Next Article in Journal
Synergy between Sulfonic Functions and Ru Nanoparticles Supported on Activated Carbon for the Valorization of Cellulose into Sorbitol
Next Article in Special Issue
Structure-Sensitive Behavior of Supported Vanadia-Based Catalysts for Combustion of Soot
Previous Article in Journal
Machine Learning: A Suitable Method for Biocatalysis
Previous Article in Special Issue
Investigation of Co3O4 and LaCoO3 Interaction by Performing N2O Decomposition Tests under Co3O4-CoO Transition Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Assessment of Zeolite Framework Effect for Low-Temperature NOX Adsorbers

1
Laboratory of Catalysis and Catalytic Processes, Department of Energy, Politecnico di Milano, Via La Masa 34, 20156 Milano, Italy
2
Department of Chemistry, NIS Center and INSTM Reference Center, Università di Torino, Via P. Giuria 7, 10125 Torino, Italy
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(6), 962; https://doi.org/10.3390/catal13060962
Submission received: 30 April 2023 / Revised: 28 May 2023 / Accepted: 30 May 2023 / Published: 1 June 2023

Abstract

:
Pd-promoted zeolites (Y, ZSM-5, FER, SSZ-13) were prepared and characterized to analyze their properties as low-temperature NOx adsorbers. The samples were investigated by BET and XRD and by in situ FT-IR spectroscopy of CO and NO adsorption to probe the Pd sites and the nature of the adsorbed NOx species. The NOx adsorption/desorption performances at low temperatures were examined by microreactor measurements upon NO/O2 adsorption followed by TPD in the presence of water and carbon dioxide. It was enlightened that: (i) the zeolite framework influences the Pd dispersion: the smaller the zeolite cage, the higher the Pd dispersion, irrespective of the Si/Al ratio. Accordingly, the following Pd dispersion order has been observed, inversely to the zeolite cage size: Pd/SSZ-13 > Pd/ZSM-5 ~ Pd/FER >> Pd/Y; (ii) Pd is present as isolated Pdn+ species and in PdOx particles; (iii) the Pd dispersion governs the NOx storage capacity: the smaller the zeolite cage, the higher the Pd dispersion and the storage capacity; (iv) NO adsorbs mainly in the form of Pd nitrosyls and nitrates; (v) NO desorption occurs both at a temperature below 200 °C and in a high-temperature range (near 350 °C).

1. Introduction

The severe NOx emission regulations enforced in the automotive sector [1] call for highly efficient NOx reduction techniques. Currently adopted technologies for lean-burn engines are the NOx-storage reduction (NSR) technique, also referred to as Lean NOx traps (LNT) [2,3], and the urea selective catalytic reduction (SCR) [4]. In fact, both techniques are not efficient for the reduction in NOx emissions at low temperatures, as required by recent regulations, especially for hybrid vehicles where the endothermic engine is continuously subjected to start and stop cycles [5,6].
The use of passive NOx adsorber (PNA) systems installed upstream of the SCR or LNT systems has been recently proposed to mitigate cold-start NOx emissions. In this configuration, PNAs store NOx at low temperatures; the trapped NOx are then released at higher temperatures when the NOx reduction catalysts placed downstream effectively accomplish the NOx abatement [7,8,9,10].
In 2016, Johnson Matthey [11] reported a very efficient formulation for PNA systems based on acidic zeolites doped with Pd. In that work, the authors compared the behavior of different zeolite structures, including MFI (Mobil-type five), CHA (chabasite), and BEA (Beta). These systems showed remarkable NOx storage capacity at 100 °C, with release of the trapped NOx at a temperature suitable for practical application, i.e., near 200 °C. It was found that the framework structure of the zeolites affects the NOx storage capacity and the desorption temperature; in particular, the authors pointed out that at 100 °C the zeolites with the largest pore size exhibit the highest storage capacity, and the lowest NOx desorption temperature during heating.
Starting from this pioneering work, research on zeolite-based LTNAs has intensified in recent years [12,13,14,15,16,17]. Zheng et al. [13] compared Pd-doped ZSM-5 (Zeolite Socony Mobil-type 5) and SSZ-13 (framework type code CHA) zeolites, finding out that in these materials Pd is present as atomically dispersed species in the cationic sites of the zeolite, and as PdOx particles on the external surfaces. In particular, it was reported that while Pd is largely dispersed in ZSM-5, the small pore opening of SSZ-13 inhibits Pd diffusion, favoring the formation of PdOx particles on the external surfaces. Lee et al. [16] also showed differences in the Pd mobility within ZSM-5 and SSZ-13: during oxidative and hydrothermal treatments, the smaller pore size of SSZ-13 contributed to the lower Pd mobility with respect to ZSM-5. Along similar lines, Pace et al. [17] compared Beta and CHA zeolites, finding that these materials show different distributions of Pd species. Notably, it is generally agreed that Pd loadings near 1 wt% lead to the formation of isolated Pd ions, which are recognized as active sites for low-temperature NOx storage [18,19,20,21].
The literature survey points out the extreme complexity of these systems as the nature of Pd surface sites is influenced by many factors, including the zeolite structure. Aiming at a better understanding of how the morphological characteristic of the zeolite affects the nature of the Pd sites and their involvement in the NOx storage, in this study, a large-pore zeolite (Y-zeolite, with Faujasite FAU structure), two medium-pore zeolites (ZSM-5, with MFI structure, and Ferrierite FER), and a small-pore zeolite (SSZ-13, with CHA structure) have been considered. These zeolites have been doped with Pd, and the effect of the zeolite framework on the NOx adsorption/desorption performances has been investigated. In particular, NOx adsorption and desorption experiments under simulated exhaust conditions (i.e., in the presence of both H2O and CO2) have been performed, along with in situ FT-IR spectroscopy studies to point out the NOx adsorption activity/nature of the zeolite substrate relationships. The effect of the Si/Al ratio on the sample performances could also be considered in that zeolites with the same pore size but different Si/Al ratios (FER and ZSM-5) have been considered and compared.

