Next Article in Journal
Recyclable TiO2–Fe3O4 Magnetic Composites for the Photocatalytic Degradation of Paracetamol: Comparative Effect of Pure Anatase and Mixed-Phase P25 TiO2
Previous Article in Journal
Tailoring Electrocatalytic Pathways: A Comparative Review of the Electrolyte’s Effects on Five Key Energy Conversion Reactions
Previous Article in Special Issue
Effects of SiO2, Al2O3 and TiO2 Catalyst Carriers on CO-SCR Denitration Performance of Bimetallic CuCe Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

H2-SCR over Low Loaded Platinum-Based Catalysts: Investigations in the Reaction Pathways

Institut de Chimie des Milieux et Matériaux de Poitiers (IC2MP), Université de Poitiers, CNRS, 4 rue Michel Brunet TSA 51106, 86073 Poitiers Cedex 9, France
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(9), 838; https://doi.org/10.3390/catal15090838
Submission received: 22 July 2025 / Revised: 22 August 2025 / Accepted: 26 August 2025 / Published: 1 September 2025
(This article belongs to the Special Issue Heterogeneous Catalysis in Air Pollution Control)

Abstract

The pathways and mechanistic aspects of H2-SCR over precious metal-based catalysts is still under debate. This study focusses on low loaded platinum-based catalysts (0.07–0.3%) in a large temperature range (50–500 °C), with special focus on (i) the role of NH3 as a possible intermediate species, (ii) the origin of the undesired N2O emission and (iii) the platinum sites involved in the H2-SCR deNOX reactions. Up to 60 °C, the N2O selectivity was close to 100%, with no influence of the presence of oxygen in the 50–100 °C temperature range. Ammonia formation was observed at relatively low temperatures (from 60 °C), but its reactivity was then limited. All these low temperature reactions were associated with the same platinum sites, probably a mix of edge and face sites. The maximum outlet NH3 was observed around 100 °C and the role of the NH3-SCR in the whole H2-SCR process appeared very limited. On the contrary, the ammonia oxidation by O2, which started near 120 °C, significantly contributed to the H2-SCR process and appeared responsible for the second N2O emission peak (150–500 °C). This reaction did not imply the same platinum sites and appears mainly dependant on the platinum particle size.

1. Introduction

To reduce the CO2 emissions responsible for the greenhouse gas (GHG) footprint of the automotive sector, three main carbon zero-emission technologies for vehicles are proposed: battery electric vehicles (BEVs), hydrogen fuel cell electric vehicles (FCEVs), and hydrogen internal combustion engines (H2-ICEs). Achieving the objective of the GHG emission mitigation necessitates that hydrogen production come from non-fossil fuels and renewable energy. The use of hydrogen as a carbon-free fuel in internal combustion engines is gaining significant interest due to its high energy density, its lower sensibility to H2 quality (compared to fuel cells) and the possibility of the reconversion of current robust and mature diesel and natural gas engine technologies on a possibly hybrid powertrain coupled system. As a result, H2-ICEs may serve as an attractive transitional solution toward the large-scale deployment of hydrogen-based propulsion systems [1,2].
However, some nitrogen oxides (NOX) are still emitted by H2-ICEs as harmful pollutants, with emissions possibly exceeding those of conventional Diesel engines when a stoichiometric air/H2 mixture is used [3]. Fortunately, NOX emissions decrease significantly as the air/H2 ratio increases, dropping from approximately 4000 ppm under stoichiometric mixture to around 100 ppm in lean conditions [4]. Moreover, operating H2-ICEs under lean conditions not only significantly limits NOX emissions—potentially achieving a 90% reduction compared to diesel engines—but also enhances thermal efficiency [3]. Nevertheless, achieving a zero-emission H2-ICE that can compete with clean fuel cells is still a challenge. One attractive option is to develop a robust, passive and cost-effective catalytic for the NOX selective catalytic reduction by H2 (H2-SCR) to convert the remaining NOX traces into N2.
According to the literature, the H2-SCR mechanism can be divided into two distinct routes, namely (i) the NO adsorption/dissociation pathway based on a Langmuir–Hinshelwood mechanism and (ii) the oxidation–reduction pathway [5]. Regardless of the considered mechanism, a typical volcano-shaped curve in NOX conversion is obtained. This behaviour is attributed to the intensified competition with the H2 combustion reaction as the temperature increases, particularly above 150 °C on precious metals.
The NO adsorption/dissociation mechanism is primarily associated with noble metal-supported catalysts which have been developed for lean-deNOX systems for Diesel engines after-treatment [6]. Pt and Pd-based catalysts exhibit the highest activity when hydrogen is used as the reductant. However, the operating temperature range for these catalysts is rather limited (typically between 100 and 300 °C) and their selectivity at low temperatures suffers from the generation of undesirable N2O emissions, especially in the presence of water [7]. The key elementary step for the NO adsorption/dissociation route involves the NO dissociative adsorption, leading to the formation of adsorbed O and N atoms. The N2 formation occurs through the recombination of two adsorbed nitrogen atoms (Nads) on the catalyst surface. The role of H2 is then to remove adsorbed oxygen atoms (Oads) from the active sites, thereby cleaning the surface. It is generally admitted that the N2O formation arises from the reaction between Nads and NO [5,8]. However, it has also been proposed that N2O formation can result from the reaction between NOads and NO, leading to N2O and Oads [9]. Indeed, the N2O emission via the formation and the decomposition of a (NO)2 dimer species at a low temperature is supported by the lower energetic barrier of this pathway compared to NO dissociation [10]. Moreover, it was also observed on Rh/SiO2 that the suppression of H2 in the reaction mixture led to N2O formation, while outlet N2 decreased [11]. It was concluded that NOad and Oad neighbouring species were involved in N2O formation. The increase in Oad leads to a decrease in Nad and also promotes hydrogen combustion [12]. These features illustrate competitions in neighbouring platinum active sites, which must be addressed to gain a deeper understanding of H2-SCR pathways at low temperatures.
The oxidation–reduction mechanism is based on adsorbed intermediate N-containing species including nitrosonium, nitrites, nitrates or ammonia-like (NHX) species [13]. Interestingly, this NO reduction mechanism appears similar to the NH3-SCR pathway, in which the generation of nitrosamine intermediates (NH2NO) plays a key role in the enhancement of the SCR performance [14]. Whatever the reaction mechanism considered, the reaction pathways remain a topic of ongoing debate and NHX type species are also reported to play a significant role in the H2-SCR process. NHX formation takes place on metallic sites via the reaction between adsorbed nitrogen species from the NO dissociative adsorption and active hydrogen. Subsequently, NHX compounds react with adsorbed NO species to yield N2 [15]. At low temperatures, the dissociation of NO remains the rate-determining step (rds) in the H2-SCR process, while the in situ formation of NHX species becomes a crucial step reaction at higher temperatures. In fact, N2 selectivity was proposed to depend on the H(ad)/NOX(ad) ratio [16]. As expected, since the NHX species is generated, the NOX conversion is favoured with catalysts that exhibit acidic behaviours, which promote the NH3-SCR activity. The acidic properties of the support have also been reported to influence the H2-SCR behaviour by impacting the metal oxidation state. Li et al. [17] showed that platinum-based catalysts supported on acidic supports promoted the formation of metallic platinum particles, thereby promoting the dissociative adsorption of NOX. However, the reaction selectivity was reported to be mainly dependent on the acidity of the support rather than the intrinsic properties of platinum. These observations were attributed to a bifunctional mechanism: at a low temperature (75° C), the formation NH4+ intermediate species occurs via the reaction of NO with H2 on the metallic platinum surface. These NHX species subsequently adsorb onto the acidic sites of the support, where they facilitate the selective reduction in NO, ultimately yielding N2 as the main reaction product. This NH3-SCR involvement is notably reported in the very recent review from Jabłońska et al. [18].
Finally, the pathways and mechanistic aspects of the H2-SCR over precious metal-based catalysts remain the subject of ongoing debate. Among the usual precious metals (Pt, Pd, Ir), Shao et al. recently confirmed that Pt exhibited the highest catalytic activity but also offered the widest operational temperature window [19]. Consequently, platinum was selected in this study to investigate the reaction pathways of the H2-SCR over low loaded platinum-based catalysts, with a special focus on (i) the role of NH3 as a possible intermediate species, (ii) the undesired N2O emission and (iii) the platinum sites involved in the deNOX reactions. To address these aspects, platinum loading was limited to 0.07%–0.3 wt% due to its high cost, and various silica-alumina supports with different silica contents were selected to study the influence of acidity on H2-SCR activity and selectivity, in comparison with the use of alumina alone. This study is mainly based on catalytic measurements and a statistical cuboctahedron platinum cluster model. To assess the H2-SCR reaction pathways depending on the temperature, various masses of catalyst were involved, and various associated reactions, such as the NO + H2 reaction (without oxygen), NH3-SCR, the NH3 oxidation by O2, were separately examined. Very recently, Dong et al. developed a similar approach on Pt/SSZ-13 catalysts [20], varying the platinum loading between 0.01 wt% and 0.5 wt%. They proposed that the H2-SCR reaction over this zeolite-supported catalyst involves hydrogen spillover and the generation of NHX+ species, and that NH3-SCR and NHX+ oxidation may occur in the whole reaction.

2. Results

2.1. Investigation in Silica-Alumina and Al2O3 Supports Properties for H2-SCR over 0.30 wt%Pt-Based Catalysts

In order to select the more appropriate support to investigate the H2-SCR reaction pathways, an alumina support and three silica-alumina supports were selected and impregnated with 0.30 wt% platinum. Catalytic tests were then first performed with 200 mg of various samples in the 100–500 °C temperature range. The catalytic test set-up and protocols are described in Section 3.3 and Appendix A.

