Skip to main content
Erschienen in: Topics in Catalysis 13-14/2023

Open Access 31.10.2022 | Original Paper

Low-Temperature NOx Reduction by H2 on Mo-Promoted Pt/ZrO2 Catalysts in Lean Exhaust Gases

verfasst von: D. Schröder, S. Kureti

Erschienen in: Topics in Catalysis | Ausgabe 13-14/2023

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

This article aims to improve the low-temperature H2-deNOx performance of the active Pt/ZrO2 catalyst using MoOx as a promoter. For this purpose, a systematic series of Pt/ZrO2 samples were prepared with a Pt content of 0.25 wt% and Mo loads from 0 to 10 wt%. The samples were physico-chemically characterized by means of powder X-ray diffraction, N2 physisorption, temperature-programmed desorption of CO and NH3, Raman spectroscopy and diffuse reflectance infrared spectroscopy using NH3 as probe molecule, while the H2-deNOx efficiency was investigated in a lean synthetic exhaust. The Pt/ZrO2 catalyst with a Mo load of 3 wt% showed the best performance, including H2-deNOx between 80 °C and 150 °C, a maximum NOx conversion of 90% and N2 selectivity up to 78%. Isolated MoOx species predominately present at Mo loads below 4 wt% were found to act as structural promoter by stabilizing the BET surface area, while also providing smaller Pt particles and more active Pt sites, respectively. By contrast, the aggregated Mo oxide moieties found at higher Mo loads exhibit a clearly weaker promotional effect. The structure–activity-selectivity correlations also suggest that the promoter additionally enables a SCR-related mechanistic pathway to be followed, including the spill-over of NHx species from the Pt sites to strong Lewis acid sites in the case of highly dispersed MoOx entities followed by reaction with NOx.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1007/​s11244-022-01720-4.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

For the reduction of CO2, which is the most significant greenhouse gas, combustion engines are progressively being replaced by fuel cell and battery electric vehicles. However, mobile and stationary combustion engines are still essential to maintain transportation and the energy supply in the long term. Thus, one special challenge lies in replacing fossil with sustainable fuels such as biofuels or “e-fuels” originating from power-to-X strategies. Moreover, H2 is a CO2-neutral fuel, particularly when it is produced from H2O electrolysis powered by wind and solar energy. Combustion engines are most efficient under lean burn conditions, but the output of pollutants such as hydrocarbons, CO and NOx remains an issue [1]. Consequently, automotive pollution control techniques are required to meet international emission standards. To remove NOx from O2-rich exhaust gases, SCR (Selective Catalytic Reduction) and NOx storage catalyst (NSC) technologies are well established. These operate effectively between approx. 175 °C and 500 °C. However, both procedures perform only to a limited extent at the lower temperatures relevant for cold starts and low load engine operation such as city driving.
NOx reduction with H2 (H2-deNOx) using Pt and Pd catalysts is the only available technique (2 NO + 2 H2 → N2 + 2 H2O) providing outstanding conversion in the low-temperature regime of lean engine exhaust gases (< 175 °C). However, one current issue of H2-deNOx is the formation of by-products like N2O and NH3 as exemplarily known for Pt/Al2O3, Pt/H-ZSM-5 and Pt/SiO2 catalysts [24]. Marked N2 selectivity above 90% has been reported for Pt/MgO-CeO2 [5, 6], Pt/La0.7Sr0.2Ce0.1FeO3 [7] and Pd/LaCoO3 catalysts [8]. Additionally, WOx-promoted Pt catalysts supported by TiO2 [9], HZSM-5 [10] and ZrO2 [11] have been shown to be highly active, including the preferential formation of N2, while Mo-modified Pt/SiO2 and Pt/Al2O3 samples [12] also exhibited improved N2 selectivity. Moreover, Pt catalysts modified with metal oxide promoters such as 1st and 2nd group elements (Na, K, Ba, etc.) [13, 14] and transition metals (Co, Mn, W, Mo, etc.) [11, 15, 16] enhanced N2 formation during H2-deNOx.
The present article aims to improve the low-temperature H2-deNOx performance under O2-rich conditions using Pt/ZrO2, which was recently reported to be a highly active catalyst. For this purpose, MoOx was evaluated as a potential promoter [12], requiring a systematic series of Pt/ZrO2 catalysts with different Mo loads. The prepared samples were physico-chemically characterized and studied in terms of their lean H2-deNOx activity and selectivity. From these investigations, structure–activity relations were derived to reveal the determining properties of the Mo promoter, which control the H2-deNOx activity leading to N2.