2. Results and Discussion

2.1. Sample Characterization

2.1.1. Structural and Morphological Characterization

Y zeolite exhibits the Faujasite (FAU) framework; it is a large-pore zeolite (pore openings of 7.4 Å), and its structure contains a 3D pore system limited by 12-member rings. Ferrierite (FER framework) and ZSM-5 (MFI framework) are medium-pore zeolites whose structure contains a 2D system of intersecting pores limited by 8- and 10-member rings for FER and 10-member rings for ZSM-5. SSZ-13 is a small-pore zeolite (ring opening 3.6 Å) with a chabazite framework; its structure contains a 3D pore system limited by 8-member rings. The main characteristics of these zeolitic frameworks are reported in Table 1, together with the compositional and textural characterization.
As expected, all the zeolites are characterized by a large surface area and very low pore volume; the addition of Pd does not substantially modify either the surface area or the pore volume.
X-ray Diffraction (XRD) patterns obtained on the bare zeolites and on the Pd-doped samples (Figure 1) are in agreement with the diffraction patterns for the reference zeolite frameworks [22]. The zeolite structure is maintained in all the catalysts even after Pd addition; additionally, peaks corresponding to crystalline PdO, whose most intense peak should be observed at 33.8° (JCPDS 75-0584), are not observed, suggesting that particles are too small to be identified by the XRD technique or absent.
SEM (Scanning Electron Microscopy) images reported in Figure 2 clearly show the maintenance of the zeolite morphologies also after Pd addition; moreover, the associated energy-dispersive X-ray (EDX) analysis confirms the Si/Al ratio given by the supplier and the Pd loading near the nominal value (1 wt%, see Table 1). Pd loading has also been confirmed by ICP-MS analysis in the case of the SSZ-13 sample.