2.1.1. Characterization

The textural properties of the alumina support and the three selected silica-alumina supports with silica weight contents of 5%, 20% and 40% (all provided by Sasol) are reported in Table A1 (Appendix B.1). Pure alumina exhibited the lowest specific surface area (157 m2 g−1), while increasing the silica content from 5% to 40% led to a progressive rise in surface area, from 296 m2 g−1 to 422 m2 g−1.
The acidic properties of the supports were evaluated by pyridine adsorption monitored by infrared spectroscopy, which allows the quantification of both Lewis acid sites (LAS, associated with PyL species) and Brønsted acid sites (BAS, associated with the formation of PyH+ species). The corresponding spectra obtained after pyridine adsorption and evacuation at 150 °C are presented Figure A1 (Appendix B.2).
Alumina shows a typical spectrum with characteristic bands associated with the Lewis acid sites (LAS) at 1454 cm−1, 1624 and 1618 cm−1. The doublet with peaks at 1622 and 1618 cm−1 reveals the presence of multiple types of LAS, differing by their acidity strength and Al coordination (AlIV and AlVI). Additionally, as expected, no Brønsted acidity was detected on Al2O3. The spectrum obtained with Siral-5 is closely similar to that of pure alumina, except that the components observed at 1624 and 1618 cm−1 now merged into a single band centred at 1622 cm−1, which suggests a homogenization of the Lewis acid sites in terms of acidity strength. Increasing SiO2 content progressively led to a decrease in the intensity of the LAS-related bands and to a shift in the band at 1454 cm−1 to higher frequencies, reaching 1457 cm−1 for Siral-40. This shift indicates a change in the nature of the surface Lewis acid sites, likely due to alterations in their electronic structure and chemical environment. Additionally, a weak ν19b band at around 1548 cm−1, attributed to pyridinium ions formed by a pyridine interaction with Brønsted acid sites, was evidenced in the spectra of Siral-20 and Siral-40. Note that the intensity of this ν19b band was very low for the Siral-5 sample.
The quantification of the amount of each LAS and BAS depending on the evacuation temperature is plotted in Figure 1 for each support, normalized by gram of catalyst and determined from ν19b area, using its molar coefficient. As expected, the increase in the evacuation temperature led to a decrease in the amount of absorbed species. Some pyridine adsorption was still observed on LAS at 450 °C, while Brønsted sites exhibited nearly no adsorption beyond 300 °C. Among the studied supports, Siral-5 exhibited the highest amount of acid sites, with a low proportion of Brønsted sites. In contrast, Siral-20 and Siral-40 exhibited significant Brønsted acidity, which persisted from pyridine evacuation up to 250 °C.
The platinum loading of the 0.30 wt%Pt/support samples (Table 1) shows that the expected amount of platinum was deposited whatever the considered support. Hydrogen chemisorption measurements and TEM were also performed to evaluate the platinum dispersion. Table 1 shows that the higher platinum dispersion was observed on alumina (19%). The platinum dispersion decreased with the silica content, from 13% on siral-5 to 10% on Siral-40. These results are consistent with those previously reported by Shibata et al. [21] who also observed a decrease in platinum dispersion with the increase in the silica content in SiO2-Al2O3 supports.

2.1.2. H2-SCR Behaviour

The 0.30 wt%Pt supported catalysts were evaluated in H2-SCR in the 100–500 °C temperature range (mcata: 200 mg). NOX conversion, N2 and N2O yields are reported in Figure A2 (Appendix B2). For all catalysts, the NOX conversion reached 100% at 100 °C. The NOX conversion then decreased with the temperature increase, except for 0.30Pt/Siral-5 which maintained a complete NOX conversion until 200 °C. The N2 and N2O formations in the 100–200 °C temperature range were dependent on the nature of the support. The major product was N2O for 0.30Pt/Al2O3, 0.30Pt/Siral-20 and 0.30Pt/Siral-40 (mean selectivity around 78%, 54% and 58%, respectively). On the contrary, the 0.30Pt/Siral-5 sample showed N2 as the major product (mean N2 selectivity: 58%). Figure 2 illustrates that NOX conversion and especially N2 selectivity were mainly dependant on acidic properties at 250 °C. Note that ammonia emission was not observed during these tests involving 200 mg of catalyst. These results are in accordance with Zhang et al. [22] which showed by diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) that LAS can contribute to NH3 adsorption, enhancing the N2 selectivity and deNOX performances by assisted NH3-SCR. However, this particular aspect about the ammonia formation is further examined in the next Section 2.2 by decreasing the involved weight of the catalyst.
Finally, these preliminary results showed that Siral-5 support offered the best compromise between NOX conversion and N2 selectivity. This support exhibited the higher acidity, mainly as Lewis sites with traces of Brønsted sites, a high specific surface area compared to pure alumina, and an intermediate platinum dispersion (12%, slightly higher than on the two other alumina-silica supports). The Siral-5 support was therefore selected for further investigations about the deNOX pathways involved in the H2-SCR reaction.

2.2. Investigation of the H2-SCR Behaviour Depending on the Involved Amount of Catalyst (0.30 wt%Pt/Siral-5)

To investigate the involvement of intermediate species and by-side reactions in the H2-SCR process, the mass of the catalyst was varied to increase the space velocity (same feed flow and composition). Starting from 200 mg, additional catalytic tests were performed decreasing the amount of the catalyst to 100 mg, 30 mg and 5 mg. Moreover, the minimum catalytic test temperature was decreased to almost 60 °C to insure NOX conversion below 100% in the low temperature range. The resulting NOX conversion profiles and the calculated yields in N2, N2O and NH3 are reported in Figure 3 depending on the temperature.
Considering the initial catalyst mass of 200 mg (Figure 3a), a high NOX conversion was recorded at low temperature, reaching 86% at 30 °C. However, the main product was N2O, with a yield of 71%. The NOX conversion remained higher than 95% across the 50–250 °C temperature range, with maximum N2 yield (66%) obtained around 250 °C. At higher temperatures, NOX conversion decreased due to the competition with hydrogen oxidation by O2 (see thereafter Section 2.3.4). No outlet ammonia was detected whatever the considered temperature in this experiment (200 mg of catalyst).
As the catalyst mass decreased (Figure 3a–d), the decrease in the number of available active sites resulted in a narrower operating temperature range. The NOX conversion was maintained above 90% in the 40–240 °C temperature ranges for 100 mg and 200 mg of involved catalyst, and this activity window decreased to 85–240 °C and 110–210 °C for 30 mg and 5 mg of catalyst, respectively.
Interestingly, decreasing the catalyst mass significantly influenced the distribution of the products. Particularly, Figure 3c,d shows the detection of NH3 centred around 100 °C, with an increasing amount recorded as the involved mass of the catalyst decreased. This observation indicates that NH3 is mainly formed at the beginning of the catalyst bed before further reaction. Note that a significant ammonia concentration was recorded at the lowest tested temperature when using 5 mg of catalyst (20 ppm at 60 °C). This raises important questions regarding the role of NH3 as an intermediate or by-product, which warrants further investigations. Additionally, decreasing the amount of the catalyst also revealed that N2O emissions split into two distinct broad peaks, one below and one above 100 °C. Surprisingly, the maximum N2 yield remained around 65% regardless of the mass of catalyst ranging from 200 mg to 5 mg. However, the temperature at which this yield was obtained was shifted to lower values, from 250 °C to 120 °C, as the catalyst mass was decreased.
Finally, these catalytic results clearly illustrate the participation of various reactions depending on the temperature. Further investigations are required to (i) determine at which temperature NH3 is produced, according to Equation (1), especially in comparison with the NOX conversion on-set temperature; (ii) elucidate the subsequent reactions of NH3 with other compounds present in the reaction mixture since NH3 may react with the remaining NOX (Equations (2) and (3)) and/or O2 (Equations (4) and (5)) to form N2 and/or N2O. Additionally, the H2 reactivity also deserves particular attention. Note that the possible reaction between H2 and N2O was also checked using 200 mg of catalyst: in a rich mixture (900 ppm N2O + 1% H2), the N2O conversion started below 100 °C and reached 100% at 250 °C (100% N2 selectivity), but the conversion became negligible with the addition of 2% O2 in the mixture (lean mixture). These results show no N2O reactivity in oxidative media.
NO + 5/2 H2 →NH3+ H2O
4 NH3 + 4 NO + O2 →4 N2 + 6 H2O
4 NH3 + 4 NO + 3 O2 → 4 N2O + 6 H2O
2 NH3 + 3/2 O2→ N2 + 3 H2O
2 NH3 + 2 O2 → N2O + 3 H2O
To address these questions, an individual and rigorous investigation of the various reactions which are supposed to intervene is necessary. This specific study under various reaction mixtures was performed with 30 mg of 0.30 wt%Pt/Siral-5 sample.

2.3. Investigation on H2-SCR Reaction Pathways: H2-SCR, (NO + H2) Reaction, NH3-SCR, NH3 Oxidation Reactions

2.3.1. H2-SCR vs. (NO + H2) Reaction Behaviour

To investigate ammonia formation in the H2-SCR process, a catalytic test was performed without oxygen in the reaction mixture (NO + H2 reaction, Table 2). Particular attention was paid to determine the initial ammonia emission temperature (Equation (1)) in relation to the NOX conversion onset temperature.
Figure 4a,b compare NOX conversions and yields in N2, N2O and NH3 depending on the reaction mixture, i.e., NO + H2 + O2 and NO + H2. Both figures show that the NOX conversion started at the same temperature with or without O2 in the feed stream, and both NOX conversion curves remained superposed until reaching full NOX conversion at around 100 °C. Similarly, the N2O yield curves were also superposed until reaching their maximum (56% at 80 °C), indicating that O2 is not involved in the “low temperature” behaviour. Furthermore, the NH3 emission also began at the same temperature of 70 °C for both reaction mixtures. Given the excess of reductant in the NO + H2 mixture, NOX conversion remained complete within the 115–500 °C temperature range. Since NOX conversion reached 100%, NH3 was the main product with a maximum yield of 80% (at 375 °C), while the N2 yield was between 7% and 13% in the 200–500 °C temperature range. Conversely, in the presence of O2 (H2-SCR, Figure 4a), the obtained ammonia can then react along the catalytic bed as illustrated in Figure 3, either with NOX or O2 (Equations (2)–(5)). Both reactions can influence the distribution of reaction products. These strongly temperature-dependent reactions are evaluated in the two following sections.