2 Experimental

2.1 Catalyst Preparation

The samples were prepared by means of incipient wetness impregnation using monoclinic ZrO2 as a support (Saint-Gobain NorPro). First, the Mo promoter was introduced by an aqueous solution of (NH4)6Mo7O24 · 4 H2O (Merck) with different concentrations to obtain Mo loads between 0 and 10 wt%. After drying for at least 4 h at 90 °C in air, each sample was impregnated with a Pt(NO3)2 solution (Chempur), resulting in a Pt proportion of 0.25 wt.% referred to the bare support. Note that preliminary investigations, shown in section S1, indicated no significant re-dissolution of the Mo precursor during the subsequent impregnation with the aqueous Pt solution. Pt activation of the 2 g samples was carried out in a flowing mixture of 10 vol% H2 and 90 vol% N2 (400 ml/min, STP) while ramping the temperature to 300 °C at a rate of 1.7 K/min and holding that temperature for 30 min. Finally, the samples were calcined for 5 h at 500 °C in static air. In this paper, the Pt-containing catalysts are denoted as Pt/xMo/ZrO2, where x refers to the loading of Mo.

2.2 Catalyst Characterization

The characterization of the catalysts was performed prior to the H2-deNOx investigations, since exposure to the lean model exhaust gas at a maximum temperature of 280 °C is unlikely to alter the samples calcined at 500 °C before. This aspect was checked in preliminary studies.
Powder X-ray diffraction (PXRD) was performed on a D8 Discover (Bruker-AXS) with a Bragg–Brentano configuration, Fe-filtered Co-Kα radiation and a VANTEC-1 detector. Diffractograms were taken from 10° to 80° in 2θ mode with a step width of 0.5° and an integration time of 250 s. The PXRD patterns were evaluated using the Powder Diffraction Files (PDF) database.
N2 physisorption was carried out on a TriStar II (Micromeritics). Respective sample was degassed in vacuo (10–1 mbar) at 350 °C for 16 h to remove adsorbed components, then the N2 adsorption isotherms were taken at − 196 °C. The BET surface area (SBET) was determined from the adsorption data recorded at p/p0 ratios from 0.05 to 0.20.
Laser Raman spectroscopy (LRS) was conducted with an inVia Raman microscope (Renishaw) equipped with a Nd:YAG laser (532 nm, 100 mW), a grating with 1800 lines per mm and a CCD array detector. The spectra were collected under ambient conditions from 10 to 1800 cm−1 at a resolution of 1.6 cm−1, an exposure time of 120 s and a laser power of 10 mW, accumulating 3 scans per spectrum.
Temperature-programmed desorption of CO (CO-TPD) was performed on a home-made laboratory rig. A granulated sample (300 mg, 125–250 µm) was introduced into the U-shaped quartz glass tube reactor (i.d. 8 mm), packed as a fixed bed and pre-treated in 2 vol% O2 (He balance) at 450 °C for 30 min (400 ml/min, STP). Then, it was cooled to − 196 °C with liquid N2 to reduce spill-over effects from Pt to the promoter or support during the CO exposure. Saturation with CO was performed by applying pulses of 200 ppm CO in He for 9 s each followed by intermediate purging with He for 3 min (400 ml/min, STP). After the final flushing with He, the TPD was started (150 ml/min, STP) by heating the sample to 550 °C at a rate of 20 K/min. The temperature was recorded by a K-type thermocouple located directly above the sample. Desorbing CO and CO2 were monitored using a non-dispersive infrared (NDIR) spectrometer (X-Stream Enhanced XEGK, Emerson). Blank CO-TPD studies of the bare Mo/ZrO2 samples provided CO desorption profiles very similar to the Pt/xMo/ZrO2 samples without significant amounts of CO2. The evolution of CO2 upon TPD is ascribed to the oxidation of CO by the MoOx/ZrO2 surface. The blank COx profiles were subtracted from the respective TPD traces of the Pt catalysts. The resulting COx desorption profiles were integrated to quantify the available Pt sites (na(Pt)), assuming a CO/Pt adsorption stoichiometry of 1 [17]. From na(Pt) and the total abundance of platinum (nPt), the Pt dispersion (DPt = na(Pt) / nPt) was estimated, while the mean size of Pt particles (dPt = 6 VPt/(aPt ∙ DPt)) was calculated assuming spheres; VPt is the volume of a Pt atom present in the bulk metal (15.10 Å3), aPt represents the surface area of a Pt atom located on a polycrystalline surface (8.07 Å2) [17].
The temperature-programmed desorption of NH3 (NH3-TPD) was also carried out on a home-made rig. A granulated sample (500 mg, 125–250 µm) was introduced into the quartz glass tube reactor (i.d. 8 mm), fixed with quartz wool and pre-treated with 2 vol% O2 in N2 at 450 °C for 30 min (500 ml/min, STP). After cooling to 130 °C in a N2 flow, the sample was saturated with 1000 ppm NH3 in N2 (500 ml/min, STP). Weakly bound NH3 was removed by flushing with N2 until the outlet NH3 fraction was below 2 ppm. For the final NH3-TPD, the sample was heated to 550 °C at a rate of 10 K/min in N2 (500 ml/min, STP). The temperature was measured by two K-type thermocouples located directly in front of and behind the catalyst bed, respectively. Desorbing NH3 was monitored by means of NDIR spectroscopy (X-Stream, Emerson). Assuming a molar NH3 to acid site stoichiometry of 1, the number of surface acid sites was estimated from the integrated NH3 traces.
Diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) was conducted using NH3 as a molecular probe. The investigations were performed on a Tensor 27 FTIR spectrometer (Bruker) equipped with Praying Mantis reflectance optics (Harrick Scientific) and a MCT detector. The sample compartment was continuously purged with N2 to avoid the diffusion of air. Each sample was introduced into the heatable stainless steel IR cell (Harrick Scientific) equipped with ZnSe windows. After pre-treatment with 2 vol% O2 in N2 at 450 °C for 30 min (200 ml/min, STP), the sample was cooled to 180 °C, 130 °C, 90 °C and 50 °C in a N2 flow, a background spectrum being collected at each temperature. Subsequently, the sample was exposed to 1000 ppm of NH3 (N2 balance) for 30 min at 50 °C. After flushing with N2 for another 30 min at 50 °C to remove physisorbed NH3, an IR spectrum was recorded. The temperature was then increased to 90 °C and the sample was purged again with N2 for 30 min before another spectrum was taken. The same procedure was adopted at 130 °C and 180 °C. The total gas flow was always kept at 200 ml/min (STP). Spectra were collected from 800 to 4000 cm−1 with a resolution of 1 cm−1, and 200 scans were accumulated per spectrum, resulting in an acquisition time of approx. 5 min.