2.1.2. FT-IR Characterization

CO adsorption—In Figure 3, the FT-IR spectra of CO (20 mbar) adsorbed at Room Temperature (RT) on the activated samples are displayed. The spectra are characterized by complex envelopes of bands, which are related to Pd carbonyl species in different oxidation states and coordination, as discussed in our previous paper [23], where the precise assignment of each component in the spectrum of Pd/SSZ-13 was made. As discussed in ref. [23], with reference to the Pd/SSZ-13 sample (trace a), it is possible to divide the bands into different spectral regions, namely 2220–2170 cm−1 and 2170–2110 cm−1, associated with linear carbonyls of Pd2+ and Pd+, respectively [13,24,25,26]. The absorption near 2228 cm−1 is assigned to carbonyls of extra-framework Al3+ [27].
In the spectral region 2220–2170 cm−1 (where very strong bands are present in the case of Pd/SSZ-13 sample, Figure 3 trace a), it has to be considered that, due to the σ-donation contribution in the Pd2+-CO bond, the higher the band frequency, the more coordinatively unsaturated the site. With this in mind and on the base of the literature data, the bands at 2211 and 2192 cm−1 are assigned to di-carbonyls (νasym(CO) and νsym(CO) modes, respectively) of isolated Pd2+ ions formed by ion exchange with the Brønsted acid sites of the zeolite [24,28]; the shoulder at 2187 cm−1 is instead assigned to CO adsorbed on Pd2+ sites of PdOx particles [23].
Along similar lines, for the spectral region 2170–2110 cm−1, due to the π-backdonation contribution to the Pd+-CO bond, the lower the band frequency, the more coordinatively unsaturated the site. For this reason, in the spectrum of Pd/SSZ-13 (Figure 3, trace a), the band at 2163 cm−1 is assigned to CO adsorbed on Pd+ sites belonging to PdOx particles. The bands at 2150, 2142, 2127, and 2111 cm−1 are related to the presence of two different Pd+ di-carbonyl species (two couple of bands, one component of the couple related to νasym(CO) mode, the other component to νsym(CO) mode), whose presence is well demonstrated in our previous paper [23]. This implies, for Pd/SSZ-13, the existence of two different isolated Pd+ sites formed by ion exchange with the Brønsted acid sites of the zeolite: indeed, Pd+ can be located into both 8- and 6-member rings of the zeolite framework [17,29].
For Pd/ZSM-5 and Pd/FER samples (Figure 3, traces b and c, respectively), the bands observed in the same regions are less intense, but their assignments are similar to that examined for the Pd/SSZ-13 sample. In particular, for the Pd/ZSM-5 sample, the bands at 2214 and 2194 cm−1 are assigned to di-carbonyls of isolated Pd2+ ions; the band at 2173 cm−1 to CO bonded on Pd2+ belonging to PdO particles; the bands at 2142 and 2127 cm−1 to di-carbonyls of isolated Pd+ sites. For the Pd/FER sample (Figure 3, trace c), the broadness of the bands means higher surface heterogeneity. Notably, the two bands related to di-carbonyls of isolated Pd2+ ions at about 2214 and 2194 cm−1 show shoulders, pointing out the presence of exchanged Pd2+ ions located in different zeolite positions (e.g., 10- and 8-ring channels).
Finally, it is worth noting that the spectrum of Pd/Y (Figure 3, trace d) shows very weak bands, enlarged in the inset of Figure 3. Only bands related to di-carbonyls of isolated Pd+ ions are present at 2142 and 2111 cm−1, respectively. The band at 2228 cm−1 related to carbonyls of extra-framework Al3+ is present as well.
The area of the carbonyl bands is proportional to the number of available Pd sites, both isolated Pd species inside the zeolite framework and Pd species belonging to external PdOx particles. Inspection of Figure 3 clearly indicates that the availability of the Pd sites (i.e., the dispersion) is highest for the Pd/SSZ-13 sample and follows the order: Pd/SSZ-13 > Pd/ZSM-5 ~ Pd/FER >> Pd/Y. This suggests a relationship between pore size and the Pd dispersion: the smaller the pore size, the higher the Pd dispersion. In this respect, it has to be recalled that Zheng et al. [13] characterized Pd on zeolites with different pore sizes and found that Pd is largely present as isolated ions in the ZSM-5 framework, while the small pore opening of SSZ-13 inhibits the exchange with Brønsted acid sites. In our case, the availability of the Pd sites follows a different order, as reported above, apparently in contrast with the results of Zheng et al. This discrepancy might be related to the different pretreatment of the samples: in fact, while they evaluated the dispersion of isolated Pd ions on the freshly calcined samples, at variance, we evaluated the dispersion after an activation treatment in vacuum and oxygen at 500 °C.
In addition, our results seem to indicate that the Pd dispersion is not driven by the Si/Al ratio for zeolites having similar pore size: in fact, the Pd/ZSM-5 and Pd/FER samples, characterized by similar pore size but different Si/Al ratio, exhibit similar Pd dispersion. This suggests that the Pd dispersion, for the Pd loading conditions here adopted, is primarily affected by the zeolite pore size and only to a minor extent by the Si/Al ratio. At variance, Khivantsev et al. [28] analyzed the Pd dispersion over 1 wt% Pd-doped zeolites having different Si/Al ratios (and the same framework) but prepared in a different way and observed that the Si/Al ratio significantly affects the Pd dispersion. In particular, these authors, using a modified-ion exchange method to prepare 1 wt% Pd on NH4-SSZ-13 with Si/Al ratios ranging from 6 to 30, observed the increasing formation of PdO nanoparticles (i.e., low dispersion) upon increasing the Si/Al ratio. This is very likely related to the different preparation method used by these authors, which significantly differs with respect to our preparation route.
NO and NO/O2 adsorption—Traces “a” in Figure 4 show FT-IR spectra of NO adsorbed at RT on activated Pd/Y (A), Pd/ZSM-5 (B), Pd/FER (C), and Pd/SSZ-13 (D). In all cases, bands in the region of 1900–1800 cm−1 are present with intensities that vary from one sample to another. These spectral features are well documented in the literature and are assigned to nitrosyl species of Pd2+ and Pd+, even if there is no definitive consensus on the precise assignments of the single bands to either the Pd2+ or Pd+ species [28,29,30,31,32]. Accordingly, it is worth noting that, contrarily to CO, NO adsorption is not able to distinguish the different coordinations of Pdn+ ions (isolated ions or sites on particles). In fact, the complex envelopes of bands present in the spectra of adsorbed CO do not find correspondence in the spectral features shown upon NO adsorption. However, the intensities of the Pd nitrosyl bands in the various samples well match those of carbonyls observed upon CO adsorption. This confirms the Pd dispersion order already discussed in the case of CO adsorption, i.e., Pd/SSZ-13 > Pd/ZSM-5 ~ Pd/FER >> Pd/Y.
Along with Pd nitrosyl bands, for Pd/FER and Pd/SSZ-13 samples (Figure 4, traces “a”, sections C and D, respectively), peaks at 2250–2247, 2161–2155, 1972 (only for Pd/FER), and at about 1630 cm−1 are also present. The bands at 2250–2247 cm−1 and at 1972 cm−1 are related to adsorbed N2O and N2O3, respectively, present as impurities in the gas [33] or formed as detailed below. On the other hand, the bands at 2161–2155 and 1630 cm−1 are assigned to the stretching mode of NO+ and to the bending mode of adsorbed molecular water, respectively. Hadjiivanov et al. [34] proposed the formation of these species upon the reaction of NO with Brønsted acid sites of an H-ZSM-5 in the presence of oxygen:
2NO + ½O2 + 2Olattice–H+ → H2O + 2Olattice–NO+
Oxygen is not present in the gas phase, but the samples were oxidized before NO admission: accordingly, oxygen species are likely present on the Pd phase. This is supported by the fact that the bands related to NO+ and adsorbed molecular water show lower intensities on the sample with lower Pd dispersion, i.e., Pd/FER with respect to Pd/SSZ-13. As a matter of fact, NO+ species are observable with very low intensities also in the Pd/ZSM-5 sample (NO+ band at 2128 cm−1 and water band at 1635 cm−1), and they are not present (except for a very small amount of water) for Pd/Y sample, for which Pd site availability is almost nil. It is important to point out that the presence of reactive oxygen species on the Pd phase can also induce the formation of NO2 that, along with NO, can form the adduct N2O3 (indeed observed for Pd/FER catalyst) and/or can form NO+ and surface nitrates according to the following reactions [35,36,37]:
2NO2 → NO+ + NO3
NO+ + NO2 + O2−lattice → NO3 + NO(g)
Consequently, the observed NO+ can be derived from both reactions (1) and (2), along with nitrates that should show characteristic vibrational modes in the region below 1700 cm−1. As a matter of fact, nitrates could contribute to the band at about 1630 cm−1, assigned to the bending mode of molecular water. All these observations were confirmed by oxygen admission, as described in the following section.
Starting from the condition obtained in the presence of 2 mbar of NO (spectra “a”), 10 mbar of oxygen were added: traces “b” and “c” in the different sections of Figure 4 show the evolution of bands after 5 and 60 min, respectively. First of all, the increase in bands below 1700 cm−1, assigned to surface nitrates formed by reactions (2) and (3) mentioned above, where NO2 is formed upon NO oxidation, is well evident for all the samples. In this region, the band at 1630 cm−1 is related to both nitrates and molecular water formed by reaction (1). Moreover, reactions (1) and (2) account for the formation of NO+, whose band suddenly increases after oxygen admission for Pd/SSZ-13 and Pd/FER samples. Even in the presence of oxygen, the band of NO+ at 2128 cm−1 remains very weak for Pd/ZSM-5 and not detectable for Pd/Y sample. Indeed, reaction (1) needs Brønsted acid sites to occur: as a matter of fact, the intensity of the band and hence the amount of adsorbed NO+ is directly related to the Brønsted acid site amount, i.e., to the Si/Al ratio of the zeolite. Moreover, it has to be taken into account the higher stability of adsorbed NO+ in the smaller channels of SSZ-13 and FER [38].
Concerning nitrosyl species, oxygen admission causes an intensity increase in the bands for Pd/SSZ-13 and Pd/FER samples. Moreover, for Pd/SSZ-13 sample, nitrosyl bands blue-shift upon oxygen admission. The latter can be ascribed to the increase in the average positive charge of the Pd sites, especially of the Pd+ sites, induced by oxygen adsorption, which could also cause changes in the absorption coefficients of nitrosyl species, justifying the increase in band intensities.
The areas (integrated intensities) of the bands related to nitrosyls and nitrates are indicative of the Pd dispersion, as also evidenced by CO adsorption. Indeed, along with nitrosyl formation, the presence of Pd is required to form nitrate species as well. In fact, nitrates are not formed in the absence of Pd, as evidenced by the comparison between the NO/O2 adsorption on Pd/SSZ-13 and the bare SSZ-13 shown in Figure 5. This comparison also points out the markedly lower amount of NO+ (band at 2161 cm−1) on Pd/SSZ-13 with respect to the bare zeolite. In fact, NO+ is formed on the zeolite Brønsted acid sites, and their concentration has been reduced as a result of Pd exchange with the formation of Pd2+ and Pd+ sites.