2.3.2. H2-SCR vs. NH3-SCR Reaction

Since NH3 was formed under the H2-SCR condition, the NH3-SCR reaction (Figure 4c) may occur along the catalytic bed. However, considering the measurements obtained in H2-SCR with 5 or 30 mg of catalyst (Figure 3c,d), the simultaneous presence of NOX and NH3 was restricted to a narrow temperature range, from approximately 70 °C to 100–150 °C (the oxidation by O2, evaluated in the next section, is then more likely).
According to the results reported in Figure 4c, NOx reduction by NH3 significantly started at approximately 100 °C, which also corresponds to the drop in NH3 emission under the H2-SCR conditions (Figure 4a). Consequently, ammonia was not involved in low temperature NOX conversion (below 100 °C) during the H2-SCR process. The conversion of NOX by NH3 (NH3-SCR, Figure 4c) reached a maximum at around 200 °C, a temperature at which the NH3 conversion became complete. Thereafter, the conversion gradually declined, becoming negligible at approximately 300 °C. At higher temperatures, the amount of detected NOX exceeded the inlet NOX, which can be attributed to the ammonia oxidation by O2 into NOX (orange line). However, the major product of the NH3-SCR reaction was clearly N2O, while the N2 yield remained around 11–18% in the 200–500 °C temperature range.

2.3.3. H2-SCR vs. NH3 Oxidation (By O2)

As previously mentioned, ammonia consumption during the H2-SCR reaction can also be associated with its oxidation by oxygen. As shown in Figure 4d, NH3 oxidation by oxygen (NH3 + O2, Table 2) started slowly at around 100 °C, and then increased significantly from 150 °C, reaching 100% at 190 °C. The ammonia conversion remained complete up to 500 °C. This reaction led mainly to N2, with a maximum yield of 80% recorded when the ammonia conversion reached 100%.
The N2O emission started to increase from 150 °C, reaching a maximum close to 30% at around 240 °C. The obtained N2O yield profile is fairly similar to that observed during the H2-SCR reaction at similar temperatures. This similarity suggests that NH3-oxidation by O2 would largely contribute to the second wave of N2O formation observed at high temperatures during H2-SCR.
Outlet NOX species were recorded from 200 °C during the NH3 oxidation test. This temperature also corresponds to the decrease in NOX conversion observed during the H2-SCR reaction. This suggests that the NH3 oxidation by O2 may contribute to the decrease in the H2-SCR efficiency at high temperatures, by the potential re-oxidation of NH3 into NO. However, another plausible explanation is the insufficient availability of H2 at these high temperatures.
In summary, results in Figure 4d show that the oxidation of NH3 by O2 appears to play a major role in the H2-SCR behaviours, especially at temperatures above 150 °C. This contribution could likely explain the high N2 selectivity observed in the H2-SCR reaction and the formation of the second N2O phase at high temperatures. In contrast, the NH3-SCR appears to be favoured only in a narrow temperature range, between 100 °C and 150 °C.

2.3.4. H2-Reactivity

The H2 conversion curve during the H2-SCR test on 0.30 wt%Pt/Siral-5 is reported in Figure 5. Considering the various reactions involving H2 (Equations (1) and (6–8)) and the outlet gas composition, an estimation of the hydrogen consumption in the H2-SCR was calculated and reported in Figure 5a. As expected, due to the large amount of H2 compared to NOX, only a small percentage (<5%) of the inlet hydrogen was used to reduce NOX, with the predominant reaction being the H2 oxidation by O2 regardless of the temperature.
H2 + 1/2 O2→H2O
2 H2 + 2 NO → N2 + 2 H2O
H2 + 2 NO → N2O + H2O
5/2 H2 + NO → NH3 + H2O
It is important to note that the H2 inlet concentration was 1% compared to 400 ppm NO. At 50 °C, the H2 conversion was very limited, while the NOX conversion had already reached 10% (a theoretical H2 conversion of 0.2% is required to convert 10% of the inlet NO into the N2O, Equation (8)). The H2 conversion curve (black line) showed a more significant increase from 80 °C, which also corresponded to significant ammonia outlet detection. The highest ratio of hydrogen utilization toward NOX reduction was observed at 100 °C, when there was sufficient H2 activation, but limited O2 activation. Maximum ammonia emission was also observed at this temperature. The initial increase in N2 yield followed the increase in H2 conversion. For temperatures above 250 °C, the NOX conversion decreased, indicating that the available reductant for NOX reduction became insufficient, likely due to competition with the H2 oxidation by O2.
The H2 + O2 reaction (without NO) was subsequently examined on 0.30%Pt/Siral-5. The H2 conversion curves with or without NO in the feed stream (H2-SCR vs. H2 + O2 reactions) are compared in Figure 5b. The H2 conversions at 50 °C were 4% and 82% in H2-SCR (black line) and H2 + O2 reactions (red line), respectively (shift of T50 from 28 °C to 98 °C). These results clearly indicate that the presence of NO in the feed stream inhibited the hydrogen conversion by oxygen at a low temperature. These specific statements are discussed in the next section which is a focused on the low-temperature behaviours considering the N2, NO, O2 interactions over platinum-based catalysts.

2.4. Discussion

Many studies have examined the interactions of O2, H2 and NO with a platinum surface in the past. Although reported heat of adsorption or sticking coefficient values vary significantly depending on the considered platinum-based sample, the general trends clearly indicate that, at room temperature, oxygen shows the strongest interaction with platinum. O2 adsorption is dissociative on platinum, with a reported heat of adsorption ranging from 300 to 370 kJ mol−1 [23] to 210 kJ mol−1 [24]. For supported platinum species, adsorption enthalpies are typically in the 210–280 kJ mol−1 range [24,25,26], mainly depending on the platinum dispersion. Notably, Frank et al. [8] reported a significantly lower adsorption enthalpy of 97 kJ mol−1. They explained this value by the fact that the heat of adsorption is lower than the activation energy required for O2 desorption because the dissociative oxygen adsorption is an activated process. Only in the case of non-activated adsorption, the heat of adsorption can be directly compared to the desorption energy. Interestingly, it has also been reported that the oxygen coverage of Pt(111) surface is limited to 0.25 [27], in accordance with the fact that the saturation coverage of oxygen has been reported to be lower than that of hydrogen [28]. Indeed, a complete hydrogen surface coverage of platinum via H2 dissociative adsorption can be obtained [29]. Again, reported heats of adsorption for H2 on platinum vary significantly across the studies, but they appear always lower than those of O2, typically in the 56–87 kJ mol−1 range [8,27,29,30]. No significant trend with platinum crystallite size was observed over a range of 1.3 to 22.3 nm Pt deposited on SiO2 [30].
Concerning the NO adsorption on platinum, both molecular and dissociative adsorption have been proposed at room temperature. Fiorin et al. [26] reported that NO initially adsorbs molecularly on Pt(111), and dissociatively on Pt(211) (stepped surfaces) and Pt(411) (unit cell with two terraces), with a higher amount of dissociated NO observed on Pt(411) compared to Pt(211), This aligns with previous studies showing that step and/or terraces sites enhance the reactivity of NO [31,32]. The calculated adsorption energies at the fcc threefold sites are around 170–180 kJ mol−1 [23,33,34] and smaller at the bridge site (140 kJ mol−1) and at the atop site (62 kJ mol−1) [33]. Consequently, as mentioned by Dhainaut et al., the competition between H2 and NO for adsorption on Pt(111) was largely in favour of NO, especially considering that the probability for finding two adjacent sites for the H2 dissociative adsorption would be very low [35].
Finally, considering O2, H2 and NO, it can be concluded that O2 exhibits the strongest interactions with platinum, but with a limited maximum coverage (0.25 on Pt(111)) which allows the co-adsorption of other molecules. This limited coverage by oxygen is consistent with the fact that NO conversion at low temperatures appeared unaffected by the presence of oxygen (Figure 4a,b). Considering the competition between H2 and NO, the NO adsorption is favoured, which is consistent with the fact that the presence of NO inhibited the H2 + O2 reaction at a low temperature (Figure 5b). Additionally, in the presence of O2 and NO, the hydrogen adsorption should be limited, which could suggest an Eley–Rideal type mechanism for the NOX reduction. High nitric oxide and oxygen coverage is also consistent with the high N2O selectivity observed at a low temperature (close to 100% below 70 °C), probably via the formation of the (NO)2 dimer species as mentioned in the Introduction section [10].
The possible participation of NOX species stored on the Siral-5 support was also examined through NOX-TPD experiments. The NOX desorption profiles recorded with the Siral-5 support and the 0.15 wt%Pt/Siral-5 catalyst are reported Figure 6. Despite its acidic behaviour (Section 2.1.1, Appendix B.2), the Siral-5 support is able to store significant amounts of NOX species, around 950 µmol g−1. The NOX desorption from Siral-5 support without platinum started near 200 °C and showed a maximum at 425 °C (blue curve). The addition of platinum mainly led to a shift in the starting desorption temperature down to 180 °C, and the whole profile was shifted to approximately 30 °C to lower temperatures compared to the support alone (red curve). Consequently, NOX adsorption clearly occurred on the silica-alumina support, but these species appear stable at 100 °C and below. Therefore, we can assume that they do not intervene to a large extent in the deNOX activity at a low temperature.

