Recently, Pd/zeolite serving as passive NOx adsorbers (PNAs) was first proposed by Chen et al. [4]. This catalyst is emerging as effective passive NOx adsorbent technology because of its NOx storage/release capabilities, resistance to sulfur poisoning and hydrothermal deactivation [4,5,6,7,8]. Due to these excellent characteristics, Pd/zeolite has attracted great interest recently and has been further optimized [6,7,8,9,10,11,12,13,14,15,16,17,18].
Several recent studies show that isolated Pd ions in Pd/zeolite are the main active sites for NOx trapping [6,7,11,13,16]. It is reported that there are nine skeletal T sites with various chemical environments in BEA [19]. So, isolated Pd ions located in various positions of zeolite framework are likely to be formed. As reported by Gao et al. and Giordanino et al. [20,21,22], two kinds of isolated Cu species (Cu2+ and [Cu(OH)]+) are identified on various framework positions of Cu/SSZ-13, which indicates that species of isolated cations may be influenced by their locations in zeolites. They also mentioned that Cu species were significantly affected by Cu loading. So, various isolated Pd species may co-exist in Pd/BEA, and Pd loading may be capable of influencing the content of them. Actually, two kinds of active isolated Pd ions (bare Pd2+ and Pd(OH)+) have been observed by Zheng et al. [10] and Khivantsev et al. [12]. However, they did not point out the difference in adsorption between these two species.
Therefore, these two kinds of isolated Pd ions were further studied in this research. A series of in situ FTIR experiments in CO and NH3 were carried out, and a semiquantitative method was adopted to distinguish these two species. Based on this method, the NOx storage capacity of each isolated Pd species was compared. Further, by increasing the proportion of isolated Pd species with higher NOx storage capacities, the atom utilization of Pd can be improved.

2. Results

2.1. Ex-Situ FTIR

Ex-situ FTIR spectra of each sample are exhibited in Figure 1. Compared with 0-Pd (Figure 1a), an extra vibration at 926 cm−1 is observed on the FTIR spectrum of 1-Pd. This peak is attributed to the vibration distortion of the skeletal T–O–T bond due to the strong interaction of Pd ions [4]. This is an obvious piece of evidence for the existence of Pd ions on the exchange sites of zeolites. Peaks at 926 cm−1 appear on spectra of 0.2-Pd, 0.5-Pd and 2-Pd, too (Figure 1b). So, the existence of Pd ions on exchange sites in these samples is confirmed.
Compared with 1-Pd (Figure 1a), the peak at 926 cm−1 disappears in the spectrum of Na-1-Pd, which indicates the elimination of isolated Pd ions on exchange sites. This is firm evidence of which isolated Pd ions in Na-1-Pd have been completely exchanged by Na+ ions through the titration process. So, Pd loaded on Na-1-Pd mainly exists in form of Pd oxidations.

2.2. Na+ Titration

Na+ titration was adopted to measure the content of isolated Pd ions in each sample [23], and the proportion of isolated Pd ions in total Pd loading (marked as Isolated Pd/Pd loading in Table 1) was further calculated. Considering that Pd is sensitive to Cl [24,25], the titration process was carried out in a NaNO3 solution. Besides, this measurement is believed to have no negative effect on zeolite structures since the titration process was carried out in mild conditions and no calcination was done. As shown in Table 1, lower Pd loading leads to less isolated Pd2+ content, whereas the proportion of isolated Pd ions in total Pd loading becomes larger. This phenomenon indicates that the increase of Pd loading leads to the formation of more Pd oxidations (PdOx), which is consistent with Jaeha Lee et al.’s study [11].