2.2. NOx Uptake and Desorption Experiments

2.2.1. NO/O2 Adsorption Phase

The NOx storage performances of Pd/zeolite were determined under realistic conditions, i.e., at 80 °C with a NO/O2 mixture in the presence of CO2 and H2O.
The concentration traces obtained upon NOx adsorption and subsequent Temperature Programmed Desorption (TPD) are shown in Figure 6A–D. For all samples, upon the addition of the NO/O2 mixture, the NO concentration recorded at the reactor outlet rapidly increases without showing any dead time, eventually reaching the inlet concentration value with a dynamic response that depends on the catalyst sample. The response is very fast in the case of Pd/Y (Figure 6A) and sluggish for the Pd/SSZ-13 sample (Figure 6D), which, in fact, shows a large NOx uptake evaluated from the amounts of desorbed NOx during the TPD (Figure 7). Lower amounts of NOx are stored on the other Pd-doped samples, and the following order is observed: Pd/SSZ-13 > Pd/ZSM-5 ~ Pd/FER >> Pd/Y (Figure 7). This ranking is in good agreement with the Pd dispersion evaluated from FT-IR measurements of CO and NO adsorption.
During NO adsorption, NO2 evolution is also observed, although with different features. In fact, a small NO2 peak is observed in the case of Pd/ZSM-5 and Pd/FER (Figure 6B, C, respectively), with evolution of 8 and 5 µmol/gcat NO2, respectively, whereas a sluggish NO2 evolution is seen in the case of Pd/SSZ-13. The overall amount of NO2 evolved, in this case, is near 19 µmol/gcat. Almost no NO2 formation is observed in the case of Pd/Y.
NO2 evolution observed during the adsorption phase is likely related to the reduction of Pd2+ to Pd+ operated by NO according to the following overall reaction [13,39,40,41]:
3NO + 2[Pd(II)OH]+ → NO2 + 2[Pd(I)NO]+ + H2O
According to the stoichiometry of reaction (4), 3 moles of NO are consumed and lead to the formation of 1 mol of NO2 evolved in the gas phase and to the storage of 2 moles of Pd nitrosyls. Reaction (4) also envisages the reduction of Pd2+ with the formation of Pd+ nitrosyls upon NO adsorption.
Notably, as expected from the stoichiometry of reaction (4), the amounts of evolved NO2 represent half of the amounts of formed nitrosyls. However, as will be shown in the following section, the amounts of stored NOx are higher than those expected from the stoichiometry of reaction (4) based on the NO2 formation. This points out that other routes are active in the storage of NOx, possibly involving the direct NO uptake on Pd2+ sites as well.

2.2.2. NOx Desorption Phase

During the thermal desorption (TPD) after the NO/O2 storage at 80 °C, NOx release occurs mainly as NO, although NO2 is also observed to desorb. The amounts of evolved NOx were evaluated from the area of the TPD peaks, and the results are summarized in Figure 7, along with the estimated desorbed NOx/Pd molar ratio.
For all samples, at the beginning of the temperature ramp a desorption of H2O is observed due to the heating ramp; notably, the amounts of water desorbed is inversely proportional to the pores size, i.e., the bigger the pores, the lower the amount of desorbed water. This is very likely related to water capillary condensation that is favored within the smaller pores.
In the case of the Pd/SSZ-13 sample (Figure 6D), three main desorption regions can be identified: (i) from the closure of the NO/O2 feed up to 120 °C a tail is apparent due to the physisorbed species, counting for 13 µmol/gcat NO; (ii) from 120 °C to 240 °C a well-defined peak is observed with a maximum at 180 °C, composed of NO and NO2 (32 µmol/gcat and 6 µmol/gcat, respectively); and (iii) a high temperature peak (in the range of 240–500 °C) formed only by NO (41.4 µmol/gcat) and centered near 360 °C. The overall amount of evolved NOx is near 93 µmol/gcat (see Figure 7). As previously recalled, the overall amounts of stored NOx are well above those expected from the amounts of evolved NO2, based on the stoichiometry of reaction (4). This indicates the occurrence of other adsorption reactions besides reaction (4), e.g., direct NO uptake on Pd2+ with no reduction of Pd2+ to Pd+ (and no NO2 formation).
The Pd/ZSM-5 and Pd/FER present two desorption regions (Figure 6B, C, respectively), the former in the low-temperature region below 240 °C and the latter in the high-temperature region above 300 °C; negligible amounts of NO2 are observed for these samples. The overall amounts of desorbed NOx are 41 and 50 µmol/gcat, respectively. Finally, in the case of Pd/Y (Figure 6A), any NOx desorption is observed, as expected, due to the negligible storage capacity of this sample. These results are well in line with the CO and NO FTIR adsorption experiments previously discussed, showing a higher NO adsorption capacity for Pd/SSZ-13.
From the amounts of stored NOx, we have also estimated the adsorbed NOx/Pd molar ratio (Figure 7) under the hypothesis that all NOx are stored over Pd. In the case of the Pd/SSZ-13 sample, a value of desorbed NOx near 93 µmol/gcat was evaluated, corresponding to a NOx/Pd molar ratio of 0.86. This calls for a large utilization of Pd, thus indicating a good dispersion of the metal. The Pd/ZSM-5 and Pd/FER showed lower stored NOx/Pd values, whereas, as already stated the storage capacity for Pd/Y is negligible.
Concerning the stability of the adsorbed NOx species, low- and high-temperature contributions are observed in all cases, excluding the Pd/Y sample that shows no NO adsorption. NOx desorption peaks at low and high temperatures (near 150 °C and 360 °C) have also been observed, e.g., by Gupta et al. over Pd/SSZ-13 [39]; the authors attributed these two peaks to at least two different adsorption sites with distinct binding energies. In particular, they attributed the lower temperature NO peak to NO bound to ZPd2+Z (where Z- denotes the zeolite anionic aluminum site (Al-O-Si)-) and the higher temperature NO peak to NO bound to ZPd+. Similar conclusions have been drawn by Villamaina et al. [37], who pointed out that more stable nitrosyls are formed over Pd+ species. Moreover, as reported by Castoldi et al. in the case of Pd/FER [42], and by Matarrese et al. [23] for Pd/SSZ-13, the low-temperature NOx desorption peak can also be associated with the decomposition of nitrate species. On this basis, it is speculated that the low-T desorption peaks can be associated with the decomposition of nitrates and/or Pd2+ nitrosyls, whereas the high desorption peak is related to the decomposition of stable Pd+ nitrosyls.
The results reported above clearly indicate that the zeolite framework affects both the amounts and the stability of the stored NOx. The characterization data previously discussed point out the existence of different Pd species in terms of different structures and oxidation states, including isolated Pd cations and PdOx particles on the external zeolite surface, whose distribution varies with the zeolite framework [11,13]. Additionally, according to FT-IR characterization, the investigated samples show different Pd dispersion, and evidence has been provided on the fact that this parameter is influenced by the pore size of the zeolitic framework. The Pd dispersion is the key parameter affecting the NOx storage, having observed a clear correlation between NOx adsorption capacity and Pd dispersion. Notably, a very high NO adsorption capacity has been measured for the Pd/SSZ-13 sample, in spite of the fact that Pd species belonging to PdOx particles have been identified by CO adsorption. This suggests that also non-isolated Pd species participate in the NOx storage.

