Next Article in Journal
ortho-Substituted 2-Phenyldihydroazulene Photoswitches: Enhancing the Lifetime of the Photoisomer by ortho-Aryl Interactions
Previous Article in Journal
Towards the Use of Adsorption Methods for the Removal of Purines from Beer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress on Noble Metal-Based Catalysts Dedicated to the Selective Catalytic Ammonia Oxidation into Nitrogen and Water Vapor (NH3-SCO)

by
Magdalena Jabłońska
Institute of Chemical Technology, Universität Leipzig, Linnéstr. 3, 04103 Leipzig, Germany
Submission received: 21 September 2021 / Revised: 19 October 2021 / Accepted: 25 October 2021 / Published: 26 October 2021
(This article belongs to the Topic Nanomaterials for Sustainable Energy Applications)

Abstract

:
A recent development for selective ammonia oxidation into nitrogen and water vapor (NH3-SCO) over noble metal-based catalysts is covered in the mini-review. As ammonia (NH3) can harm human health and the environment, it led to stringent regulations by environmental agencies around the world. With the enforcement of the Euro VI emission standards, in which a limitation for NH3 emissions is proposed, NH3 emissions are becoming more and more of a concern. Noble metal-based catalysts (i.e., in the metallic form, noble metals supported on metal oxides or ion-exchanged zeolites, etc.) were rapidly found to possess high catalytic activity for NH3 oxidation at low temperatures. Thus, a comprehensive discussion of property-activity correlations of the noble-based catalysts, including Pt-, Pd-, Ag- and Au-, Ru-based catalysts is given. Furthermore, due to the relatively narrow operating temperature window of full NH3 conversion, high selectivity to N2O and NOx as well as high costs of noble metal-based catalysts, recent developments are aimed at combining the advantages of noble metals and transition metals. Thus, also a brief overview is provided about the design of the bifunctional catalysts (i.e., as dual-layer catalysts, mixed form (mechanical mixture), hybrid catalysts having dual-layer and mixed catalysts, core-shell structure, etc.). Finally, the general conclusions together with a discussion of promising research directions are provided.

1. Introduction

Ammonia (NH3) is a corrosive, highly toxic, and reactive inorganic gas with a pungent odor under ambient conditions. It is an atmospheric pollutant that is dangerous to health and life because it could corrode skin, eyes, and lungs, and cause permanent injury or even death when the concentration is higher than 300 ppm [1,2]. Ammonia is also reported to be the most common pollutant found in water streams, further affecting human health as the consequence of eating toxic fish and drinking water [3]. The toxic action of ammonia on aquatic animals can lead to the extinction of the entire population, threatening many important ecosystems and fisheries worldwide [4,5]. NH3 is referred to as one of four major atmospheric pollutants together with NOx, SOx, and nonmethane volatile organic compounds (NMVOC). Approximately 5600 kt y−1 of ammonia are emitted into the atmosphere each year, i.e., up to 4-times higher emission levels than in the previous century, and it continues to increase [6]. NH3 is emitted by several various processes, including nitric acid production, urea manufacturing, nitrogen fertilizer production, biomass, and coal gasification, petroleum refining and refrigeration, livestock waste, and animal agriculture, transport (as a gas slip from the process of selective catalytic reduction of NOx using NH3 or urea (SCR) in DeNOx applications), etc. More attention was given to the removal of NH3 from gaseous and waste streams, e.g., through its oxidation.
The selective catalytic oxidation of ammonia (NH3-SCO) into nitrogen and water vapor is considered as the most promising method for the elimination of NH3 from oxygen-containing exhaust gases (Equation (1)) [7]:
4 NH3 + 3 O2 → 2 N2 + 6 H2O
NH3 is generated by an onboard aqueous urea dosing system. The obtained ammonia acts as a NOx reductant in the DeNOx process. There is a serious risk that unreacted ammonia is released into the atmosphere. Thus, the active SCO catalysts (so-called guard catalyst, ammonia slip catalyst—ASC, ammonia oxidation catalyst—AMOX) should operate in a broad temperature range (up to 600–700 °C—in the cycle of diesel particulate filter regeneration) in the presence of typical components of the exhausts (H2O, COx, and SOx), and additionally should selectively direct the reaction to the formation of N2 and H2O. Euro VI emission standards for heavy-duty vehicles (HDVs) introduced for the first time limits for NH3 emissions up to 10 ppm [8]. Currently, there are no limits for NH3 emitted from light-duty vehicles (LDVs, i.e., passenger cars), despite their high levels of emissions (e.g., [9,10]). Thus, potentially NH3 will be considered next to ultra-fine particles smaller than 23 nanometers (PN10) and nitrous oxide (N2O) in the upcoming regulations, e.g., upcoming Euro emission standards.
Various kinds of catalysts were studied for NH3-SCO, including noble metals (e.g., Ru [11]), supported noble metals (e.g., Au/Nb2O5, Au/ZrO2 [12]), (mixed) metal oxides, supported metal oxides, modified zeolites (e.g., Pt-CuMgAlOx [13], Ag-USY [13]), etc. These groups of catalysts investigated in NH3-SCO were summarized by Jabłońska et al. [7,14], Gao et al. [15] and Lan et al. [16]. In general, noble metal-based catalysts tend to possess high activity at low temperatures (< 300 °C), while their high cost and relatively low N2 selectivity have restrained their widespread application. Transition metal-containing oxides and zeolites show improved selectivity toward N2 than noble metal-based catalysts; however, they need higher operating temperatures (300–600 °C). Consequently, the proper combination of these two metals (in the form of bifunctional catalysts) could produce the catalysts with enhanced activity and N2 selectivity. In general, the concept of bifunctional catalysts is based on the internal selective catalytic reduction (i-SCR) mechanism. This mechanism consists of two main steps. In the first step, part of ammonia is oxidized to NO and NO2—minor by-product, while in the second step, NO and NO2 are reduced by ammonia (unreacted in the first step) with the formation of N2 and H2O (DeNOx, NH3-SCR). In this step, also the formation of N2O is possible. Besides the i-SCR mechanism, other major reaction pathways, i.e., the imide (NH, in which NH3 transforms to N2 and N2O as final products, with nitrosyl (HNO) as an intermediate) mechanism and the hydrazine (with the formation of a hydrazine-type (N2H4) intermediate) mechanism were proposed for NH3-SCO over different types of catalysts. The details of the above-mentioned reaction mechanisms can be found in previous review articles (e.g., [7,14]). Additionally, due to lack of identification of the characteristic intermediates of the aforementioned three mechanisms, recently Wang et al. [17] proposed a N2 mechanism of NH3-SCO on a Ag/nano-Al2O3. Li et al. [18] postulated an Eley–Ridel mechanism over perovskite-based catalysts, where gaseous NH3 reacts with adsorbed -ONH2 species to form the surface diazo species (-N=N-) with the rate-determining step depending on the catalysts composition.
Recent interest focuses on bifunctional catalysts consisting of noble metal-based catalyst and transition metal-based catalyst. Previous review articles, including ammonia oxidation, give a clear statement about high activity, N2 selectivity, and stability over Cu-containing materials (Jabłońska et al., 2016, Jabłońska, 2020) [7,19]. From the noble metal-based catalysts the most frequently applied—Pt/Al2O3, provides high activity but also significant selectivity to byproducts (N2O and NOx). Besides Pt/Al2O3, also the Ag/Al2O3 catalyst has received extensive concerns on the low-temperature NH3-SCO. Other systems were less investigated for NH3-SCO than Pt/Al2O3 or Ag/Al2O3. Thus, the present mini-review aims to provide a broad picture of the property-activity correlations of noble metal-based catalysts investigated for NH3-SCO up to now (including Pt-, Pd-, Ag- and Au-, Ru-based catalysts). The NH3-SCO catalysts are classified considering their full NH3 conversion and N2 selectivity in the same temperature range. If not provided in the references, the catalytic activity and N2 selectivity were roughly estimated based on NH3 conversion-temperature profiles. This overview will shed light on future research directions regarding catalyst composition and architecture that maximizes the oxidation of NH3 into N2 over a broad temperature range and in the presence of the typical components of exhaust, such as H2O, SOx, and COx.