2.3. CO In Situ FTIR

Pd species were further probed by CO using in situ FTIR, and the result is displayed in Figure 2. Peaks below 2100 cm−1 are attributed to CO signals on metallic Pd (Pd0) formed via CO reduction [10,26], among which Pd0−CO (atop) are found at 2098 cm−1, 2076 cm−1 and Pd20−CO (bridging) are found at 1951 cm−1. Besides, the peak at 2117 cm−1 is attributed to the C-O vibration on Pd+ [10,27]. As reported by Vu et al. [9], Pd+ is formed by the CO reduction of ion-exchanged Pd species [9]. CO signals on isolated Pd2+ species are observed above 2100 cm−1. The peak at 2152 cm−1 with a shoulder peak at 2137 cm−1 is attributed to the C-O vibration on isolated Pd2+ bonded with the hydroxy of zeolites (marked as Z-Pd2+-Z) [10]. The peak at 2179 cm−1 is attributed to the vibration of C-O adsorbed by another kind of isolated Pd2+ [10], which was first determined as Z-Pd(OH)+ by Okumura et al. [28].
In short, two kinds of isolated Pd2+ (Z-Pd2+-Z and Z-Pd(OH)+) were identified in 0.2-Pd, 0.5-Pd, 1-Pd and 2-Pd. Note that all spectra in Figure 2 were obtained when the steady state had been achieved, and corresponding peaks of Z-Pd2+-Z and Z-Pd(OH)+ were still observed. So, these two isolated Pd2+ species cannot be completely reduced by CO at this temperature, which indicated that the reduction reaction is a reversible one. As shown in Figure S1, the existence of Z-Pd2+-Z in 1-Pd-80 is also demonstrated by CO whereas no Z-Pd(OH)+ is observed. Since the complete reduction of Z-Pd(OH)+ cannot be achieved by CO, there is only one isolated Pd2+ species, Z-Pd2+-Z, in 1-Pd-80 (see Figure S1).

2.4. Catalyst Evaluation

Profiles of the NO adsorption stage are shown in Figure 3. NOx storage capacities are calculated by the integration of negative peaks on these profiles, and the dead volume has been subtracted. The result is displayed in Table 2. Note that samples with the same Pd loading as much as 1 wt % (see Figure S2) exhibit entirely different NOx storage capacities (53.3 μmol/gcat and 9.9 μmol/gcat respectively), which indicates that only part of Pd loaded on samples is efficient. As shown in Figure 3, NOx storage capacities of 0-Pd and Na-1-Pd are very low. So, Brønsted hydroxyl group and PdOx are inefficient active centers for NOx storage, whereas they do trap NOx in this condition as reported [10,27]. Since the NOx storage capacity of 1-Pd is much higher, isolated Pd2+ ions are likely to be the main active sites for NO trapping.