3. Materials and Methods

3.1. Sample Preparation

The Pd-doped zeolites (Pd/Y, Pd/ZSM-5, Pd/FER, and Pd/SSZ-13; Pd loading 1% w/w) used in this study were prepared starting from commercial samples (Zeolyst International, Kansas City, KS, USA and ACS Material LLC, Pasadena, CA 91106) in their NH4- or H-form. The zeolites in the NH4-form were transformed into H-form by calcination at 700 °C in the air (heating rate: 1 °C/min).
Then, the Pd-doped zeolite samples were prepared by impregnation of the zeolite supports in excess of solvent using aqueous Pd(NO3)2 (solution at 4.5% Pd, Sigma-Aldrich, St. Louis, MO, USA), followed by drying in air at 80 °C overnight [42]. The dried precursors were calcined from room temperature up to 500 °C (rate: 1 °C/min, at 500 °C 3 h) and then from 500 °C up to 750 °C (rate 5 °C/min, at 750 °C 2 h).

3.2. Sample Characterization

BET analysis was performed by adsorption–desorption of nitrogen at −196 °C using a Micromeritics Tristar 300 instrument (Alfatest SRL, Rome, Italy). Before the measurement, samples were outgassed at 120 °C and 0.1 torr for 3 h.
Powder X-ray diffraction (XRD) patterns of bare and Pd/zeolite samples were obtained using a Bruker D8 diffractometer () with Cu Kα radiation. The XRD patterns were obtained on the as prepared samples in the range of 10–50° (2θ) with a step size of 0.05° at a speed of 12.5°/min.
SEM and elemental compositions analysis of bare zeolite and Pd-doped samples were performed using a ZEISS EVO 50 instrument (Oberkochen, Germany) equipped with energy-dispersive X-ray analyzer (EDX) model Quantax 200 6/30 (Bruker, Karlsruhe, Germany).
ICP-MS analysis has been performed on selected samples to evaluate the Pd loading. The Perkin Elmer NexION 2000 single quadrupole ICP-MS instrument (Shelton, CT, USA) was employed upon calibration with a limit of quantification (LOQ) of 0.040 ppb for Pd.
IR spectra were run in transmission mode at RT on a Perkin-Elmer FT-IR 2000 spectrophotometer (Shelton, CT, USA) equipped with a Hg-Cd-Te cryodetector, working in the range of wavenumbers 7200–580 cm−1 at a resolution of 2 cm−1 (number of scans ~60). Self-supporting disks (10–15 mg cm−2) were obtained from the powder samples and placed in a quartz IR cell allowing thermal treatments in vacuum or in a controlled atmosphere. The samples were activated by treatment at 500 °C in vacuum, followed by oxidation in dry O2 (40 mbar) at the same temperature. The samples were cooled down to room temperature (RT) in oxygen. CO adsorption was performed to characterize the nature and availability of Pd sites; NO and NO/O2 adsorptions were performed to evidence the species formed during NOx storage. IR spectra were recorded at RT in static conditions before and after interaction with CO (20 mbar), NO (2 mbar), and NO/O2 (1:5, pNO = 2 mbar) mixture. In the figures, spectra are reported as difference spectra, where the subtrahend spectrum is that of the sample after the activation, before the admission of the gas.

3.3. Gas phase Analysis during NO/O2 Adsorption at Low Temperature

NOx adsorption experiments were conducted using 100 mg of powdered samples, sieved at 70–100 µm, and loaded in a quartz microreactor with an internal diameter of 8 mm. The outlet of the reactor is directly connected to a mass spectrometer (ThermoStar 200, Pfeiffer Vacuum GmbH, Aßlar, Germany)), a UV-analyzer specific for NO, NO2, NH3 (Limas 11 HW, ABB, Zurig, Swizzerland), and a micro gas-chromatograph (3000 A, Agilent, Santa Clara, CA 95051 United States). In a typical experiment, the sample was initially pre-treated at 500 °C for 30 min under a flow of O2 (3% v/v), CO2 (2% v/v), and H2O (2.5% v/v) (balance He) at a flow rate of 100 Ncm3/min. The sample was then cooled down to the adsorption temperature (50 °C, 80 °C, or 120 °C) without changing the gas flow composition, and at this temperature 300 ppm of NO were stepwise added to the gas mixture for 75 min. Then, the NO supply was closed, and the sample was heated under temperature programming (TPD) up to 500 °C (heating rate 10 °C/min).