2.3 Catalytic H2-deNOx Studies

Catalytic investigations were performed on a laboratory bench using a lean model exhaust gas. The granulated samples (200 mg, 125–250 μm) were introduced into the quartz glass tube reactor (i.d. 8 mm), fixed with quartz wool and heated in an N2 flow to 280 °C for 20 min. Subsequently, the synthetic exhaust, composed of 200 ppm NO, 2000 ppm H2, 5 vol% O2, 10 vol% H2O and N2 (balance), was dosed using independent mass flow controllers (Bronkhorst). The total flow was kept at 400 ml/min (STP), corresponding to a space velocity (GHSV) of 160,000 h−1. Once a steady state was reached at 280 °C, the temperature was linearly reduced to 60 °C at a rate of 1.7 K/min. Note that this test procedure provided a quasi-steady state, as proven by preliminary studies. The temperature was measured using two K-type thermocouples located directly in front of and behind the catalyst bed. The difference between inlet and outlet temperature was always below 8 K. All H2-deNOx data are referred to the averaged temperature. To investigate the impact of CO on the H2-deNOx performance of Pt/xMo/ZrO2, 200 ppm CO was added to the lean model exhaust. For reference purposes, SCR experiments with and without 2000 ppm H2 were also performed by adding 100 ppm NO and 100 ppm NH3 to the model exhaust. After pre-treatment at 280 °C, each temperature was kept constant for 45 min to achieve steady-state conditions excluding NH3 storage effects.
Gas-phase analysis was performed using a Multi-Gas 2030 FTIR spectrometer (MKS Instruments), detecting NO, NO2, N2O, NH3 and H2O, whereas O2 was monitored with an LSU 4.9 lambda sensor (Bosch). Inlet concentrations were checked employing the reactor bypass. To assess the catalysts, NOx conversion (X(NOx) = 1- y(NOx)out/y(NOx)in), maximum NOx conversion (Xmax) and a temperature of Xmax (T(Xmax)) were used. Additionally, the selectivity of N2O (S(N2O) = 2·y(N2O)out/(y(NOx)in – y(NOx)out)), NH3 (S(NH3) = y(NH3)out/(y(NOx)in – y(NOx)out)) and N2 (S(N2)) were calculated. Since N2 is not detectable using FTIR, S(N2) was derived from the mass balance of N including the measured species NO, NO2, NH3 and N2O, i.e. S(N2) = 1- (2·y(N2O)out + 2·y(NH3)out) /(y(NOx)in – y(NOx)out).