2.5. NH3 In Situ FTIR

Acidity over samples is probed by NH3 with in situ FTIR, and the result is exhibited in Figure 4. The peak at 1463cm−1 observed in both spectra of 1-Pd and 0-Pd (Figure 4a) is attributed to the vibration of NH4+ in Brønsted hydroxyl groups (NH4+-B) [29]. In the spectrum of 1-Pd, two additional peaks at 1625 cm−1 and 1313cm−1 corresponding to the vibration of NH3 on Lewis acid (NH3-L) [30,31] is observed. Nevertheless, these two peaks do not exist in Na-1-Pd in which isolated Pd2+ ions (Z-Pd2+-Z and Z-Pd(OH)+) are replaced by Na+ completely. So, it is reasonable to believe that Lewis acid is generated from Z-Pd2+-Z and Z-Pd(OH)+ species.
As shown in Figure S3, there is only one peak at 1625cm−1 corresponding to the vibration of NH3-L observed in the spectrum of 1-Pd-80. Since there is only one kind of isolated Pd2+, Z-Pd2+-Z, in this sample as discussed above, the peak at 1625cm−1 should be assigned to the vibration of NH3-L originated from Z-Pd2+-Z. In this case, the peak at 1313cm−1 should be attributed to the vibration of NH3-L originating from Z-Pd(OH)+.
Since Z-Pd2+-Z and Z-Pd(OH)+ are capable of being probed by NH3, the ratio of the height of the peaks at 1625 cm−1 (PZ-Pd2+-Z) and 1313 cm−1 (PZ-Pd(OH)+) in Figure 4b can be defined as the relative content between these two isolated Pd2+ species. Considering that extinction coefficients for NH3 molecules adsorbed on Z-Pd2+-Z and Z-Pd(OH)+ are both constants, they have no effect on the tendency of peak height ratios. So, the extinction coefficient is not considered in this part [32,33,34]. The result is shown in Figure 5a. It is worth nothing that PZ-Pd(OH)+/PZ-Pd2+-Z rises in parallel with the increase of isolated Pd2+, which means that the higher content of isolated Pd2+ leads to a much more obvious increase of Z-Pd(OH)+ than that of Z-Pd2+-Z. So, isolated Pd2+ in 0.2-Pd mainly exists in the form of Z-Pd2+-Z, whereas a large amount of Z-Pd(OH)+ ions are formed in 2-Pd.
Mole ratios of NOx adsorbed to isolated Pd2+ for each sample (marked as NO/Pd2+) are calculated via data in Table 1 and Table 2. Figure 5b is plotted by PZ-Pd(OH)+/PZ-Pd2+-Z on the horizontal axis and NOx/Pd2+ on the vertical. As reported, NOx is believed to be trapped in the form of Pdƍ+-NO [12,15], so the maximum NOx/Pd2+ ratio is 1 in theory. However, the downward trend between PZ-Pd(OH)+/PZ-Pd2+-Z and NOx/Pd2+ means that higher PZ-Pd(OH)+/PZ-Pd2+-Z leads to lower NOx storage capacity of isolated Pd2+, which indicates that the NOx storage capacity of Z-Pd(OH)+ is obviously lower than that of Z-Pd2+-Z in this condition.
A further discussion is given below so as to give a reasonable explanation from the perspective of the structure-function relationship.
In BEA zeolite, nine skeletal T sites with various chemical environments are determined [19]. Among these T sites, T5 and T6 are demonstrated to have the lowest Al substitution energy [35]. Thus, Al substitution on these two sites (marked as AlT5 and AlT6) is preferential, and the amount of AlT5 and AlT6 is larger than that of Al atoms on the other T sites in BEA. Besides, exchange sites formed on these two Al atoms are more active due to the lowest deprotonation energy [36] of them. Meanwhile, AlT5 and AlT6 are located in the meta-position of the same five-membered ring [37], and the co-ion-exchange of protons on these two exchange sites formed on AlT5 and AlT6 is feasible. These characteristics of BEA can explain the ‘exchange preference’ for Z-Pd2+-Z formed in Pd/BEA with lower content of isolated Pd ions to some extent. Besides, with the increase of Pd loading, more exchange sites are taken up. Limited by the amount of protons which are capable of being exchanged, Z-Pd(OH)+ ions, which take up less exchange sites than Z-Pd2+-Z, are preferred to be formed. So, higher content of isolated Pd2+ leads to much more obvious increase of Z-Pd(OH)+ than that of Z-Pd2+-Z.
As discussed above, the maximum NOx/Pd2+ ratio is 1 in theory. As reported by Khivantsev et al. [12], H2O molecules can occupy NOx storage sites due to strong hydration of isolated Pd2+ species. Since isolated Pd2+ is the main active site for NOx storage, the NOx/Pd2+ ratio of Pd/zeolite-based PNA materials will significantly smaller than 1 in the presence of H2O. However, as shown in Figure 5b, NOx/Pd2+ of 0.2-Pd is as much as 1, which indicates that isolated Pd2+ in this sample is not occupied by H2O. As discussed above, isolated Pd2+ in 0.2-Pd mainly exists in the form of Z-Pd2+-Z. Thus Z-Pd2+-Z may be insensitive to H2O in this condition. Furthermore, Z-Pd(OH)+ probably tends to hydrate due to the hydrogen bond interaction between H2O and hydroxy. Accordingly, the difference in NOx storage capacities of Z-Pd(OH)+ and Z-Pd2+-Z is probably caused by the different resistance to H2O in this condition.
The atom utilization of Pd can be represented by mole ratios of NOx adsorbed to total Pd loading. According to Figure 6, the Pd utilization of 0.2-Pd is as much as 100% whereas that of 2-Pd is only 25%. So, it is obvious that lower Pd loading benefits to the increase of Pd utilization. Besides, Khivantsev et al. [13] have reported that the Pd utilization of the 1 wt% Pd/SSZ-13 (Si/Al ratio = 6) sample is 100%, which is much higher than that of 1-Pd (Si/Al ratio = 16) using BEA as a support. Accordingly, zeolite structure and Si/Al ratio can also influence the Pd utilization. In this part, only the influence of Pd loading will be discussed.
As discussed above, lower Pd loading leads to lower content of isolated Pd2+, whereas the proportion of isolated Pd ions in total Pd loading becomes larger. On the on hand, a larger proportion of isolated Pd ions in total Pd loading is beneficial for the improvement of Pd utilization, since isolated Pd ions are identified as the main active center for NOx storage. On the other hand, less isolated Pd2+ results in a preference for the formation of Z-Pd2+-Z which exhibits higher NOx storage capacity. Thus, the atom utilization of Pd can be improved by decreasing Pd loading.
Nevertheless, lower Pd loading also leads to less NOx storage capacities of unit mass of Pd/BEA. Note that the coating amount of catalysts is limited in order to obtain acceptable back pressure. Accordingly, to determine optimum Pd loading for low temperature NOx adsorption, the NOx storage capacity and the atom utilization of Pd should be both considered. As shown in Figure 6, 0.2-Pd and 0.5-Pd exhibits much higher Pd atom utilization than that of the other two samples. Meanwhile, Pd atom utilization of 0.5-Pd is 92%, which is only slightly lower than that of 0.2-Pd. However, the NOx storage capacity of 0.5-Pd is larger than twice of that of 0.2-Pd. So, 0.5wt% should be determined as the optimum Pd loading for Pd/BEA (Si/Al ratio = 16) served as PNA material.