2.4.1. Identification of the Main Active Sites Responsible for the Low Temperature H2-SCR Activity and the NH3-SCR Activity

Varying the catalyst, and particularly the platinum dispersion, can reveal a possible sensitivity of a reaction pathway to the size and shape of metal Pt particles. With this aim, several catalysts were prepared. For the four considered alumina and silica-alumina supports, a new series of catalysts with 0.15% platinum loading were prepared. Additionally, the platinum loading was also decreased down to 0.07% for the selected Siral-5 support. Their mean platinum particle sizes were determined and reported in Table 1. However, probably due to the rather severe hydrothermal treatment, the obtained platinum size range remained relatively limited. To enlarge it, additional catalysts were considered: a 0.30%Pt/Siral-5 sample treated only under pure N2 at 700 °C (without hydrotreatment at 700 °C), and a commercial reduced 1%Pt/Al2O3. This allowed for a broader distribution of platinum particle sizes (Table 1).
All these 11 catalysts were evaluated in H2-SCR and NH3 oxidation tests. To avoid full NOX conversion at a low temperature, catalysts with platinum loading until 0.30% were evaluated involving 30 mg, while 5 mg were used for the 1% Pt/Al2O3 catalyst. To study the initial low temperature H2-SCR behaviours, the selected temperature was 60 °C for which the NOX conversions were between 31% and 73%, while the ammonia production was evaluated at 100 °C. For the NH3 oxidation tests, a temperature of 180 °C was selected. The corresponding catalytic data are reported in Table A2 (Appendix C).
The turnover frequency (TOF), which corresponds to the deNOX activity (mol s−1 g−1) divided by the amount of surface platinum sites (mol g−1) is usually suitable to evaluate which kind of platinum sites are involved in a reaction when plotted depending on the metallic particle size. With this aim, Figure A4a,b report the evolution of the surface site concentration and distribution depending on the particle size of a perfect fcc truncated octahedron (from [36]). The turnover frequency (TOF) for the deNOX activity at 60 °C was plotted in Figure A5 depending on the platinum particle size. This figure shows a general increase in the TOF with the particle size increase, but the points are rather dispersed, and the trend does not match well with any evolution of specific platinum sites.
Then, since the considered catalysts exhibited large ranges of specific surface areas and platinum loadings, the influence of the platinum density was examined by plotting the NOX conversion rate at 60 °C (mol s−1 gcata−1) depending on the amount of platinum per square metre (Figure 7a). This figure shows that a high platinum density favours the NOX conversion. It can be then assumed that interactions between platinum particles are beneficial, probably via the mobility of the surface species. It should be noted that no relevant relationship between the NOX conversion rate at 60 °C and the acidity of the catalysts was obtained, which may be attributed to (i) the relatively narrow range in support acidities or (ii) a metal-intensive catalytic reaction.
To take into account the platinum density previously evidenced, the NOX conversion rate was expressed in mol s−1 m−2 and normalized by the amount of platinum (molPt g−1). Results are plotted in Figure 8a, depending on the platinum particle size. Practically, the total amount of surface platinum globally varies with the square of the particle size. Consequently, the total surface site normalized by mass of platinum decreases with the particle size, as also illustrated in Figure A4a considering the addition of all the sites. However, the profile reported in Figure 8a does not respect this evolution and the NOX conversion rate appears depending mainly on specific sites, for which the proportion varies with the particle size. Figure A4 shows the surface site evolution calculated for corners, edges, (100) faces and (111) faces. Clearly, corners cannot be responsible for the NOX conversion at 60 °C. In contrast, the trend observed in Figure 8a appears consistent with those reported for edges, (100) faces and (111) faces. However, Figure A4 predicts the maximum surface concentration for edges and (111) faces for platinum particle sizes around 2 nm, which does not perfectively match the evolutions reported in Figure 8a. Then, this suggest that NOX conversion probably depends on a mix of these face and edge sites.
As expected, since the major product at 60 °C was N2O, Figure 8b shows that same platinum sites are involved in N2O formation. Note that the observed trend takes into account all the various supports, with no specific behaviour depending on the oxide support. This observation indicates that the reaction at low temperature is governed predominantly by the platinum phase, rather than the support properties.
Since ammonia appeared to play a key role as the intermediate species in the H2-SCR pathway, its formation was also tentatively considered. Due to the very narrow temperature range between the ammonia formation and its subsequent reactivity, a temperature of 100 °C was selected for evaluation, as it generally corresponds to the maximum of ammonia emission. The NH3 formation rates (molNH3 s−1 m−2) normalized by the amount of platinum (molPt g−1) are plotted in Figure A6 depending on the platinum particle size. Once again, it appears that the ammonia emission at 100 °C followed the same trend as the NO reduction at 60 °C (mainly into N2O), indicating that the same platinum sites were involved. However, points in Figure A6 are more scattered, likely due to the partial reactivity of ammonia even at 100 °C.
Ammonia was previously shown to react mainly with oxygen (Section 2.3.3). Therefore, the ammonia oxidation was also investigated depending on the platinum particle size. With this aim, the NH3 + O2 reaction was examined at 180 °C, a temperature for which the NH3 conversions were between 9% and 56% (5 mg of catalyst were used for 1% Pt loaded catalyst while 30 mg were used for the sample with lower platinum loadings). As previously performed for the H2-SCR tests, the NH3 conversion rates at 180 °C (mol s−1 m−2) normalized by the amount of platinum (molPt g−1) were plotted depending on the platinum particle size (Figure A7). Unlike the previous obtained trends, points appear scattered, and a general increase can be observed, especially when the 0.3Pt/Al2O3 catalyst (dPt = 7 nm) is not considered. This result suggests that ammonia oxidation does not occur on the same platinum sites compared to the previous considered H2-SCR reaction. It is assumed that catalytic behaviour is multifactorial (specific surface area, acid properties, platinum dispersion…). Particularly, Figure 2 (H2-SCR over 0.30Pt/support catalysts at 250 °C, mcata = 200 mg) showed that NOX conversion in H2-SCR and especially N2 selectivity appeared mainly dependant on the acidic properties of the support. More precisely, Figure A2 shows that the more acidic support maintained a high level of NOX conversion over a wider temperature window. In this relatively high temperature range, the NH3 + O2 reaction was demonstrated to be the main one responsible for the observed behaviour (Figure 4). However, despite extensive calculations involving the NH3 + O2 conversion rates at 180 °C and the acidity behaviour, no clear correlation has been obtained. The only general trend obtained for the ammonia oxidation behaviour at 180 °C is an increase in the NH3 conversion rate normalized by the amount of platinum (molNH3 s−1 molPt−1) with the increase in the platinum particle size (Figure A8). This indicates that large platinum particle sizes appear more active in ammonia oxidation by oxygen. Considering the platinum site distribution (Figure A4), this trend does not match the contribution of corners and edges, but rather suggests the involvement of faces. In addition, a large particle size could also be associated with a higher metallic state. Note that the catalysts exhibiting the larger platinum particle sizes are not associated with any specific support (Table 1).

2.4.2. Overview of the H2-SCR Reaction Pathways Depending on Temperature

In order to have an overview of the varied reactions involved in the H2-SCR process, the main conversion/yields curves are compared in Figure 9 depending on the reactional mixture.
Oxygen from the gas phase had no influence on the low temperature NO conversion (50–100 °C). This observation strongly suggests that NO oxidation was not involved in the catalytic activity, in contrast to the oxidation–reduction mechanism. In accordance with this assumption, no NO2 was detected in the gas phase. In addition, O2 in the gas phase did not affect the N2O yields (Figure 5) and the selectivity in N2O remained close to 100% up to 60 °C. A few ppm of ammonia were detected from this temperature. The reactivity of ammonia was then very low. Indeed, the NH3-SCR activity was observed from around 70 °C and became significant only from 100 °C. Consequently, below 100 °C, the dissociative mechanism (including the formation of NH3) is more probable than the oxidation–reduction one.
Since the NOX conversion in H2-SCR reached 100% from 100 °C, the availability of NOX for the NH3-SCR reaction was limited to a very narrow temperature window (mostly below 100 °C, while the activity in NH3-SCR was still limited). As a result, the maximum outlet NH3 was observed around 100 °C and the role of the NH3-SCR in the whole H2-SCR process appeared very limited. On the contrary, the ammonia oxidation by O2, which started near 120 °C, significantly contributed to the H2-SCR process. This reaction appeared responsible for the second N2O emission peak (150–500 °C, Figure 5), and also facilitated the N2 formation as the main product. The optimal N2 yield was observed when H2 conversion reached 100%, around 150 °C, which can be attributed to the optimal ammonia formation followed by its oxidation by O2. These results are in accordance with the very recent work of Dong et al. on Pt/SSZ-13 catalysts [20]. They propose that the NH3-SCR route, resulting mainly in N2O formation and limited N2 selectivity occurs mainly below 150 °C, while the oxidation of the generated NHX+ species at a higher temperature favours N2 formation. These various reaction pathways are finally summarized in Figure 10.

3. Materials and Methods

3.1. Catalysts

Alumina and three silica-alumina supports with silica weight loadings of 5%, 20% and 40% (denoted as Siral-5, Siral-20 and Siral-40, respectively), all provided by Sasol, were selected for this study. These supports were in the form of fine powders with grain sizes below 100 μm. Prior to use, they were firstly hydrothermally treated at 700 °C for 4 h (synthetic air flow + 10% H2O; heating rate: 5 °C min−1). Platinum was then added using the incipient wetness impregnation method. The desired amount of metal precursor (Pt(NH3)2(NO2)2; Sigma Aldrich, Darmstadt, Germany) was dissolved in ultrapure water and then added dropwise to the support, which had been previously moistened with ultrapure water, to reach Pt loadings of 0.07, 0.15 and 0.30 wt%. The mixture was thereafter stirred magnetically for 30 min to ensure a homogeneous distribution of the precursor. The solvent was then evaporated at 80 °C using a sand bath. The resulting powder was dried overnight at 120 °C in an oven and then treated at 700 °C under a nitrogen atmosphere (60 mL min−1) for 4 h to decompose the nitrate precursors and stabilize the metal on the support surface [37]. Finally, the catalyst was hydrotreated at 600 °C for 4 h under synthetic air with 10% H2O.
The resulting powder was then subjected to pelleting, grinding and sieving to achieve a controlled grain size between 400 and 200 μm. Additionally, a commercial 1%Pt/Al2O3 catalyst (Alfa Aesar, reduced, 300 m2 g−1) was also evaluated after a treatment under 10% H2 at 600 °C for 1 h.