3 Results and Discussion

3.1 H2-deNOx Performance of the Pt/Mo/ZrO2 Catalysts

The most important H2-deNOx features of the Pt/xMo/ZrO2 catalysts are demonstrated in Fig. 1, indicating that the maximum NOx conversion and N2 selectivity at Xmax increase when the Mo load (w(Mo)) grows to 3 wt%, and decrease at higher Mo contents. As a result, Pt/3Mo/ZrO2 displays the highest H2-deNOx activity and N2 selectivity, including Xmax of approx. 90%, T(Xmax) of 130 °C and S(N2) of 78% at Xmax. For clarity, the H2-deNOx performance of Pt/3Mo/ZrO2 is shown in Fig. 2. This reveals a broad activity window with deNOx above 30% between 105 °C and 200 °C. Furthermore, Pt/3Mo/ZrO2 strongly suppresses the NH3 formation (maximum S(NH3) = 2%), an effect that is markedly pronounced for the 0.25Pt/ZrO2 reference (maximum S(NH3) = 50% at 160 °C). The catalysts with Mo proportions above 3 wt% do not produce NH3 at all, while a N2O selectivity of 22% still remains.
The performance of Pt/3Mo/ZrO2 is similar to that of the highly active WOx-modified Pt/ZrO2 catalysts, which also provide NOx conversions up to 90% between 100 °C and 180 °C, but are characterized by higher N2 selectivity (80–90%) [11, 18]. A N2 selectivity above 90% was reported for Pt/MgO-CeO, albeit using a markedly higher H2 proportion of 1 vol% [6]. Note that for the WOx-promoted catalysts, a rise in S(N2) with increasing H2 concentration was also demonstrated [9, 11, 18], while at the same time the selectivity of H2 to deNOx drastically decreases.
Additional H2-deNOx investigations performed with 200 ppm CO in the model exhaust led to a drastic decline in the catalytic activity of the Pt/xMo/ZrO2 catalysts, especially at low temperatures (Fig. S2). Since CO strongly covers active Pt sites at low temperatures [19], significant deNOx activity is only observed above 135 °C, when CO is almost completely oxidized and free Pt sites become accessible for the NOx reduction. Note that this experiment was performed with decreasing temperatures. Contrary, when heating up the sample the CO light-off might be slightly shifted to higher temperatures potentially compromising the NOx conversion. However, even above the CO light-off temperatures, the presence of CO causes a diminished NOx conversion due to the competing adsorption on the active Pt sites; for instance, the 0.25/3Mo/ZrO2 sample only achieves a Xmax of 50% appearing at 163 °C. CO and HC were reported as having a similar effect on the H2-deNOx reaction in the case of the Pt/W/ZrO2 catalysts [1, 18]. Thus, for engines operating with hydrocarbon-based fuels such as gasoline and diesel, an oxidation catalyst upstream of the H2-deNOx stage is mandatory to ensure that the H2-deNOx is efficient at low temperatures.