3. Materials and Methods

3.1. Catalyst Preparation

H-BEA zeolite (Si/Al = 16.2) was supplied by Novel Chemistry. Pd/BEA samples with various Pd loading were prepared by incipient wetness impregnation, and Pd(NO3)2 solution (15.47wt% Pd, Heraeus Materials Technology Shanghai Ltd., Shanghai, China) was used. Then, the powders obtained were dried under ambient temperature followed by a 4-h calcination at 550 °C in air. Finally, samples were stabilized at 750 °C for 12 h in air with 10% H2O. Pd loading of each sample was detected by inductively coupled plasma (ICP) analysis (USA Agilent 5100 ICP-OES, Santa Clara, CA, USA). Si/Al ratios were measured by X-ray fluorescence (XRF, ThermoFisher PERFORM’X, Waltham, MA, USA) analysis. Detailed information is exhibited in Table 3. In the following text, samples are abbreviated as x-Pd, where “x” represents the Pd loading of corresponding samples. Besides, the sample treated by Na+ titration was named as Na-1-Pd.

3.2. Catalyst Characterization

Na+ titration was used to quantify isolated Pd ions as reported by Ogura et al. [23]. Each sample was mixed with NaNO3 (Purity above 99.0%, Tianjin Jiangtian Chemical Technology Co., Ltd., Tianjin, China) solution (0.1M) at 80 °C. The solution was stirred for 4 h followed by suction filtration. At last, the powders obtained were washed by deionized water and dried at 100 °C for 6 h. The whole process above was repeated three times. By analyzing the change of Pd loading in each sample after the titration process, the content of isolated Pd can be measured.
NOx storage capacities were measured by standard catalyst evaluation tests carried out in a plug flow reactor system. 0.25 g sample (60–80 mesh) mixed with 0.75 g quartz (60–80 mesh) was loaded in a quartz reactor with a thermocouple. The sample was oxidized in the flow of 10% O2/N2 for 30 min at 500 °C and was cooled to 80 °C in N2. Then the flow (200 ppm NOx, 200 ppm CO, 5% CO2, 5%H2O, 10% O2, balanced with N2) mixed in the bypass in advance was introduced into the reactor. The NOx adsorption stage continued for 3 min. After that, the sample was heated to 500 °C with a ramping rate of 10 °C/min in the flow of N2. Gas concentrations were measured by an online MKS MultiGas 2030 FTIR gas analyzer in the whole process. Besides, a space velocity of 28,800 h−1 was adopted.
Ex-situ FTIR was carried out on a Nicolet iS10 FTIR spectrometer equipped with a liquid N2 cooled mercury cadmium telluride (MCT) detector to identify the existence of Pd ions. The sample (20 mg) was pressed into a self-supporting wafer with a diameter of 13 mm and was inserted into a cell sealed with ZnSe windows connected with a gas manifold. KBr was used to obtain background spectra. All samples, as well as KBr, were oxidized in the flow of 10% O2/N2 for 30 min at 500 °C in advance. Spectra were obtained at 200 °C in N2.
In situ FTIR was carried out on a Nicolet iS10 FTIR spectrometer, too. Test temperatures and feed compositions varied according to the needs of different experiments, and detailed information will be reported below.

4. Conclusions

In this work, two isolated Pd2+ ions, Z-Pd2+-Z and Z-Pd(OH)+, on exchange sites of zeolites are confirmed as the main active sites for NO trapping in cold-start applications. Lower Pd loading leads to a lower content of isolated Pd2+ whereas the proportion of isolated Pd ions in total Pd loading becomes larger. Lower content of isolated Pd2+ further leads to a sharp decline of Z-Pd(OH)+, which is attributed to the ‘exchange preference’ for Z-Pd2+-Z in BEA. Besides, the higher NOx storage capacity of Z-Pd2+-Z is demonstrated compared with that of Z-Pd(OH)+, which is caused by the different resistance to H2O. In conclusion, the atom utilization of Pd can be improved by using lower Pd loading. 0.5wt% should be determined as the optimum Pd loading for Pd/BEA (Si/Al ratio = 16) serving as PNA material among all samples.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/3/247/s1, Table S1: Detailed information of 1-Pd-80, Figure S1: CO in situ FTIR spectra of 1-Pd-80, Figure S2: NOx adsorption profiles of 1-Pd and 1-Pd-80, Figure S3: NH3 in situ FTIR spectra of 1-Pd-80.