3.2. Characterizations

Elemental analyses were performed using an Agilent 5100 apparatus (Santa Clara, CA, USA). Samples were previously mineralized in a HNO3-HCl mixture, assisted by a microwave heating.
Textural behaviours were assessed by nitrogen physisorption performed with a TRISTAR 2 set-up (Micromeritics, Norcross, Georgia, USA). The sample was first degassed under vacuum at 250 °C for 2 h to remove previously adsorbed gases or contaminants. Subsequently, the sample was then exposed to increasing doses of nitrogen gas at −196 °C. Once adsorption was completed, the pressure was gradually reduced. The resulting adsorption and desorption isotherms were analyzed to determine the material’s specific surface area and pore size distribution.
Acidic properties were evaluated by means of pyridine adsorption monitored by FTIR spectroscopy. IR spectra were recorded in a Nexus Nicolet spectrometer (Themo Fisher, Waltham, MA, USA) equipped with a DTGS detector (Deuterium TriGlyceride Sulphur) and KBr beam splitter with a resolution of 4 cm−1 and 64 scans. The spectra presented were normalized to a disc of 10 mg cm−2. After sample activation at 450 °C under vacuum, pyridine was adsorbed at 50 °C (P = 200 Pa) and desorption was performed up to 450 °C. From pyridine adsorption, the Lewis and Brønsted acid sites (BAS and LAS, respectively) concentration are determined by the evolution of the ν19b band area (1456 or 1546 cm−1, respectively) using its molar coefficients previously reported [38].
The platinum particle mean sizes were estimated from transmission electron microscopy measurements. Micrographs were recorded on a JEOL 2100 instrument (Peabody, MA, USA) operated at 200 kV with a LaB6 source. For grid preparation, the sample was dispersed in ethanol under an ultrasonic field, and a drop of the upper suspension was deposited on the grid. The mean particle size was determined by considering generally between 300 and 600 particles for each catalyst. Hydrogen chemisorption measurements were also performed using a Micromeritics Autochem 2920II instrument (Norcross, GA, USA) to evaluate the accessible metallic surface. A catalyst sample weighing between 200 and 500 mg was placed in a U-shaped quartz reactor between two quartz wool beds. Downstream of the reactor, a metallic water trap containing magnesium perchlorate was installed to prevent water from reaching the detector. The sample was in situ pretreated for 1 h at 500 °C (5 °C min−1) under a hydrogen flow (30 mL min−1) to ensure a metallic state. Following this, the sample was purged under argon at the same temperature and flow rate (30 mL min−1) for 30 min and subsequently cooled to 30 °C. The measurements were then performed at 30 °C by pulsed injections (0.0551 mL) of 10% H2/Ar until saturation (20 pulses). After a purge under argon at 30 °C, a second series of pulses was performed to subtract the possible physisorbed hydrogen contribution. Hydrogen consumption was monitored using a thermal conductivity detector (TCD). Calculations for dispersion and particle size were based on the model developed by Le Valent et al. [39]. Due to the low platinum loading, the reported results correspond to a mean value of at least three measurements.

3.3. Catalytic Tests

The same set-up was used to evaluate various reactions such as H2-SCR, ammonia formation under NO + H2 (without O2), NH3-SCR, NH3 oxidation by O2, and NO oxidation. The corresponding mixtures are depicted in Table 2. For all experimental conditions, the total flow rate was 150 mL min−1 regulated by electronic mass-flow controllers.
Various controlled amounts of catalyst, ranging from 5 to 200 mg, were used for the catalytic tests. All tests were performed maintaining a constant catalytic bed volume, corresponding to that obtained with 200 mg of catalyst. For catalyst masses below 200 mg, silicon carbide (SiC) with the same grain size as the catalyst was added to maintain the bed volume. The catalyst was then placed on a bed of quartz wool inside the reactor.
Before each test, the sample underwent a pretreatment until 500 °C for 30 min (heating rate: 5 °C min−1) using the same reactional mixture as for the subsequent catalytic test. Catalytic behaviours were then evaluated by decreasing the temperature from 500 °C to 50 °C, with stepwise measurements at constant temperatures (even if results are later discussed in order of increasing temperature). The continuous monitoring of the feed gas and effluent stream compositions was performed using an online MKS 2030 Multigas infrared analyser for NO, NO2, N2O and NH3. Additionally, an OmniStar mass spectrometer (Pfeiffer Vacuum, Asslar, Germany) was used to record the H2 signal. N2 formation was calculated with the assumption that only NO, NO2, N2O and NH3 were formed as N-compounds. All equations employed for calculating conversions, yields and selectivity are displayed in Appendix A.

4. Conclusions

The pathways and mechanistic aspects of the H2-SCR over low loaded platinum-based catalysts (0.07–0.3%) was investigated over a large temperature range (50–500 °C), with a special focus on the role of NH3 as a possible intermediate species, the origin of the undesired N2O emission and the platinum active sites involved in the deNOX reactions. It was found that oxygen from the gas phase had no influence on the low temperature NOX conversion (50–80 °C), with the same N2O yields close to 100% up to 60 °C. Ammonia formation was observed at relatively low temperatures (starting from 60 °C) but its reactivity was then limited. All these low temperature reactions were associated with the same platinum sites, i.e., mainly the platinum faces and/or edges. The maximum outlet NH3 was observed around 100 °C. In opposition to conclusions frequently reported in the literature, the role of the NH3-SCR in the whole H2-SCR process appeared limited to a very narrow temperature range (more or less below 120 °C). On the contrary, the H2-SCR process at a high temperature appeared governed by the ammonia oxidation by O2, which started near 120 °C and contributed to the second N2O emission peak (150–500 °C). This reaction does not involve similar active sites to the initial H2-SCR reaction route and appeared mainly dependant on the platinum particle size.

Author Contributions

Conceptualization, F.C. and X.C.; Data curation, F.C. and X.C.; Formal analysis, A.B.A., F.C. and X.C.; Funding acquisition, F.C. and X.C.; Investigation, A.B.A. and F.C.; Methodology, F.C. and X.C.; Project administration, F.C. and X.C.; Resources, F.C. and X.C.; Supervision, F.C. and X.C.; Validation, A.B.A., F.C. and X.C.; Writing—original draft, A.B.A. and X.C.; Writing—review and editing, A.B.A., F.C. and X.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the French National Agency for Research (ANR, H-to Clean Project, ref. ANR-21-CE05-0012) and the Regional Council of Nouvelle Aquitaine (DeNOx-Decarb project, ref. AAPR2022-2021-17214210). The authors also acknowledge financial support from the European Union (ERDF) and Région Nouvelle Aquitaine.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Acknowledgments

Authors thank A. Le Valent from IC2MP for the discussions about the involved platinum sites in the considered reactions and data to perform Figure A4.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Catalytic Test: Calculations

Based on the data recorded by the multigas FTIR analyser, various calculations were performed to obtain conversions (C), yields (Y) and selectivity (S). Corresponding equations are reported in Equations (A1)–(A8) where “in” and “out” represent the inlet and outlet concentration, respectively. Since NO may be partially oxidized into NO2 (even in the gas phase), the NO reduction efficiency was calculated using NOX measurements.

Appendix A.1. H2-SCR Test

C   N O x   ( % ) = N O x i n N O x o u t N O x i n × 100
Y   N 2 O   % = 2 × N 2 O o u t N O x i n × 100
Y   N H 3   % = N H 3 o u t N O x i n × 100
Y   N 2   % =   C o n v .     N O x Y     N 2 O Y     N H 3
C   H 2   ( % ) = H 2 i n H 2 o u t H 2 i n × 100
S   N 2 O   % = 2 × N 2 O o u t N O x i n N O x o u t × 100
S   N H 3   % = N H 3 o u t N O x i n N O x o u t × 100
S   N 2 % = 1 2 × N 2 O o u t + N H 3 o u t N O x i n N O x o u t × 100
During the H2-SCR test, reactions with H2 can lead to various products as described by Equations (A9)–(A12):
NO   reduction   into   N 2 :   H 2 + N O ½ N 2 + H 2 O
NO   reduction   into   N 2 O :   H 2 + 2 N O N 2 O + H 2 O
NO   reduction   into   NH 3 :   H 2 + 2 5 N O 2 5 N H 3 + 2 5 H 2 O
H 2   combustion :   H 2 + 1 2 O 2 H 2 O
The use of H2 is described by Equations (A13)–(A16)
c o n v e r t e d   H 2   l e a d i n g   t o   N 2   ( % ) = N O x i n ( 2 N 2 O o u t + N H 3 o u t + N O x o u t ) H 2 i n × 100
converted H 2   l e a d i n g   t o   N 2   O ( % ) = N 2 O o u t H 2 i n × 100
c o n v e r t e d   H 2 leading   t o   N H 3   ( % ) = 5 N H 3 o u t 2 H 2 i n × 100
b u r n e d   H 2   b y   O 2   ( % ) = C H 2 ( c o n v e r t e d   H 2   l e a d i n g   t o   N 2 + c o n v e r t e d   H 2   l e a d i n g   t o   N 2 O + c o n v e r t e d   H 2 leading   t o   N H 3 )

Appendix A.2. NH3-SCR Test

C   N O x   ( % ) = N O x i n N O x o u t N O x i n × 100
C   N H 3   ( % ) = N H 3 i n N H 3 o u t N H 3 i n × 100
Y   N 2 O   % = 2 × N 2 O o u t N O x i n + N H 3 i n × 100
Y   N O x   % = N O x o u t N O x i n + N H 3 i n × 100
Y   N 2   % = ( 1 N O x o u t + N H 3 o u t + 2 × N 2 O o u t s N O x i n + N H 3 i n ) × 100

Appendix A.3. NH3-Oxidation Test

C   N H 3 = N H 3 i n N H 3 o u t N H 3 i n × 100
Y   N 2 O   % = 2 × N 2 O o u t N H 3 i n × 100
Y   N O x   % = N O x o u t N H 3 i n × 100
Y   N 2   % =   C   N H 3 Y   N 2 O Y   N O x

Appendix B. Preliminary Investigations for Support Selection

To enhance the N2 selectivity in H2-SCR on noble metal-based catalysts, one strategy involves the modification of the support (typically alumina) with metal oxides such as SiO2 or TiO2 to enhance the acidity. For instance, it was observed that the Pt/SiO2-Al2O3 catalyst (Si/Al = 5.3) exhibited a better NOX reduction rate and N2 selectivity compared to Pt/SiO2 and Pt/Al2O3 [21]. These results demonstrate that the combination of silica and alumina enhances catalytic properties. Therefore, in addition to pure alumina, three silica-alumina supports provided by Sasol were selected, with silica loading of 5%, 20% and 40%. Before use, all supports were hydrothermally treated at 700 °C for 4 h (synthetic air flow + 10% H2O; heating rate: 5 °C min−1)