4. Conclusions

In this study, Pd-promoted zeolites have been prepared by impregnation of the zeolite supports in excess of solvent and investigated for potential use in low-temperature NOx adsorption. Four zeolitic supports with different structures have been considered (Y, ZSM-5, FER, and SSZ-13). These structures have different pore openings, i.e., large (Y), medium (ZSM-5 and FER), and small (SSZ-13); samples with different Si/Al ratios have also been considered. All the samples have been characterized and tested in the adsorption of NO under relevant operating conditions.
The FT-IR characterization using CO as a probe molecule evidenced the presence of a variety of Pdn+ sites with different coordination. In particular, isolated Pd cations formed by ion exchange with the Brønsted acid sites of the zeolite and on the surface of PdOx particles have been evidenced. The Pd dispersion estimated by CO and NO adsorption follows the order: Pd/SSZ-13 > Pd/ZSM-5 ~ Pd/FER >> Pd/Y, i.e., increases upon decreasing the pore opening. These results suggest that the pore size plays a primary role in controlling the Pd dispersion. At variance, the relevance of the Si/Al ratio in affecting the Pd dispersion seems to be minor. In other words, the zeolite structure (and related pore openings) is apparently the key parameter to drive the Pd dispersion.
Reactivity experiments carried out under the relevant conditions pointed out that the amounts of adsorbed NOx follow the same order as the Pd dispersion. Notably, a very high NOx/Pd molar ratio has been estimated for Pd/SSZ-13, despite the presence of PdOx clusters and/or particles pointed out in this sample besides well-dispersed Pd sites. This suggests that non-isolated Pd species belonging to particles participate in the NOx storage along with isolated well-dispersed species.

Author Contributions

P.T.: Investigation, Methodology. L.C.: Investigation, Conceptualization, Writing—original, review & editing, Methodology. S.M.: Investigation, Writing—original, review & editing, Conceptualization, Funding acquisition. R.M.: Writing—review & editing, Conceptualization. L.L.: Supervision, Conceptualization, Writing— review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available in Catalysts 2023, 13.