Author Contributions

Conceptualization, B.Z.; methodology, J.W. (Jun Wang), M.S., and J.W. (Jianqiang Wang); software, B.Z.; validation, J.W. (Jianqiang Wang); formal analysis, J.W. (Jianqiang Wang); investigation, B.Z and J.W. (Jiaming Wang).; writing—original draft preparation, B.Z.; writing—review and editing, B.Z.; visualization, B.Z.; supervision, J.W. (Jun Wang); resources, M.S.; project administration, M.S.; funding acquisition, M.S.

Funding

This research was funded by National Key Research and Development Program, grant number 2017YFC0211002; State Key Laboratory of Advanced Technologies for Comprehensive Utilization of Platinum Metals, grant number SKL-SPM-2018017.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnson, T.; Joshi, A. Review of vehicle engine efficiency and emissions. SAE Int. J. Engines 2018, 11, 1307–1330. [Google Scholar]
  2. Gao, F.; Tang, X.; Yi, H.; Zhao, S.; Li, C.; Li, J.; Shi, Y.; Meng, X. A Review on Selective Catalytic Reduction of NOx by NH3 over Mn–Based Catalysts at Low Temperatures: Catalysts, Mechanisms, Kinetics and DFT Calculations. Catalysts 2017, 7, 199. [Google Scholar] [CrossRef]
  3. Kubiak, L.; Matarrese, R.; Castoldi, L.; Lietti, L.; Daturi, M.; Forzatti, P. Study of N2O Formation over Rh- and Pt-Based LNT Catalysts. Catalysts 2016, 6, 36. [Google Scholar] [CrossRef]
  4. Chen, H.-Y.; Collier, J.E.; Liu, D.; Mantarosie, L.; Durán-Martín, D.; Novák, V.; Rajaram, R.R.; Thompsett, D. Low temperature NO storage of zeolite supported Pd for low temperature diesel engine emission control. Catal. Lett. 2016, 146, 1706–1711. [Google Scholar] [CrossRef]
  5. Chen, H.-Y.; Mulla, S.; Weigert, E.; Camm, K.; Ballinger, T.; Cox, J.; Blakeman, P. Cold Start Concept (CSC™) A Novel Catalyst for Cold Start Emission Control. SAE Int. J. Fuels Lubr. 2013, 6, 372–381. [Google Scholar] [CrossRef]
  6. Ryou, Y.; Lee, J.; Cho, S.J.; Lee, H.; Kim, C.H.; Kim, D.H. Activation of Pd/SSZ-13 catalyst by hydrothermal aging treatment in passive NO adsorption performance at low temperature for cold start application. Appl. Catal. B 2017, 212, 140–149. [Google Scholar] [CrossRef]
  7. Lee, J.; Ryou, Y.; Hwang, S.; Kim, Y.; Cho, S.J.; Lee, H.; Kim, C.H.; Kim, D.H. Comparative study of the mobility of Pd species in SSZ-13 and ZSM-5, and its implication for their activity as passive NOx adsorbers (PNAs) after hydro-thermal aging. Catal. Sci. Technol. 2019, 9, 163–173. [Google Scholar] [CrossRef]
  8. Ryou, Y.; Lee, J.; Lee, H.; Kim, C.H.; Kim, D.H. Effect of various activation conditions on the low temperature NO adsorption performance of Pd/SSZ-13 passive NOx adsorber. Catal. Today 2019, 320, 175–180. [Google Scholar] [CrossRef]
  9. Vu, A.; Luo, J.; Li, J.; Epling, W.S. Effects of CO on Pd/BEA Passive NOx Adsorbers. Catal. Lett. 2017, 147, 745–750. [Google Scholar] [CrossRef]
  10. Zheng, Y.; Kovarik, L.; Engelhard, M.H.; Wang, Y.; Wang, Y.; Gao, F.; Szanyi, J. Low-Temperature Pd/Zeolite Passive NOx Adsorbers: Structure, Performance, and Adsorption Chemistry. J. Phys. Chem. C 2017, 121, 15793–15803. [Google Scholar] [CrossRef]