3.2 Structural Promotion of the Pt/Mo/ZrO2 Catalysts

The LRS investigations of the Pt/xMo/ZrO2 samples (Fig. 3) indicated that the ZrO2 carrier is exclusively present in the monoclinic phase as indicated by the peaks at 176–184 cm−1 (Bg), 310 cm−1 (Ag), 335 cm−1 (Bg), 382 cm−1 (Bg), 474 cm−1 (Ag), 536 cm−1 (Bg), 560 cm−1 (Ag), 618 cm−1 (Bg) and 635 cm−1 (Ag) [2022]. This interpretation is substantiated by PXRD, which shows the reflexes of monoclinic ZrO2 (PDF-04-004-4339) (Fig. S3). Moreover, the samples with low Mo loadings (≤ 3 wt% Mo) provide additional LRS peaks at 860 cm−1 (Mo-O-Mo vibrations) and 930 cm−1 (M=O vibrations), ascribed to highly dispersed Mo oxide species [23, 24]. These peaks are shifted from 860 to 889 cm−1 and from 930 to 960 cm−1 when the Mo proportion is raised to 7 wt% reasonably suggesting aggregated MoOx entities, which likely exist in the form of large clusters [23]. For the highest Mo content of 10 wt%, exceeding the estimated theoretical monolayer of MoO3 (equivalent to 8 wt% Mo) [25], the formation of crystalline MoO3 occurs. This is shown by the LRS peaks at 748 cm−1s(O-Mo-O)) and 947 cm−1as(O-Mo-O) [23, 24] and is also suggested by the PXRD reflex at 26.9° (Fig. S3), which is attributed to the (112) plane of β-MoO3 (PDF-04-007-2607).
Figure 4 (left) illustrates that the MoOx promoter steadily increases the BET surface area of the samples when the Mo proportion rises from 0 wt% (65 m2/g) to 4 wt% (101 m2/g). This effect is ascribed to the inhibition of the sintering of ZrO2 by MoOx, which is known to reduce the number and mobility of defects on the support surface [26]. However, when Mo loadings increase further, the BET surface area continuously decreases, probably due to the aggregation of the MoOx species, which have a lower specific surface area compared to ZrO2 [27].
CO-TPD studies show that the number of active Pt sites is significantly increased by promoting the bare Pt/ZrO2 (Fig. 4, right), going through a maximum of 7.1 μmol/g at a Mo load of 3 wt%. Table S4 indicates that the Pt dispersion obviously follows the same trend including maximum dispersion of 56% for the catalyst with 3 wt% Mo (14% for bare Pt/ZrO2), while the estimated Pt particle size is decreased to a minimum of 2 nm for the same sample. It is important to note that the clear decrease in the number of available Pt sites at Mo loadings above 3 wt% cannot be entirely attributed to the decline in the BET surface area (Fig. 4, left). Since the increase in the molybdenum content is connected with the aggregation of the Mo oxide species (Fig. 3), it may be supposed that already during the impregnation of the ZrO2 support Mo clusters form at the higher Mo loads. Then, electrostatic repulsion of the positively charged Mo cores of these clusters and Pt2+ cations may occur during subsequent treatment with the Pt solution. With this assumption it may be speculated that the Pt cations are more concentrated on the reduced surface area of the bare ZrO2 and at the interface to the Mo clusters, respectively, leading to agglomeration of the Pt particles upon the following activation and calcination.
Previous mechanistic studies of WOx-promoted Pt/ZrO2 catalysts using in-situ DRIFTS [11] and elementary kinetic modelling [28, 29] indicated that the H2-deNOx reaction mainly occurs on the Pt sites. With this molecular mechanistic understanding [28, 29], four neighbouring active Pt sites (*) are required for the N2 formation, including the adsorption of 2 NO molecules, the dissociation of the resulting NO adsorbates (NO*, Eq. (1)) and the recombination of the two N adsorbates (N*) (Eq. (2)). By contrast, the production of N2O only requires 3 adjacent active sites, since N2O originates from a NO and a N adsorbate (Eq. (3)). Consequently, the number of active sites is assumed to affect both the catalytic activity and the N2 selectivity. The formation of the side-product NH3 is not considered here in detail due to its minor yield when using the Mo-promoted Pt/ZrO2. For the same reason, the possible formation of N2O by the oxidation of NH3 tends to be negligible, as substantiated in Sect. 3.3. Basically, NH3 is formed by the step-by-step addition of a H atom to NHx adsorbates [30].
$$NO* \, + \, * \to N* \, + \,O*$$
(1)
$$2\,N* \, \to \,N_{2} \, + \,2*$$
(2)
$$N* \, + \,NO* \, \to \,N_{2} O\, + \,2*$$
(3)
Indeed, the Pt/xMo/ZrO2 catalysts show a clear increase in the maximum NOx conversion when the quantity of available Pt sites increases (Fig. 5, left), and the N2 selectivity at Xmax also tends to rise in line with the number of Pt sites (Fig. 5, right). The impact of the active sites on the activity as well selectivity is also reflected by the turnover frequency of NO (TOF), which expresses the NO conversion rate per available Pt site (TOF = ṅ(NOx)in ∙ X(NOx)/na(Pt)). The TOF exemplarily calculated for 130 °C provides similar values for bare Pt/ZrO2 and the Pt/xMo/ZrO2 catalysts (Table S4) substantiating that the number of the active Pt centers drives the H2-deNOx activity. In contrast to that, the N2 formation rate referred to the available Pt sites (ṅ(N2)/na(Pt)) grows significantly from 5.9∙103 s−1 (bare ZrO2) to 1.4∙104 s−1 and 1.5∙104 s−1 when the Mo load is increased to 3 and 4 wt%, but declines for the further increasing Mo contents (Table S4). This trend supports the importance of the active Pt centers in controlling the N2 selectivity of the H2-deNOx reaction. From the above shown relations it is inferred that highly dispersed MoOx species act as structural promoters, driving the H2-deNOx activity and N2 selectivity by stabilizing the BET surface area and providing higher numbers of accessible Pt sites and smaller Pt particles, respectively. However, aggregated Mo oxide moieties existing at Mo loads of 4 wt% and above clearly attenuate this promoting effect.