Appendix B.1. Textural Behaviour

The textural properties of the treated supports after hydrotreatment are reported in Table A1. Alumina showed the lowest specific surface area (157 m2/g) and it increased with the increase in the SiO2 loading, reaching 422 m2/g for Siral-40. Conversely, alumina exhibited the largest average pore size (126 Å). It decreased until 77 Å with 20% SiO2 content, and the pore size then slightly increased to 86 Å for Siral-40. The total pore volume followed a similar trend: it increased with the SiO2 content up to Siral-20 (0.96 cm3 g−1) and then slightly decreased for Siral-40 (0.91 cm g−1), while alumina exhibited the smallest total pore volume (0.50 cm3 g−1).
Table A1. Textural properties of the selected supports after hydrothermal treatment at 700 °C, 4 h.
Table A1. Textural properties of the selected supports after hydrothermal treatment at 700 °C, 4 h.
SupportSiO2 Loading (%)SBET
(m2·g−1)
Average Pore Size
(Å)
Pore Volume
(cm3·g−1)
Alumina-1571260.50
Siral-55296830.62
Siral-2020356770.96
Siral-4040422860.91

Appendix B.2. Acid Behaviours Estimated by Pyridine Adsorption

The acidity of the supports was evaluated by pyridine adsorption monitored infrared spectroscopy which allows the quantification of both Lewis acid sites (LAS, associated with PyL species) and Brønsted acid sites (BAS, associated with the formation of PyH+ species). The spectra obtained after pyridine adsorption and evacuation at 150 °C are presented in Figure A1.
Figure A1. FTIR spectra recorded after pyridine adsorption and evacuation at 150 °C (spectra were normalized by mass of catalyst).
Figure A1. FTIR spectra recorded after pyridine adsorption and evacuation at 150 °C (spectra were normalized by mass of catalyst).
Catalysts 15 00838 g0a1
Alumina shows a typical spectrum with characteristic bands for Lewis acid sites at 1454 cm−1, 1624 and 1618 cm−1. The doublet with peaks at 1622 and 1618 cm−1 suggests the presence of several types of Lewis acid sites differing in acidity strength, likely due to variations in the coordination and proximity of aluminum and oxygen atoms on the surface. Additionally, no Brønsted acid sites were detected. The spectrum of Siral-5 was similar to that of pure alumina. However, the two components observed at 1624 and 1618 cm−1 on alumina disappear to form only one band centred at 1622 cm−1, which suggests a homogenization of the Lewis acid sites in terms of acidity strength.
As the SiO2 content increased, the intensity of the bands associated with the Lewis acid sites decreased, and a shift in the band at 1454 cm−1 to higher frequencies was also observed, reaching 1457 cm−1 for Siral-40. This indicates a change in the nature of the surface Lewis acid sites, likely due to alterations in the electronic structure and chemical environment of the acid sites.
Regarding Brønsted acid sites, no characteristic bands were evidenced for alumina, indicating the absence of sufficiently strong Brønsted acid sites to protonate pyridine. Siral-5 showed a very low intensity signal around 1546 cm−1 attributed to pyridinium ions, formed through the pyridine interaction with the Brønsted acid sites. On the contrary, Siral-20 and Siral-40 spectra clearly exhibited this characteristic band.

Appendix B.3. Support Selection: Catalytic Activity of 0.30%Pt/Support

A series of catalysts with 0.30%Pt supported on the four considered support was prepared by wet impregnation. Note that X-ray diffraction (XRD) analyses were also performed using an EMPYREAN PANalytical diffractometer (Almelo, The Netherlands) equipped with a copper source. The diffractograms were recorded for 2θ angles between 10° and 80°, with a step size of 0.02° (one step every 2 s). As expected, due to the low platinum loading, no significant signal attributable to platinum was recorded whatever the considered catalyst.
The catalysts were then evaluated in H2-SCR from 500 °C to 100 °C (mcata: 200 mg; reaction mixture: 400 ppm NO, 1% H2, 2% O2; total flow rate: 150 mL min−1). NOX conversions, N2 yields and N2O yields are reported in Figure A2.
Figure A2. () NOX conversion, () N2 yield and () N2O yield on (a) 0.30%Pt/Al2O3; (b): 0.30%Pt/Siral-5; (c): 0.30%Pt/Siral-20 and (d): 0.30%Pt/Siral-40. mcata: 200 mg; reaction mixture: 400 ppm NO, 1% H2, 2% O2; total flow rate: 150 mL.min−1.
Figure A2. () NOX conversion, () N2 yield and () N2O yield on (a) 0.30%Pt/Al2O3; (b): 0.30%Pt/Siral-5; (c): 0.30%Pt/Siral-20 and (d): 0.30%Pt/Siral-40. mcata: 200 mg; reaction mixture: 400 ppm NO, 1% H2, 2% O2; total flow rate: 150 mL.min−1.
Catalysts 15 00838 g0a2

Appendix C. Platinum Characterization and Catalytic Behaviour of the Selected Platinum-Based Samples

With the aim to highlight the main active sites responsible for the low temperature H2-SCR activity and the NH3-SCR activity, all the samples described in Table 2 were characterized in terms of platinum mean particle size by H2 chemisorption and/or TEM analysis. Examples of TEM analysis for catalysts supported on Siral-5 are depicted in Figure A3.
Figure A3. TEM images and particle size distribution for Pt/Siral-5 samples. (a): 0.07 wt%Pt/Siral-5; (b): 0.15 wt%Pt/Siral-5; (c): 0.30 wt%Pt/Siral-5; (d): 0.30 wt%Pt/Siral-5 without hydrotreatment at 600 °C.
Figure A3. TEM images and particle size distribution for Pt/Siral-5 samples. (a): 0.07 wt%Pt/Siral-5; (b): 0.15 wt%Pt/Siral-5; (c): 0.30 wt%Pt/Siral-5; (d): 0.30 wt%Pt/Siral-5 without hydrotreatment at 600 °C.
Catalysts 15 00838 g0a3aCatalysts 15 00838 g0a3b
All considered samples were also evaluated in H2-SCR and NH3 oxidation tests. To avoid full NOX conversion at a low temperature, catalysts with platinum loading up to 0.30% were evaluated involving 30 mg, while 5 mg was used for 1%Pt/Al2O3 catalysts. The data used for figures presented in Section 2.4.1 and Figure A5, Figure A6, Figure A7 and Figure A8 are reported in Table A2.
Table A2. Platinum characterization and catalytic behaviours of the various selected catalysts.
Table A2. Platinum characterization and catalytic behaviours of the various selected catalysts.
CatalystSBET
(m2 g−1)
Pt
Loading *
(wt%)
Mean Pt Particle Size
(nm) **
Pt
Dispersion
(%)
H2 SCR Test ***:
NOx Conv. at 60 °C (%)
H2 SCR Test ***:
N2Ooulet at 60 °C (ppm)
H2 SCR Test ***:
NH3 oulet at 100 °C (ppm)
NH3 Oxidation Test ***:
NH3 Conv. at 180 °C (%)
0.30wt%Pt/Al2O31440.397.0/-199514014056
0.30wt%Pt/Siral-52180.3410.4/10.613639514941
0.30wt%Pt/Siral-202450.3011.8/-11731298099
0.30wt%Pt/Siral-402320.3213.4/-107611314155
0.15wt%Pt/Al2O31490.13-/4.527578513613
0.15wt%Pt/Siral-52250.1611.4/11.21231486040
0.15wt%Pt/Siral-202390.15-/10.91242646938
0.15wt%Pt/Siral-402300.1314.3/13.59334911633
0.07wt%Pt/Siral-52690.078.2/8.41533527613
0.30wt%Pt/Siral-5 w/o hydrotreatment2240.231.9/-5637567118
1wt%Pt/Al2O3 (commercial) **1420.90-/2.64446761499
* From ICP analysis. ** mean value from H2 chimisorption and TEM analysis when both pieces of data were available. *** Catalytic tests were performed with 30 mg of catalyst, except for the 1%Pt/Al2O3 sample (5 mg).
To evaluate the platinum sites involved in the deNOX reactions, the evolution of the surface site concentration and distribution is depicted in Figure A4a,b, respectively, depending on the particle size of a perfect fcc truncated octahedron (from Ref. [36])
Figure A4. Evolution of the surface site concentration (a) and distribution (b) depending on the platinum particle size of a perfect fcc truncated octahedron: (): corner; (): edge; (): face 100; (): face 111.
Figure A4. Evolution of the surface site concentration (a) and distribution (b) depending on the platinum particle size of a perfect fcc truncated octahedron: (): corner; (): edge; (): face 100; (): face 111.
Catalysts 15 00838 g0a4
From data reported in Table A2, the turnover frequency (TOF) for the deNOX activity at 60 °C, which corresponds to the amount of converted NOX (mol s−1 g−1) normalized by the amount of surface platinum sites (mol g−1), was plotted in Figure A5 depending on the platinum particle size.
Figure A5. Evolution of the turnover frequency (TOF) at 60 °C for the NOx conversion depending on the platinum particle size.
Figure A5. Evolution of the turnover frequency (TOF) at 60 °C for the NOx conversion depending on the platinum particle size.
Catalysts 15 00838 g0a5
Additionally, Figure A6 illustrates the NH3 formation behaviour depending on the platinum particle size, while Figure A7 and Figure A8 deal with the NH3 conversion rate at 180 °C.
Figure A6. NH3 formation rate at 100 °C in H2-SCR (molNH3 s−1 m−2) normalized by the amount of platinum (molPt g−1) and plotted depending on the platinum particle size.
Figure A6. NH3 formation rate at 100 °C in H2-SCR (molNH3 s−1 m−2) normalized by the amount of platinum (molPt g−1) and plotted depending on the platinum particle size.
Catalysts 15 00838 g0a6
Figure A7. NH3 conversion rate at 180 °C in the NH3 oxidation test (molNH3 s−1 m−2) normalized by the amount of platinum (molPt g−1) and plotted depending on the platinum particle size.
Figure A7. NH3 conversion rate at 180 °C in the NH3 oxidation test (molNH3 s−1 m−2) normalized by the amount of platinum (molPt g−1) and plotted depending on the platinum particle size.
Catalysts 15 00838 g0a7
Figure A8. NH3 conversion rate at 180 °C in the NH3 oxidation test normalized by the amount of platinum (molNH3 s−1 molPt g−1) and plotted depending on the platinum particle size.
Figure A8. NH3 conversion rate at 180 °C in the NH3 oxidation test normalized by the amount of platinum (molNH3 s−1 molPt g−1) and plotted depending on the platinum particle size.
Catalysts 15 00838 g0a8