  11. Lee, J.; Ryou, Y.; Cho, S.J.; Lee, H.; Kim, C.H.; Kim, D.H. Investigation of the active sites and optimum Pd/Al of Pd/ZSM–5 passive NO adsorbers for the cold-start application: Evidence of isolated-Pd species obtained after a high-temperature thermal treatment. Appl. Catal. B 2018, 226, 71–82. [Google Scholar] [CrossRef]
  12. Khivantsev, K.; Gao, F.; Kovarik, L.; Wang, Y.; Szanyi, J. Molecular Level Understanding of How Oxygen and Carbon Monoxide Improve NOx Storage in Palladium/SSZ-13 Passive NOx Adsorbers: The Role of NO+ and Pd (II)(CO)(NO) Species. J. Phys. Chem. C 2018, 122, 10820–10827. [Google Scholar] [CrossRef]
  13. Khivantsev, K.; Jaegers, N.R.; Kovarik, L.; Hanson, J.C.; Tao, F.; Tang, Y.; Zhang, X.; Koleva, I.Z.; Aleksandrov, H.A.; Vayssilov, G.N.; et al. Achieving Atomic Dispersion of Highly Loaded Transition Metals in Small-Pore Zeolite SSZ-13: High-Capacity and High-Efficiency Low-Temperature CO and Passive NOx Adsorbers. Angew. Chem. 2018, 130, 16914–16919. [Google Scholar] [CrossRef]
  14. Mei, D.; Gao, F.; Szanyi, J.; Wang, Y. Mechanistic insight into the passive NOx adsorption in the highly dispersed Pd/HBEA zeolite. Appl. Catal. A 2019, 569, 181–189. [Google Scholar] [CrossRef]
  15. Khivantsev, K.; Jaegers, N.R.; Kovarik, L.; Prodinger, S.; Derewinski, M.A.; Wang, Y.; Gao, F.; Szanyi, J. Palladium/Beta zeolite passive NOx adsorbers (PNA): Clarification of PNA chemistry and the effects of CO and zeolite crystallite size on PNA performance. Appl. Catal. A Gen. 2019, 569, 141–148. [Google Scholar] [CrossRef]
  16. Ryou, Y.; Lee, J.; Kim, Y.; Hwang, S.; Lee, H.; Kim, C.H.; Kim, D.H. Effect of reduction treatments (H2 vs. CO) on the NO adsorption ability and the physicochemical properties of Pd/SSZ-13 passive NOx adsorber for cold start application. Appl. Catal. A 2019, 569, 28–34. [Google Scholar] [CrossRef]
  17. Porta, A.; Pellegrinelli, T.; Castoldi, L.; Matarrese, R.; Morandi, S.; Dzwigaj, S.; Lietti, L. Low Temperature NOx Adsorption Study on Pd-Promoted Zeolites. Top. Catal. 2018, 61, 2021–2034. [Google Scholar] [CrossRef]
  18. Theis, J.R.; Lambert, C.K. Mechanistic assessment of low temperature NOx adsorbers for cold start NOx control on diesel engines. Catal. Today 2019, 320, 181–195. [Google Scholar] [CrossRef]
  19. Newsam, J.; Treacy, M.M.; Koetsier, W.; De Gruyter, C. Structural characterization of zeolite beta. Proc. R. Soc. Lond. A 1988, 420, 375–405. [Google Scholar] [CrossRef]
  20. Gao, F.; Washton, N.M.; Wang, Y.; Kollár, M.; Szanyi, J.; Peden, C.H. Effects of Si/Al ratio on Cu/SSZ-13 NH 3-SCR catalysts: Implications for the active Cu species and the roles of Brønsted acidity. J. Catal. 2015, 331, 25–38. [Google Scholar] [CrossRef]
  21. Giordanino, F.; Vennestrøm, P.N.; Lundegaard, L.F.; Stappen, F.N.; Mossin, S.; Beato, P.; Bordiga, S.; Lamberti, C. Characterization of Cu-exchanged SSZ-13: A comparative FTIR, UV-Vis, and EPR study with Cu-ZSM-5 and Cu-β with similar Si/Al and Cu/Al ratios. Dalton Trans. 2013, 42, 12741–12761. [Google Scholar] [CrossRef] [PubMed]
  22. Gao, F.; Peden, C.H.F. Recent Progress in Atomic-Level Understanding of Cu/SSZ-13 Selective Catalytic Reduction Catalysts. Catalysts 2018, 8, 140. [Google Scholar] [Green Version]