3.3 Acid Promotion of the Pt/Mo/ZrO2 Catalysts

It was shown in Sect. 3.1 that the MoOx promoter strongly affects the H2-deNOx activity as well as selectivities towards N2, N2O and NH3. One notable feature is the decreasing NH3 formation in the presence of the promoter, with complete suppression at a Mo content of 4 wt% and beyond accompanied by increased N2O selectivity (Fig. 1).
The decrease in the NH3 production may be related to the enhanced NH3 oxidation activity of the Mo-modified catalysts; a similar effect of the WOx promoter was observed for the related Pt/W/ZrO2 catalysts [31]. Indeed, the NH3 oxidation studies show that the conversion of NH3 is shifted to lower temperatures when bare Pt/ZrO2 is compared with Pt/4Mo/ZrO2 (Fig. S5). For instance, the temperature at which a NH3 conversion of 50% is achieved (T50) is shifted markedly from 240 °C for bare Pt/ZrO2 to 185 °C for Pt/4Mo/ZrO2. Moreover, in the additional presence of 2000 ppm H2 the NH3 oxidation activity of Pt/4Mo/ZrO2 is further enhanced, as illustrated by the shift in T50 to approx. 150 °C. Pt/ZrO2 shows a similar effect (Fig. S5). This H2-assisted NH3 oxidation may be due to the reaction of H2 with oxygen bound to Pt, resulting in a larger number of free Pt sites, which are available for the adsorption and subsequent conversion of NH3. Additionally, the light-off temperatures of the NH3 oxidation in the presence of H2 are in line with the temperature range in which NH3 is observed during the H2-deNOx reaction, i.e. no NH3 appears in H2-deNOx above the NH3 light-off temperature. Moreover, the conversion of NH3 in the presence of H2 is continuously shifted to lower temperatures when the Mo loading grows, including N2O selectivities between 40 and 50% for all the Pt/xMo/ZrO2 catalysts (Fig. S6).
Furthermore, the reduced NH3 formation during H2-deNOx on the Mo-promoted catalysts may also be explained by an additional deNOx mechanism, which involves NHx species formed on the Pt sites reacting with NOx [9, 3234]. Thus, NH3 was taken as a reducing agent to evaluate the potential of the catalysts for a SCR-related pathway of this kind. Interestingly, Fig. 6 (left) indicates superior activity (without H2) of the promoted Pt catalysts with Mo loads of 2, 4 and 7 wt% compared to Pt/ZrO2, while the highest deNOx is achieved for the sample with 4 wt% Mo. By contrast, the Pt-free 4Mo/ZrO2 sample reveals no performance at all (Fig. 6, left), substantiating the importance of Pt for the SCR reaction [30]. As expected from the literature [30], the SCR reaction on Pt catalysts also leads to the strong evolution of N2O, which clearly represents the major product for the samples with Mo loads of 2 and 4 wt%; for instance, at 170 °C, the N2O selectivity amounts to 71% and 83%, respectively (Fig. 6, right). The addition of 2000 ppm H2 to the SCR feed gas also significantly decreases the N2O selectivity of the Pt/xMo/ZrO2 catalysts (Fig. 6, right). As an example, for Pt/4Mo/ZrO2 it declines from 83 to 31% at 170 °C (X(NOx) = 90%). This performance is close to the H2-deNOx reaction on the same catalyst, showing a N2O selectivity of 25% and a NOx conversion of 80% at 150 °C.
These investigations clearly show that the Pt/xMo/ZrO2 catalysts are substantially active in the SCR reaction, whereas the role of the Pt component obviously lies in the activation of NH3, since the SCR reaction is totally inhibited in the absence of Pt. Therefore, it is deduced that the NHx intermediates (x = 1 – 3) originating from the H2-deNOx reaction [30] also participate in the NOx reduction. Additionally, it is possible that these NHx species may spill over to the acid sites of the Mo/ZrO2 support, so that the NO-NHx reaction occurs at the interface between the Pt particles and the substrate [32, 33]. Mechanistic and kinetic studies showed that below 200 °C, NO is the most abundant adsorbate on the Pt sites of a related Pt/W/ZrO2 catalyst under very similar reaction conditions [29]. Moreover, the basic NHx species are believed to preferentially coordinate to the acid sites on the support. To gain a deeper understanding of how NH3 interacts with Mo/ZrO2 and NHx may spill over from the Pt particles, DRIFTS studies were performed with NH3 as the molecular probe, along with NH3-TPD examinations.
DRIFT spectroscopy was conducted to check the NHx adsorbates formed during the adsorption of NH3. The Pt/xMo/ZrO2 samples were exposed to 1000 ppm NH3 at 50 °C followed by flushing and heating to 130 °C in a N2 flow, then the spectra were recorded. Figure 7 demonstrates the spectral range of the NHx deformation vibrations, revealing a weak DRIFTS band at approx. 1600 cm−1as) and an intense band at approx. 1185 cm−1s) for all the samples. Both bands are related to the NH3 molecularly coordinated to Lewis acid sites, whereas some contribution could be provided from NH3 adsorbed on Pt sites [3537]. Particularly, the intense σs(NH3) band located between 1180 cm-1 and 1270 cm−1 is significantly increased for the samples containing Mo, the highest intensity being detected for the Mo load of 4 wt% (Fig. 7, inset). This indicates that the highly dispersed Mo oxide entities existing in the samples with the low Mo loads contribute additional Lewis acid sites to the support, probably in the form of Mo6+. However, the σs(NH3) vibration markedly decreases above 4 wt% Mo, suggesting that the number of Lewis acid sites is lower at high Mo loads, likely associated with the stronger aggregation of the MoOx entities. Note that the shift in the σs(NH3) band from 1186 cm−1 (Pt/ZrO2) to 1235 cm−1 (Pt/7Mo/ZrO2) is in good accordance with the literature (ZrO2: 1160 cm−1, MoO3: 1270 cm−1), indicating a stronger Lewis acidity of the MoOx moieties [35]. This interpretation is supported by the fact that the σs(NH3) band of the Mo-promoted samples does not decline when the temperature is raised from 50 to 180 °C, while it is significantly reduced for bare Pt/ZrO2 (Fig. S7).