References

  1. Turner, W.G. Future technological directions for hydrogen internal combustion engines in transport applications. Appl. Energy Combust. Sci. 2025, 21, 100302. [Google Scholar] [CrossRef]
  2. Sari, R.; Shah, A.; Kumar, P.; Cleary, D.; Rairikar, S.; Balkrishna, S.; Thipse, S.S. Hydrogen Internal Combustion Engine Strategies for Heavy-Duty Transportation: Engine and System Level Perspective. SAE Int. J. Adv. Curr. Prac. Mobil. 2024, 6, 2937–2953. [Google Scholar] [CrossRef]
  3. Safari, H.; Jazayeri, S.A.; Ebrahimi, R. Potentials of NOX emission reduction methods in SI hydrogen engines: Simulation study. Int. J. Hydrogen Energy 2009, 34, 1015–1025. [Google Scholar] [CrossRef]
  4. Rakopoulos, C.D.; Kosmadakis, G.M.; Demuynck, J.; De Paepe, M.; Verhelst, S. A combined experimental and numerical study of thermal processes, performance and nitric oxide emissions in a hydrogen-fuelled spark-ignition engine. Int. J. Hydrogen Energy 2011, 36, 5163–5180. [Google Scholar] [CrossRef]
  5. Liu, Z.; Li, J.; Woo, S.I. Recent advances in the selective catalytic reduction of NOx by hydrogen in the presence of oxygen. Energy Environ. Sci. 2012, 5, 8799–8814. [Google Scholar] [CrossRef]
  6. Savva, P.G.; Costa, C.N. Hydrogen Lean-DeNOx as an Alternative to the Ammonia and Hydrocarbon Selective Catalytic Reduction (SCR). Catal. Rev. 2011, 53, 91–151. [Google Scholar] [CrossRef]
  7. Burch, R.; Coleman, M.D. An investigation of the NO/H2/O2 reaction on noble-metal catalysts at low temperatures under lean-burn conditions. Appl. Catal. B 1999, 23, 115–121. [Google Scholar] [CrossRef]
  8. Frank, B.; Emig, G.; Renken, A. Kinetics and mechanism of the reduction of nitric oxides by H2 under lean-burn conditions on a Pt–Mo–Co/α-Al2O3 catalyst. Appl. Catal. B 1998, 19, 45–57. [Google Scholar] [CrossRef]
  9. Pirug, G.; Bonzel, H.P. A low-pressure study of the reduction of NO by H2 on polycrystalline platinum. J. Catal. 1977, 50, 64–76. [Google Scholar] [CrossRef]
  10. Burch, R. The Investigation of Mechanisms in Environmental Catalysis Using Time-Resolved Methods. Top. Catal. 2003, 24, 97–102. [Google Scholar] [CrossRef]
  11. Papp, H.; Sabde, D.P. An investigation on the mechanism of NO decomposition over Rh/SiO2 catalysts in presence of pulse injected H2. Appl. Catal. B 2005, 60, 65–71. [Google Scholar] [CrossRef]
  12. Zhou, S.; Varughese, B.; Eichhorn, B.; Jackson, G.; McIlwrath, K. Pt-Cu Core-Shell and Alloy Nanoparticles for Heterogeneous NOx Reduction: Anomalous Stability and Reactivity of a Core-Shell Nanostructure. Angew. Chem. Int. Ed. 2005, 117, 4615–4619. [Google Scholar] [CrossRef]
  13. Burch, R.; Shestov, A.A.; Sullivan, J.A. A Steady-State Isotopic Transient Kinetic Analysis of the NO/O2/H2 Reaction over Pt/SiO2 Catalysts. J. Catal. 1999, 188, 69–82. [Google Scholar] [CrossRef]
  14. Li, J.; Chang, H.; Ma, L.; Hao, J.; Yang, R.T. Low-temperature selective catalytic reduction of NOx with NH3 over metal oxide and zeolite catalysts—A review. Catal. Today 2011, 175, 147–156. [Google Scholar] [CrossRef]
  15. Qi, G.; Yang, R.T.; Rinaldi, F.C. Selective catalytic reduction of nitric oxide with hydrogen over Pd-based catalysts. J. Catal. 2006, 237, 381–392. [Google Scholar] [CrossRef]
  16. Zhang, X.; Wang, X.; Zhao, X.; Xu, Y.; Gao, H.; Zhang, F. An investigation on N2O formation route over Pt/HY in H2-SCR. Chem. Eng. J. 2014, 252, 288–297. [Google Scholar] [CrossRef]
  17. Li, X.; Zhang, X.; Xu, Y.; Liu, Y.; Wang, X. Influence of support properties on H2 selective catalytic reduction activities and N2 selectivities of Pt catalysts. Chin. J. Catal. 2015, 36, 197–203. [Google Scholar] [CrossRef]
  18. Jabłońska, M.; Osorio Hernández, A. Selective Catalytic Reduction of Nitrogen Oxides with Hydrogen (H2-SCR-DeNOx) over Platinum-Based Catalysts. ChemCatChem 2024, 16, e202400977. [Google Scholar] [CrossRef]
  19. Shao, J.; Ho, P.H.; Creaser, D.; Olsson, L. Novel catalysts expanding the temperature range for NO selective catalytic reduction by H2. Appl. Catal. O 2024, 188, 206947. [Google Scholar] [CrossRef]
  20. Dong, Y.; Zhang, T.; Zhang, Y.; Du, J.; Sun, Y.; Liu, Z.; Ding, W.; Shan, Y.; Shan, W.; He, H. Revealing the reaction mechanism of H2 selective catalytic reduction of NOx in the presence of O2 over Pt/SSZ-13 zeolites. Appl. Catal. B 2025, 371, 125217. [Google Scholar] [CrossRef]
  21. Shibata, J.; Hashimoto, M.; Shimizu, K.; Yoshida, H.; Hattori, T.; Satsuma, A. Factors Controlling Activity and Selectivity for SCR of NO by Hydrogen over Supported Platinum Catalysts. J. Phys. Chem. B 2004, 108, 18327–18335. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Xu, S.; Li, J.; He, E.; Liu, Z. Unraveling the Promotional Effect of Co on the Pd/TiO2 Catalyst for H2-SCR of NOx in the Presence of Oxygen. J. Phys. Chem. C 2023, 127, 7248–7256. [Google Scholar] [CrossRef]
  23. Fiorin, V.; Borthwick, D.; King, D.A. Microcalorimetry of O2 and NO on flat and stepped platinum surfaces. Surf. Sci. 2009, 603, 1360–1364. [Google Scholar] [CrossRef]
  24. Sen, B.; Vannice, M.A. Enthalpy changes during O2 adsorption and H2 titration of adsorbed oxygen on Platinum. J. Catal. 1991, 129, 31–37. [Google Scholar] [CrossRef]
  25. Basset, J.M.; Theolier, A.; Primet, M.; Prettre, M. Proceedings of the 5th International Congress on Catalysis, Palm Beach, FL, USA, 21–25 August 1972; Hightower, J.W., Ed.; North-Holland: Amsterdam, The Netherlands, 1973; Volume 2, p. 915. [Google Scholar]
  26. Briot, P.; Auroux, A.; Jones, D.; Primet, M. Effect of particle size on the reactivity of oxygen-adsorbed platinum supported on alumina. Appl. Catal. 1990, 59, 141–152. [Google Scholar] [CrossRef]
  27. Norton, P.R.; King, D.A. The Chemical Physics of Solid Surfaces and Heterogeneous Catalysis; King, D.A., Woodruff, D.P., Eds.; Elsevier: Amsterdam, The Netherlands, 1982; Volume 4, p. 27. [Google Scholar]
  28. O’Rear, D.J.; Loffler, D.G.; Boudart, M. Stoichiometry of the titration by dihydrogen of oxygen adsorbed on platinum. J. Catal. 1990, 121, 131–140. [Google Scholar] [CrossRef]
  29. Christmann, K. Interaction of hydrogen with solid surfaces. Surf. Sci. Rep. 1998, 9, 1–163. [Google Scholar] [CrossRef]
  30. Sen, B.; Vannice, M.A. The influence of platinum crystallite size on H2 and CO heats of adsorption and CO hydrogenation. J. Catal. 1991, 130, 9–20. [Google Scholar] [CrossRef]
  31. Campbell, C.T.; Ertl, G.; Segner, J. A molecular beam study on the interaction of NO with a Pt(111) surface. Surf. Sci. 1982, 115, 309–322. [Google Scholar] [CrossRef]
  32. Rienks, E.D.L.; Bakker, J.W.; Baraldi, A.; Carabineiro, S.A.C.; Lizzit, S.; Weststrate, C.J.; Nieuwenhuys, B.E. Interaction of nitric oxide with Pt(1 0 0). A fast X-ray photoelectron spectroscopy study. Surf. Sci. 2002, 516, 109–117. [Google Scholar] [CrossRef]
  33. Ge, Q.; King, D.A. Energetics, geometry and spin density of NO chemisorbed on Pt{111}. Chem. Phys. Lett. 1998, 285, 15–20. [Google Scholar] [CrossRef]
  34. Yang, H.; Fdez Sanz, J.; Wang, Y.; Whitten, J.L. Adsorption Energetics of NO and CO on Pt(111). J. Clust. Sci. 1995, 10, 581–590. [Google Scholar] [CrossRef]
  35. Dhainaut, F.; Pietrzyk, S.; Granger, P. Kinetics of the NO/H2 reaction on Pt/LaCoO3: A combined theoretical and experimental study. J. Catal. 2008, 258, 296–305. [Google Scholar] [CrossRef]
  36. Le Valant, A.; Bouchet, S.; Van Assche, A.; Especel, C.; Epron, F. Description of supported metal structure sensitivity by a geometric approach. J. Catal. 2021, 397, 64–74. [Google Scholar] [CrossRef]
  37. Corbos, E.C.; Courtois, X.; Can, F.; Marécot, P.; Duprez, D. NOx storage properties of Pt/Ba/Al model catalysts prepared by different methods: Beneficial effects of a N2 pre-treatment before hydrothermal aging. Appl. Catal. B 2008, 84, 514–523. [Google Scholar] [CrossRef]