  23. Ogura, M.; Hayashi, M.; Kage, S.; Matsukata, M.; Kikuchi, E. Determination of active palladium species in ZSM-5 zeolite for selective reduction of nitric oxide with methane. Appl. Catal. B 1999, 23, 247–257. [Google Scholar] [CrossRef]
  24. Kumar, A.; Buttry, D.A. Influence of Halide Ions on Anodic Oxidation of Ethanol on Palladium. Electrocatalysis 2016, 7, 201–206. [Google Scholar] [CrossRef]
  25. Yuan, N.; Pascanu, V.; Huang, Z.; Valiente, A.; Heidenreich, N.; Leubner, S.; Inge, A.K.; Gaar, J.; Stock, N.; Persson, I. Probing the evolution of palladium species in Pd@ MOF catalysts during the Heck coupling reaction: An operando X-ray absorption spectroscopy study. J. Am. Chem. Soc. 2018. [Google Scholar] [CrossRef] [PubMed]
  26. Palazov, A.; Chang, C.; Kokes, R. The infrared spectrum of carbon monoxide on reduced and oxidized palladium. J. Catal. 1975, 36, 338–350. [Google Scholar] [CrossRef]
  27. Aylor, A.W.; Lobree, L.J.; Reimer, J.A.; Bell, A.T. Investigations of the Dispersion of Pd in H-ZSM-5. J. Catal. 1997, 172, 453–462. [Google Scholar] [CrossRef]
  28. Okumura, K.; Amano, J.; Yasunobu, N.; Niwa, M. X-ray absorption fine structure study of the formation of the highly dispersed PdO over ZSM-5 and the structural change of Pd induced by adsorption of NO. J. Phys. Chem. B 2000, 104, 1050–1057. [Google Scholar] [CrossRef]
  29. Barzetti, T.; Selli, E.; Moscotti, D.; Forni, L. Pyridine and ammonia as probes for FTIR analysis of solid acid catalysts. J. Chem. Soc. Faraday Trans. 1996, 92, 1401–1407. [Google Scholar] [CrossRef]
  30. Pawelec, B.; Campos-Martin, J.; Cano-Serrano, E.; Navarro, R.; Thomas, S.; Fierro, J. Removal of PAH compounds from liquid fuels by Pd catalysts. Environ. Sci. Technol. 2005, 39, 3374–3381. [Google Scholar] [CrossRef] [PubMed]
  31. Centeno, M.; Carrizosa, I.; Odriozola, J. NO–NH3 coadsorption on vanadia/titania catalysts: Determination of the reduction degree of vanadium. Appl. Catal. B 2001, 29, 307–314. [Google Scholar] [CrossRef]
  32. Niwa, M.; Nishikawa, S.; Katada, N. IRMS–TPD of ammonia for characterization of acid site in β-zeolite. Microporous Mesoporous Mater. 2005, 82, 105–112. [Google Scholar] [CrossRef]
  33. Datka, J.; Gil, B.; Kubacka, A. Acid properties of NaH-mordenites: Infrared spectroscopic studies of ammonia sorption. Zeolites 1995, 15, 501–506. [Google Scholar] [CrossRef]
  34. Ding, K.; Gulec, A.; Johnson, A.M.; Schweitzer, N.M.; Stucky, G.D.; Marks, L.D.; Stair, P.C. Identification of active sites in CO oxidation and water-gas shift over supported Pt catalysts. Science 2015, 350, 189. [Google Scholar] [CrossRef] [PubMed]
  35. Bare, S.R.; Kelly, S.D.; Sinkler, W.; Low, J.J.; Modica, F.S.; Valencia, S.; Corma, A.; Nemeth, L.T. Uniform catalytic site in Sn-β-Zeolite determined using x-ray absorption fine structure. J. Am. Chem. Soc. 2005, 127, 12924–12932. [Google Scholar] [CrossRef] [PubMed]
  36. Jones, A.J.; Iglesia, E. The strength of Brønsted acid sites in microporous aluminosilicates. ACS Catal. 2015, 5, 5741–5755. [Google Scholar] [CrossRef]