Acknowledgments

S.M. and P.T. acknowledge the support from Project CH4.0 under the MUR (Italian Ministry for the University) program “Dipartimenti di Eccellenza 2023–2027” (CUP: D13C22003520001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Körfer, T. Potential of Advanced, Combined Aftertreatment Systems for Light-Duty Diesel Engines to Meet Upcoming EU and US Emission Regulation; SAE Technical Paper 2013-24-0163; SAE International: Warrendale, PA, USA, 2013. [Google Scholar]
  2. Roy, S.; Baiker, A. NOx Storage-Reduction Catalysis: From Mechanism and Materials Properties to Storage-Reduction Performance. Chem. Rev. 2009, 109, 4054–4091. [Google Scholar] [CrossRef]
  3. Epling, W.S.; Campbell, L.E.; Yezerets, A.; Currier, N.W.; II, J.E.P. Overview of the Fundamental Reactions and Degradation Mechanisms of NOx Storage/Reduction Catalysts. Catal. Rev. 2004, 46, 163–245. [Google Scholar] [CrossRef]
  4. Olsson, L.; Wijayanti, K.; Leistner, K.; Kumar, A.; Joshi, S.Y.; Kamasamudram, K.; Currier, N.W.; Yezerets, A. A kinetic model for sulfur poisoning and regeneration of Cu/SSZ-13 used for NH3-SCR. Appl. Catal. B Environ. 2016, 183, 394–406. [Google Scholar] [CrossRef]
  5. Theis, J.R.; Lambert, C.K. An assessment of low temperature NOx adsorbers for cold-start NOx control on diesel engines. Catal. Today 2015, 258, 367–377. [Google Scholar] [CrossRef]
  6. Gao, Z.; Kim, M.-Y.; Choi, J.-S.; Daw, C.S.; Parks, J.E.; Smith, D.E. Cold-start emissions control in hybrid vehicles equipped with a passive adsorber for hydrocarbons and nitrogen oxides. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2012, 226, 1396–1407. [Google Scholar] [CrossRef]
  7. Tamm, S.; Andonova, S.; Olsson, L. Silver as Storage Compound for NOx at Low Temperatures. Catal. Lett. 2014, 144, 674–684. [Google Scholar] [CrossRef]
  8. Jones, S.; Ji, Y.; Crocker, M. Ceria-Based Catalysts for Low Temperature NOx Storage and Release. Catal. Lett. 2016, 146, 909–917. [Google Scholar] [CrossRef]
  9. Gu, Y.; Epling, W.S. Passive NOx adsorber: An overview of catalyst performance and reaction chemistry. Appl. Catal. A Gen. 2019, 570, 1–14. [Google Scholar] [CrossRef]
  10. Lee, J.; Theis, J.R.; Kyriakidou, E.A. Vehicle emissions trapping materials: Successes, challenges, and the path forward. Appl. Catal. B Environ. 2019, 243, 397–414. [Google Scholar] [CrossRef]
  11. 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]
  12. Murata, U.; Morita, T.; Wada, K.; Ohno, H. NOx Trap Three-Way Catalyst (N-TWC) Concept: TWC with NOx Adsorption Properties at Low Temperatures for Cold-Start Emission Control. SAE Int. J. Fuels Lubr. 2015, 8, 454–459. [Google Scholar] [CrossRef]
  13. Zheng, Y.; Kovarik, L.; Engelhard, M.H.; Wang, Y.L.; 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]
  14. Ryou, Y.; Lee, J.; Lee, H.; Kim, C.H.; Kim, D.H. Low temperature NO adsorption over hydrothermally aged Pd/CeO2 for cold start application. Catal. Today 2018, 307, 93–101. [Google Scholar] [CrossRef]
  15. 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 Environ. 2018, 226, 71–82. [Google Scholar] [CrossRef]
  16. 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]
  17. Pace, R.B.; Lardinois, T.M.; Ji, Y.; Gounder, R.; Heintz, O.; Crocker, M. Effects of Treatment Conditions on Pd Speciation in CHA and Beta Zeolites for Passive NOx Adsorption. ACS Omega 2021, 6, 29471–29482. [Google Scholar] [CrossRef]
  18. Wang, Y.; Shi, X.; Liu, Z.; Shan, Y.; Shi, W.; Yu, Y.; He, H. The study of Pd-SSZ-13 as low-temperature passive NO adsorber materials: High dispersal of Pd in small-pore CHA zeolites by thermal treatment. Appl. Catal. B Environ. 2023, 324, 122254. [Google Scholar] [CrossRef]
  19. Theis, J.R.; Ura, J.; Getsoian, A.B.; Prikhodko, V.Y.; Thomas, C.R.; Pihl, J.A.; Lardinois, T.M.; Gounder, R.; Wei, X.; Ji, Y.; et al. Effect of framework Al pairing on NO storage properties of Pd-CHA passive NOx adsorbers. Appl. Catal. B Environ. 2023, 322, 122074. [Google Scholar] [CrossRef]
  20. Li, Y.; Chen, D.; Xu, X.; Wang, X.; Kang, R.; Fu, M.; Guo, Y.; Chen, P.; Li, Y.; Ye, D. Cold-Start NO(x) Mitigation by Passive Adsorption Using Pd-Exchanged Zeolites: From Material Design to Mechanism Understanding and System Integration. Environ. Sci. Technol. 2023, 57, 3467–3485. [Google Scholar] [CrossRef]
  21. Huang, S.; Wang, Q.; Shan, Y.; Shi, X.; Liu, Z.; He, H. Effects of Si/Al Ratio on Passive NOx Adsorption Performance over Pd/Beta Zeolites. Molecules 2023, 28, 3501. [Google Scholar] [CrossRef]
  22. Treacy, M.M.J.; Higgins, J.B. Collection of Simulated XRD Powder Patterns for Zeolites, 5th ed.; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  23. Matarrese, R.; Castoldi, L.; Morandi, S.; Ticali, P.; Valsania, M.C.; Lietti, L. NO adsorption/desorption pathways over Pd/SSZ-13 low temperature NOx adsorbers investigated by operando FT-IR spectroscopy and microreactor study. Appl. Catal. B 2023, 331, 122723. [Google Scholar] [CrossRef]
  24. Khivantsev, K.; Jaegers, N.R.; Koleva, I.Z.; Aleksandrov, H.A.; Kovarik, L.; Engelhard, M.; Gao, F.; Wang, Y.; Vayssilov, G.N.; Szanyi, J. Stabilization of Super Electrophilic Pd+2 Cations in Small-Pore SSZ-13 Zeolite. J. Phys. Chem. C 2020, 124, 309–321. [Google Scholar] [CrossRef]
  25. Hadjiivanov, K.I.; Vayssilov, G.N. Characterization of oxide surfaces and zeolites by carbon monoxide as an IR probe molecule. Adv. Catal. 2002, 47, 307–511. [Google Scholar]
  26. 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]
  27. Onida, B.; Geobaldo, F.; Testa, F.; Crea, F.; Garrone, E. FTIR investigation of the interaction at 77 K of diatomic molecular probes on MCM-22 zeolite. Micropor. Mesopor. Mat. 1999, 30, 119–127. [Google Scholar] [CrossRef]
  28. Khivantsev, K.; Jaegers, N.R.; Kovarik, L.; Hanson, J.C.; Tao, F.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. Int. Ed. Engl. 2018, 57, 16672–16677. [Google Scholar] [CrossRef]
  29. Mandal, K.; Gu, Y.; Westendorff, K.S.; Li, S.; Pihl, J.A.; Grabow, L.C.; Epling, W.S.; Paolucci, C. Condition-Dependent Pd Speciation and NO Adsorption in Pd/Zeolites. ACS Catal. 2020, 10, 12801–12818. [Google Scholar] [CrossRef]
  30. 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]
  31. Chakarova, K.; Ivanova, E.; Hadjiivanov, K.; Klissurski, D.; Knözinger, H. Co-ordination chemistry of palladium cations in Pd-H-ZSM-5 as revealed by FTIR spectra of adsorbed and co-adsorbed probe molecules (CO and NO). Phys. Chem. Chem. Phys. 2004, 006, 3702–3709. [Google Scholar] [CrossRef]