In contrast to bare Pt/ZrO2, the Mo-promoted Pt/ZrO2 samples show a weak DRIFTS band at 1660 cm−1s(NH4+) and a prominent signal at approx. 1445 cm−1as(NH4+)). Both bands are attributed to NH4+ species originating from the reaction of NH3 with Brønsted acid sites [38]. Moreover, the band areas, particularly that of the 1445 cm−1 feature, increase with the content of Mo (Fig. 7, inset), indicating a growing number of Brønsted acid sites. This may be primarily attributed to the increasing amount of aggregated MoOx species detected by LRS (Fig. 3); these are known to form Mo-OH-Zr groups that can act as proton donors [39, 40]. However, the intensity of the σas(NH4+) band substantially decreases when the temperature is raised from 50 to 180 °C, indicating that the thermal stability of the NH4+ species is relatively low, in line with the literature (Fig. S7) [39, 41].
NH3-TPD investigations were carried out to evaluate the NH3 uptake capacity of the xMo/ZrO2 supports (Fig. 8). The exposure to NH3 took place at 130 °C, corresponding to substantial H2-deNOx activity in the case of Pt/3Mo/ZrO2, which is the best catalyst of the samples tested. Figure 8 (inset) shows the highest NH3 uptake for the samples with Mo contents of 2 and 3 wt% implying that there is a significant increase in the total amount of acid sites, whereas bare ZrO2 and the samples with higher Mo fractions adsorb clearly smaller NH3 quantities. For comparison, a NH3-TPD study was also performed with the Pt/2Mo/ZrO2 catalyst showing no significant effect of the Pt component on the NH3-TPD profile (Fig. S8). Moreover, the samples with 2–4 wt% Mo also exhibit larger amounts of strongly bound NH3, desorbing above 300 °C only. Based on the DRIFTS studies, this strongly adsorbed ammonia is primarily coordinated to Lewis acid sites, whereas the NH3 desorption at lower temperatures (approx. 210 °C) is related to the decomposition of NH4+ entities originating from the reaction of NH3 with Brønsted acid sites [39, 41]. When the LRS (Fig. 3) and NH3-TPD results are combined, it can be inferred that the highly dispersed MoOx species that are particularly present in the samples with moderate Mo loads (2–4 wt% Mo) result in stronger Lewis acid sites, as indicated by the pronounced NH3 desorption at 300 °C and above.
The results of the DRIFTS and NH3-TPD studies are correlated with the activity of the Pt/xMo/ZrO2 catalysts in H2-deNOx and SCR to assess how NHx species contribute to an additional deNOx pathway. The relation of the σs(NH3) DRIFTS band area (Fig. 7) and the SCR activity of the Pt/xMo/ZrO2 catalysts at 170 °C (Fig. 6) indicates that the deNOx performance improves with a rise in the amount of NH3 bound to Lewis acid sites (Fig. 9). From this relation it is deduced that the NH3 species which are coordinated to strong Lewis acid sites, as primarily found on highly dispersed MoOx species, participate in the SCR reaction. In contrast, the Brønsted acid sites and resulting NH4+ species play a minor role: the sample containing 7 wt% Mo shows by far the largest σas(NH4+) band, but significantly lower SCR activity compared to the sample with a Mo load of 4 wt%. Considering the lack of SCR activity in the Pt-free 4Mo/ZrO2 sample, it may be further assumed that the reaction of NO adsorbed on Pt and NHx bound to Lewis acid sites takes place at the Pt sites or the interface of Pt and the Mo/ZrO2 substrate [30, 33].
Moreover, the area of the σs(NH3) DRIFTS bands correlates with the maximum N2 selectivity in the H2-deNOx investigations (Fig. 9). A similar correlation is obtained regarding the amount of ammonia desorbed from strong acid sites above 300 °C in the NH3-TPD experiments (Fig. S9). These relations suggest that the introduction of moderate Mo loads (2–4 wt%) to the Pt/ZrO2 catalyst creates strong Lewis acidity at highly dispersed MoOx moieties, thus enabling an additional deNOx pathway to form, providing high N2 selectivity that simultaneously extends to the H2-deNOx mechanism on the Pt sites stated above [28, 29]. On this SCR-related pathway, NHx or NH3 species formed on the Pt sites are expected to spill over to strong Lewis acid sites in the case of highly dispersed MoOx entities. The reaction between these activated NH3 or NHx species and the NO probably adsorbed on the Pt sites is assumed to take place at the interface between the Pt particles and the MoOx species, forming N2 and H2O as the major products. Therefore, the outstanding activity and N2 selectivity of the best Pt/3Mo/ZrO2 catalyst is attributed not just to the high Pt dispersion, but also to the presence of highly dispersed MoOx species, forming strong Lewis acid sites and thus enabling an SCR-type reaction of NO with NH3, featuring high N2 selectivity.