  37. Higgins, J.; LaPierre, R.B.; Schlenker, J.; Rohrman, A.; Wood, J.; Kerr, G.; Rohrbaugh, W. The framework topology of zeolite beta. Zeolites 1988, 8, 446–452. [Google Scholar] [CrossRef]
Figure 1. (a) Ex-situ FTIR spectra of 0-Pd, 1-Pd, Na-1-Pd; (b) Ex-situ FTIR spectra of 2-Pd, 1-Pd, 0.5-Pd, 0.2-Pd (Temperature: 200 °C; Flow: N2, 1 L/min).
Figure 1. (a) Ex-situ FTIR spectra of 0-Pd, 1-Pd, Na-1-Pd; (b) Ex-situ FTIR spectra of 2-Pd, 1-Pd, 0.5-Pd, 0.2-Pd (Temperature: 200 °C; Flow: N2, 1 L/min).
Catalysts 09 00247 g001
Figure 2. CO in situ FTIR spectra of 2-Pd, 1-Pd, 0.5-Pd, 0.2-Pd (Temperature: 80 °C; Flow: 1000 ppm CO, balanced with N2, 500 mL/min).
Figure 2. CO in situ FTIR spectra of 2-Pd, 1-Pd, 0.5-Pd, 0.2-Pd (Temperature: 80 °C; Flow: 1000 ppm CO, balanced with N2, 500 mL/min).
Catalysts 09 00247 g002
Figure 3. NOx adsorption profiles (Temperature: 80 °C; Flow: NOx, CO, H2O, O2, CO2, balanced with N2, 1 L/min).
Figure 3. NOx adsorption profiles (Temperature: 80 °C; Flow: NOx, CO, H2O, O2, CO2, balanced with N2, 1 L/min).
Catalysts 09 00247 g003
Figure 4. (a) NH3 in situ FTIR spectra of 0-Pd, 1-Pd, Na-1-Pd; (b) NH3 in situ FTIR spectra of 2-Pd, 1-Pd, 0.5-Pd, 0.2-Pd (Temperature: 80 °C; Flow: 500 ppm NH3, balanced with N2, 500 mL/min).
Figure 4. (a) NH3 in situ FTIR spectra of 0-Pd, 1-Pd, Na-1-Pd; (b) NH3 in situ FTIR spectra of 2-Pd, 1-Pd, 0.5-Pd, 0.2-Pd (Temperature: 80 °C; Flow: 500 ppm NH3, balanced with N2, 500 mL/min).
Catalysts 09 00247 g004
Figure 5. (a) tendency between isolated Pd2+ and PZ-Pd(OH)+/PZ-Pd2+-Z; (b) tendency between PZ-Pd(OH)+/PZ-Pd2+-Z and NOx/Pd2+.
Figure 5. (a) tendency between isolated Pd2+ and PZ-Pd(OH)+/PZ-Pd2+-Z; (b) tendency between PZ-Pd(OH)+/PZ-Pd2+-Z and NOx/Pd2+.
Catalysts 09 00247 g005
Figure 6. Tendency between Pd loading and the atom utilization of Pd.
Figure 6. Tendency between Pd loading and the atom utilization of Pd.
Catalysts 09 00247 g006
Table 1. The content of Pd ions on exchange sites of each sample.
Table 1. The content of Pd ions on exchange sites of each sample.
Catalyst0-Pd0.2-Pd0.5-Pd1-Pd2-Pd
Isolated Pd (wt %)0.000.220.530.870.91
Isolated Pd/Pd loading (%)0.0095.796.477.742.7
Table 2. NOx storage capacities in NOx, CO, H2O, O2 and CO2.
Table 2. NOx storage capacities in NOx, CO, H2O, O2 and CO2.
Catalyst0-Pd0.2-Pd0.5-Pd1-Pd2-Pd
NOx storage capacity (μmol/gcat)3.721.647.553.350.1
Table 3. Pd loading of samples.
Table 3. Pd loading of samples.
Catalyst0-Pd0.2-Pd0.5-Pd1-Pd2-PdNa-1-Pd
ICP Pd (wt %)00.230.551.122.130.25

Share and Cite

MDPI and ACS Style

Zhang, B.; Shen, M.; Wang, J.; Wang, J.; Wang, J. Investigation of Various Pd Species in Pd/BEA for Cold Start Application. Catalysts 2019, 9, 247. https://doi.org/10.3390/catal9030247

AMA Style

Zhang B, Shen M, Wang J, Wang J, Wang J. Investigation of Various Pd Species in Pd/BEA for Cold Start Application. Catalysts. 2019; 9(3):247. https://doi.org/10.3390/catal9030247

Chicago/Turabian Style

Zhang, Beibei, Meiqing Shen, Jianqiang Wang, Jiaming Wang, and Jun Wang. 2019. "Investigation of Various Pd Species in Pd/BEA for Cold Start Application" Catalysts 9, no. 3: 247. https://doi.org/10.3390/catal9030247

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