  32. Kim, P.; Van der Mynsbrugge, J.; Aljama, H.; Lardinois, T.M.; Gounder, R.; Head-Gordon, M.; Bell, A.T. Investigation of the Modes of NO Adsorption in Pd/H-CHA. Appl. Catal. B Environ. 2022, 304, 120992. [Google Scholar] [CrossRef]
  33. Hadjiivanov, K.I. Identification of Neutral and Charged NxOy Surface Species by IR Spectroscopy. Catal. Rev. 2000, 42, 71–144. [Google Scholar] [CrossRef]
  34. Hadjiivanov, K.; Saussey, J.; Freysz, J.L.; Lavalley, J.C. FT-IR study of NO+O2 co-adsorption on H-ZSM-5: Re-assignment of the 2133 cm−1 band to NO+ species. Catal. Lett. 1998, 52, 103–108. [Google Scholar] [CrossRef]
  35. Vos, A.M.; Mignon, P.; Geerlings, P.; Thibault-Starzyk, F.; Schoonheydt, R.A. Probing the basicity of zeolite frameworks with N2O4: A DFT approach. Micropor. Mesopor. Mat. 2006, 90, 370–376. [Google Scholar] [CrossRef]
  36. Szanyi, J.; Hun Kwak, J.; Moline, R.A.; Peden, C.H.F. The adsorption of NO2 and the NO + O2 reaction on Na-Y, FAU: An in situ FTIR investigation. Phys. Chem. Chem. Phys. 2003, 5, 4045–4051. [Google Scholar] [CrossRef]
  37. Perdana, I.; Creaser, D.; Öhrman, O.; Hedlund, J. A comparison of NOx adsorption on Na, H and BaZSM-5 films. Appl. Catal. B Environ. 2007, 72, 82–91. [Google Scholar] [CrossRef]
  38. 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]
  39. Gupta, A.; Kang, S.B.; Harold, M.P. NOx uptake and release on Pd/SSZ-13: Impact of feed composition and temperature. Catal. Today 2021, 360, 411–425. [Google Scholar] [CrossRef]
  40. Villamaina, R.; Iacobone, U.; Nova, I.; Tronconi, E.; Ruggeri, M.P.; Mantarosie, L.; Collier, J.; Thompsett, D. Mechanistic insight in NO trapping on Pd/Chabazite systems for the low-temperature NOx removal from Diesel exhausts. Appl. Catal. B Environ. 2021, 284, 119724. [Google Scholar] [CrossRef]
  41. 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 Gen. 2019, 569, 181–189. [Google Scholar] [CrossRef]
  42. Castoldi, L.; Matarrese, R.; Morandi, S.; Ticali, P.; Lietti, L. Low-temperature Pd/FER NOx adsorbers: Operando FT-IR spectroscopy and performance analysis. Catal. Today 2021, 360, 317–325. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of zeolite (red line) and Pd/zeolite on different frameworks (black line).
Figure 1. X-ray diffraction (XRD) patterns of zeolite (red line) and Pd/zeolite on different frameworks (black line).
Catalysts 13 00962 g001
Figure 2. SEM images of bare zeolites and of the corresponding Pd-doped samples.
Figure 2. SEM images of bare zeolites and of the corresponding Pd-doped samples.
Catalysts 13 00962 g002
Figure 3. FT-IR spectra of CO (20 mbar) adsorbed at RT on activated Pd/SSZ-13 (a—black line), Pd/ZSM-5 (b—blue line), Pd/FER (c—red line), and Pd/Y (d—green line).
Figure 3. FT-IR spectra of CO (20 mbar) adsorbed at RT on activated Pd/SSZ-13 (a—black line), Pd/ZSM-5 (b—blue line), Pd/FER (c—red line), and Pd/Y (d—green line).
Catalysts 13 00962 g003
Figure 4. FT-IR spectra of NO (2 mbar, curves a—red) adsorbed at RT on activated Pd/Y (A), Pd/ZSM-5 (B), Pd/FER (C), and Pd/SSZ-13 (D). Curves b (blue) and c (black): NO/O2 mixture (1:5, pNO = 2 mbar) after 5 and 60 min of contact at RT, respectively.
Figure 4. FT-IR spectra of NO (2 mbar, curves a—red) adsorbed at RT on activated Pd/Y (A), Pd/ZSM-5 (B), Pd/FER (C), and Pd/SSZ-13 (D). Curves b (blue) and c (black): NO/O2 mixture (1:5, pNO = 2 mbar) after 5 and 60 min of contact at RT, respectively.
Catalysts 13 00962 g004
Figure 5. FT-IR spectra of NO/O2 mixture (1:5, pNO = 2 mbar) adsorbed at RT after 60 min of contact on activated SSZ-13 (a) and Pd/SSZ-13 (b).
Figure 5. FT-IR spectra of NO/O2 mixture (1:5, pNO = 2 mbar) adsorbed at RT after 60 min of contact on activated SSZ-13 (a) and Pd/SSZ-13 (b).
Catalysts 13 00962 g005
Figure 6. NO/O2 adsorption at 80 °C and subsequent TPD experiment over the activated zeolite-based samples: (A) Pd/Y; (B) Pd/ZSM-5; (C) Pd/FER; and (D) Pd/SSZ-13. Experimental conditions: storage phase NO (300 ppm) + O2 (3%) + CO2 (2%) + H2O (2.5%) in He; TPD in O2 (3%) + CO2 (2%) + H2O (2.5%) in He from 80 °C to 500 °C (10°/min). Note that NO2 traces have been multiplied by 5 (panel C) or 2 (panel D).
Figure 6. NO/O2 adsorption at 80 °C and subsequent TPD experiment over the activated zeolite-based samples: (A) Pd/Y; (B) Pd/ZSM-5; (C) Pd/FER; and (D) Pd/SSZ-13. Experimental conditions: storage phase NO (300 ppm) + O2 (3%) + CO2 (2%) + H2O (2.5%) in He; TPD in O2 (3%) + CO2 (2%) + H2O (2.5%) in He from 80 °C to 500 °C (10°/min). Note that NO2 traces have been multiplied by 5 (panel C) or 2 (panel D).
Catalysts 13 00962 g006
Figure 7. Quantitative analysis of NOx desorbed (dark blue) and NOx/Pd ratio (light blue) after adsorption of NO/O2 at 80 °C.
Figure 7. Quantitative analysis of NOx desorbed (dark blue) and NOx/Pd ratio (light blue) after adsorption of NO/O2 at 80 °C.
Catalysts 13 00962 g007
Table 1. Morphological and chemico-physical characteristics of bare zeolites and Pd-doped ones.
Table 1. Morphological and chemico-physical characteristics of bare zeolites and Pd-doped ones.
SampleFramework Structures and Pore SystemPore Sizes
(Å)
Si/Al Ratio aPd Loading (%) aBET (m2/g)Vp (cm3/g)
YFAU; 3D;
12-ring channels
7.430//6160.27
Pd/Y1.076120.27
ZSM-5MFI; 2D;
10-ring channels
5.4–5.640//4070.10
Pd/ZSM-51.073900.12
FERFER; 2D;
10 and 8-ring channels
4.3–5.511//3020.13
Pd/FER1.13120.10
SSZ-13CHA; 3D;
8-ring channels
3.6–3.710//5580.09
Pd/SSZ-131.14 b5780.06
a from EDX analysis. b 1.03% from ICP-MS analysis.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Castoldi, L.; Morandi, S.; Ticali, P.; Matarrese, R.; Lietti, L. An Assessment of Zeolite Framework Effect for Low-Temperature NOX Adsorbers. Catalysts 2023, 13, 962. https://doi.org/10.3390/catal13060962

AMA Style

Castoldi L, Morandi S, Ticali P, Matarrese R, Lietti L. An Assessment of Zeolite Framework Effect for Low-Temperature NOX Adsorbers. Catalysts. 2023; 13(6):962. https://doi.org/10.3390/catal13060962

Chicago/Turabian Style

Castoldi, Lidia, Sara Morandi, Pierfrancesco Ticali, Roberto Matarrese, and Luca Lietti. 2023. "An Assessment of Zeolite Framework Effect for Low-Temperature NOX Adsorbers" Catalysts 13, no. 6: 962. https://doi.org/10.3390/catal13060962

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