  38. Khabtou, S.; Chevreau, T.; Lavalley, J.C. Quantitative infrared study of the distinct acidic hydroxyl groups contained in modified Y zeolites. Microporous Mater. 1994, 3, 133–148. [Google Scholar] [CrossRef]
  39. Le Valant, A.; Comminges, C.; Can, F.; Thomas, K.; Houalla, M.; Epron, F. Platinum Supported Catalysts: Predictive CO and H2 Chemisorption by a Statistical Cuboctahedron Cluster Model. J. Phys. Chem. C 2016, 120, 26374–26385. [Google Scholar] [CrossRef]
Figure 1. Amount of Lewis acid sites (a) and Brønsted acid sites (b) from pyridine adsorption depending on the temperature of evacuation and the considered support. (): 0.30 wt%Pt/Al2O3; (): 0.30 wt%Pt/Siral-5; (): 0.30 wt%Pt/Siral-20; (): 0.30 wt%Pt/Siral-40.
Figure 1. Amount of Lewis acid sites (a) and Brønsted acid sites (b) from pyridine adsorption depending on the temperature of evacuation and the considered support. (): 0.30 wt%Pt/Al2O3; (): 0.30 wt%Pt/Siral-5; (): 0.30 wt%Pt/Siral-20; (): 0.30 wt%Pt/Siral-40.
Catalysts 15 00838 g001
Figure 2. Relationship between the total acid sites (µmol g−1) @250 °C (yellow) and the catalytic behaviour in H2-SCR @250 °C (mcata = 200 mg): (a) NOX conversion (blue); (b) N2 selectivity (blue).
Figure 2. Relationship between the total acid sites (µmol g−1) @250 °C (yellow) and the catalytic behaviour in H2-SCR @250 °C (mcata = 200 mg): (a) NOX conversion (blue); (b) N2 selectivity (blue).
Catalysts 15 00838 g002
Figure 3. 0.30 wt%Pt/Siral-5 behaviours in H2-SCR depending on the involved mass of the catalyst: () NOX conversion; yields in () N2, () N2O and () NH3. (a): 200 mg; (b): 100 mg; (c): 30 mg; (d): 5 mg.
Figure 3. 0.30 wt%Pt/Siral-5 behaviours in H2-SCR depending on the involved mass of the catalyst: () NOX conversion; yields in () N2, () N2O and () NH3. (a): 200 mg; (b): 100 mg; (c): 30 mg; (d): 5 mg.
Catalysts 15 00838 g003
Figure 4. 0.30 wt%Pt/Siral-5 (30 mg) behaviour depending on the reactional mixture (Table 2). (a): H2-SCR; (b): NO + H2 reaction; (c): NH3-SCR; (d): NH3 + O2 reaction. (): NOX conversion (a,b) or yield (c,d); (): N2 yield; (): N2O yield; (): NH3 yield (a,b) or conversion (c,d).
Figure 4. 0.30 wt%Pt/Siral-5 (30 mg) behaviour depending on the reactional mixture (Table 2). (a): H2-SCR; (b): NO + H2 reaction; (c): NH3-SCR; (d): NH3 + O2 reaction. (): NOX conversion (a,b) or yield (c,d); (): N2 yield; (): N2O yield; (): NH3 yield (a,b) or conversion (c,d).
Catalysts 15 00838 g004
Figure 5. 0.30 wt%Pt/Siral-5 catalyst (30 mg). (a): Apparent use of H2 during the H2-SCR test. Note that the H2 consumption axis includes a break to better highlight the distribution of N-containing products. (■): H2 conversion; (): burned H2 by O2; (): converted H2 leading to NH3; (Catalysts 15 00838 i001): converted H2 leading to N2O; (Catalysts 15 00838 i002): converted H2 leading to N2; black line: total H2 conversion. (b): H2 conversion depending on the reactional mixture (Table 2): (▬) H2-SCR reaction; () H2 + O2 reaction.
Figure 5. 0.30 wt%Pt/Siral-5 catalyst (30 mg). (a): Apparent use of H2 during the H2-SCR test. Note that the H2 consumption axis includes a break to better highlight the distribution of N-containing products. (■): H2 conversion; (): burned H2 by O2; (): converted H2 leading to NH3; (Catalysts 15 00838 i001): converted H2 leading to N2O; (Catalysts 15 00838 i002): converted H2 leading to N2; black line: total H2 conversion. (b): H2 conversion depending on the reactional mixture (Table 2): (▬) H2-SCR reaction; () H2 + O2 reaction.
Catalysts 15 00838 g005
Figure 6. NOX TPD profiles of () Siral-5 support and () 0.15%Pt/Siral-5 catalyst. Adsorption performed at 100 °C under 1000 ppm NO2 and 2% O2 in N2 until saturation followed by a purge under N2 at 100 °C, mcata = 200 mg.
Figure 6. NOX TPD profiles of () Siral-5 support and () 0.15%Pt/Siral-5 catalyst. Adsorption performed at 100 °C under 1000 ppm NO2 and 2% O2 in N2 until saturation followed by a purge under N2 at 100 °C, mcata = 200 mg.
Catalysts 15 00838 g006
Figure 7. (a): NOX conversion rate in H2-SCR (mol s−1 gcata−1) at 60 °C depending on the platinum density at the catalyst surface (molPt m−2); (b): NOx conversion rate in H2-SCR (mol s−1 gcata−1) at 60 °C divided by the density of surface platinum (molsurface Pt m−2) depending on the platinum particle size.
Figure 7. (a): NOX conversion rate in H2-SCR (mol s−1 gcata−1) at 60 °C depending on the platinum density at the catalyst surface (molPt m−2); (b): NOx conversion rate in H2-SCR (mol s−1 gcata−1) at 60 °C divided by the density of surface platinum (molsurface Pt m−2) depending on the platinum particle size.
Catalysts 15 00838 g007
Figure 8. (a): NOx conversion rate at 60 °C in H2-SCR (mol s−1 m−2) and (b): N2O formation rate at 60 °C in H2-SCR (mol s−1 gcata−1), both normalized by the amount of platinum (molPt g−1) and plotted depending on the platinum particle size.
Figure 8. (a): NOx conversion rate at 60 °C in H2-SCR (mol s−1 m−2) and (b): N2O formation rate at 60 °C in H2-SCR (mol s−1 gcata−1), both normalized by the amount of platinum (molPt g−1) and plotted depending on the platinum particle size.
Catalysts 15 00838 g008
Figure 9. 0.30 wt%Pt/Siral-5 catalyst (30 mg): comparison of conversions/yields depending on the inlet reactional mixture.
Figure 9. 0.30 wt%Pt/Siral-5 catalyst (30 mg): comparison of conversions/yields depending on the inlet reactional mixture.
Catalysts 15 00838 g009
Figure 10. Proposed pathways in the H2-SCR process depending on the temperature and related products over 0.30 wt%Pt/Siral-5 catalyst.
Figure 10. Proposed pathways in the H2-SCR process depending on the temperature and related products over 0.30 wt%Pt/Siral-5 catalyst.
Catalysts 15 00838 g010
Table 1. Specific surface area, platinum loading, mean platinum particle size and dispersion.
Table 1. Specific surface area, platinum loading, mean platinum particle size and dispersion.
CatalystSBET
(m2 g−1)
Pt Loading from ICP Analysis (wt%)Mean Pt Particle Size (nm) *Pt Dispersion (%)
0.30 wt%Pt/Al2O31440.397.0/-19
0.30 wt%Pt/Siral-52180.3410.4/10.613
0.30 wt%Pt/Siral-202450.3011.8/-11
0.30 wt%Pt/Siral-402320.3213.4/-10
0.15 wt%Pt/Al2O31490.13-/4.527
0.15 wt%Pt/Siral-52250.1611.4/11.212
0.15 wt%Pt/Siral-202390.15-/10.912
0.15 wt%Pt/Siral-402300.1314.3/13.59
0.07 wt%Pt/Siral-52690.078.2/8.415
0.30 wt%Pt/Siral-5 w/o hydrotreatment2240.231.9/-56
1% wtPt/Al2O3 (commercial) **1420.90-/2.644
* From H2 chimisorption/from TEM (examples of TEM images and particle size distribution are available in Appendix C). ** reduced under 10% H2 at 600 °C for 1 h.
Table 2. Reactional mixtures depending on the type of catalytic test (total flow: 150 mL min−1).
Table 2. Reactional mixtures depending on the type of catalytic test (total flow: 150 mL min−1).
Catalytic TestNO
(ppm)
NH3
(ppm)
H2
(%)
O2
(%)
N2
H2-SCR400-12balance
NO + H2 reaction400-1-balance
NH3-SCR400400-2balance
NH3 oxidation-400-2balance
H2 + O2 reaction--12balance
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

Ben Attia, A.; Can, F.; Courtois, X. H2-SCR over Low Loaded Platinum-Based Catalysts: Investigations in the Reaction Pathways. Catalysts 2025, 15, 838. https://doi.org/10.3390/catal15090838

AMA Style

Ben Attia A, Can F, Courtois X. H2-SCR over Low Loaded Platinum-Based Catalysts: Investigations in the Reaction Pathways. Catalysts. 2025; 15(9):838. https://doi.org/10.3390/catal15090838

Chicago/Turabian Style

Ben Attia, Amira, Fabien Can, and Xavier Courtois. 2025. "H2-SCR over Low Loaded Platinum-Based Catalysts: Investigations in the Reaction Pathways" Catalysts 15, no. 9: 838. https://doi.org/10.3390/catal15090838

APA Style

Ben Attia, A., Can, F., & Courtois, X. (2025). H2-SCR over Low Loaded Platinum-Based Catalysts: Investigations in the Reaction Pathways. Catalysts, 15(9), 838. https://doi.org/10.3390/catal15090838

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