4 Conclusions

The present study shows that Mo is an effective promoter for Pt/ZrO2, enhancing the lean H2-deNOx activity and N2 selectivity at low temperatures. The best catalyst exhibits a Mo load of 3 wt% revealing high H2-deNOx efficiency between 80 °C and 150 °C with a maximum NOx conversion of 90% and a N2 selectivity of up to 78%. Thus, this Mo-promoted Pt/ZrO2 sample shows a performance similar to the best reported H2-deNOx catalysts, which are Pt/W/ZrO2 and Pt/MgO-CeO2 [6, 11].
Our studies have shown that highly dispersed Mo species have a particularly beneficial promoting effect on Pt/ZrO2 catalysts, which can be attributed to the enhancement of Pt dispersion, affecting both NOx conversion and selectivity towards N2. However, at Mo loadings above 4 wt%, the dispersed Mo species aggregate and there is a decrease in the number of Pt sites. As a result, there is a drastic decrease in the H2-deNOx activity and N2 selectivity. Moreover, our experiments provide strong evidence that the Mo promotion of Pt/ZrO2 enables an additional deNOx mechanism involving strong Lewis acid sites at the MoOx entities. It is assumed that NH3 or NHx species formed on the Pt sites during H2-deNOx spill over to the strong Lewis acid sites at highly dispersed MoOx moieties, followed by a reaction with NO at the interface between the Pt sites and MoOx. Thus, our study highlights the particular importance of acidity in enhancing the activity and selectivity of lean H2-deNOx catalysts. In particular, strong Lewis acid sites can trap active NHx species, creating an additional deNOx pathway at the interface of Pt and the promoter/support, whereas Brønsted acid sites are likely of minor importance for NHx activation. Thus, in our future research we aim to examine whether the acid promotion of Pt/ZrO2 is a universal means of enhancing the H2-deNOx reaction, particularly to produce selective N2 formation. For this purpose, other acid promoters such as tungsten, sulphur and phosphorus will be systematically investigated, in which context the adjustment of the respective promoter structures is the key to achieving high Lewis acidity.

Acknowledgements

The authors gratefully acknowledge the financial support from the Development Bank of Saxony (Saechsische Aufbaubank) as a subsidiary of the European Regional Development Fund (EFRE) under the RedNOx project (100367274).
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Anhänge

Supplementary Information

Below is the link to the electronic supplementary material.
Literatur
1.
Zurück zum Zitat Lox E (2008) Automotive exhaust treatment. In: Ertl G, Knözinger H, Schüth F et al (eds) Handbook of heterogeneous catalysis, 2nd edn. Wiley-VCH Verlag, Weinheim, pp 2286–2331 Lox E (2008) Automotive exhaust treatment. In: Ertl G, Knözinger H, Schüth F et al (eds) Handbook of heterogeneous catalysis, 2nd edn. Wiley-VCH Verlag, Weinheim, pp 2286–2331
17.
Zurück zum Zitat Bergeret G, Gallezot P (2008) Particle size and dispersion measurements. In: Ertl G, Knözinger H, Schüth F et al (eds) Handbook of heterogeneous catalysis, 2nd edn. Wiley-VCH Verlag, Weinheim, pp 738–765 Bergeret G, Gallezot P (2008) Particle size and dispersion measurements. In: Ertl G, Knözinger H, Schüth F et al (eds) Handbook of heterogeneous catalysis, 2nd edn. Wiley-VCH Verlag, Weinheim, pp 738–765
18.
Zurück zum Zitat Esser, E., Kureti, S., Heckemüller, L., Todt, A. et al (2022) Low-Temperature NOx Reduction by H2 in Diesel Engine Exhaust. SAE Technical Paper 2022–01–0538 Esser, E., Kureti, S., Heckemüller, L., Todt, A. et al (2022) Low-Temperature NOx Reduction by H2 in Diesel Engine Exhaust. SAE Technical Paper 2022–01–0538
36.
Zurück zum Zitat Davydov A (2003) Molecular spectroscopy of oxide catalyst surfaces. John Wiley & Sons Inc, Hoboken, NJCrossRef Davydov A (2003) Molecular spectroscopy of oxide catalyst surfaces. John Wiley & Sons Inc, Hoboken, NJCrossRef
37.
Zurück zum Zitat Knözinger H (2008) Infrared spectroscopy for the characterization. In: Ertl G, Knözinger H, Schüth F et al (eds) Handbook of heterogeneous catalysis, 2nd edn. Wiley-VCH Verlag, Weinheim, pp 1135–1163 Knözinger H (2008) Infrared spectroscopy for the characterization. In: Ertl G, Knözinger H, Schüth F et al (eds) Handbook of heterogeneous catalysis, 2nd edn. Wiley-VCH Verlag, Weinheim, pp 1135–1163
Metadaten
Titel
Low-Temperature NOx Reduction by H2 on Mo-Promoted Pt/ZrO2 Catalysts in Lean Exhaust Gases
verfasst von
D. Schröder
S. Kureti
Publikationsdatum
31.10.2022
Verlag
Springer US
Erschienen in
Topics in Catalysis / Ausgabe 13-14/2023
Print ISSN: 1022-5528
Elektronische ISSN: 1572-9028
DOI
https://doi.org/10.1007/s11244-022-01720-4

Weitere Artikel der Ausgabe 13-14/2023

Topics in Catalysis 13-14/2023 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.