2. Pt- and Pd-Based Catalysts

Early work on NH3 oxidation was given by Il’chenko et al. [20] Among metal-based catalysts, Pt and Pd are the most active for the ammonia oxidation (p(NH3) = 0.1 atm, p(O2) = 0.9 atm) but also the most selective to N2O—according to the following order: Pt, Pd > Ni > Fe > W > Ti (note that Rh was not mentioned in Il’chenko review). The catalytic ammonia oxidation over platinum is a key step in the industrial manufacturing of nitric acid. The ammonia oxidation exhibits a moderate structure sensitivity, while the activity decreased in the order of: Pt foil > Pt(865) > Pt(533) > Pt(443) > Pt(100) due to different oxygen sticking coefficient [21]. Novell–Leruth et al. [22] used periodic slab density functional theory (DFT) calculations and found that NH3 adsorbs preferentially on the top sites, NH2 (dehydrogenation intermediates) on the bridge sites, while NH and N species on the hollow sites on both the (111) and (100) surfaces. The ammonia oxidation with atomic or molecular oxygen over Pt(100), Pt(111), stepped Pt(111)/Pt(211) or terrace Pt(111) orientations, etc., yields N2, NO, N2O, and H2O, in varying amounts depending on reactant conditions. Steady-state reaction studies [23] under ultra-high vacuum (UHV) conditions with a stepped Pt(111) surface revealed that excess NH3 lead to N2 formation, while under excess O2, NO formation was preferred. No other nitrogen-containing products, i.e., N2O were detected in the gas phase (note that N2O was never observed under UHV conditions [24]). Similar conclusions were given by Pérez–Ramirez et al. [25], who studied the sequence of steps in NH3 oxidation (applying the isotope 15NH3) over Pt, Pd, and Rh wires in the temporal analysis of products (TAP) reactor at relevant temperatures in industrial ammonia burners. High NO selectivity is favored at a high ratio of adsorbed n(O)/n(NHx) species, e.g., at c(O2)/c(15NH3) = 0.1, the 15NO selectivity over Pt reached 45%, while at c(O2)/c(15NH3) = 10—ca. 100% selectivity. NO was found to be a primary reaction product in NH3 oxidation, while N2 and N2O originate from consecutive NO transformations. Pd and Rh were more active for the reduction of nitric oxide by ammonia than Pt (Figure 1). Additionally, DFT calculations showed that the N2O formation over Rh(100) plane needs higher activation energy than over Pt(100) or Pd(100). Furthermore, Rh(100) was more active in NH3 decomposition (possessed a lower activation barrier for the NH3 → NH2 step) than Pt and Pd surfaces, and strongly stabilized the dehydrogenated NH and N species [22].
γ-Al2O3 and ZSM-5 are often used as the supports for noble metal-based catalysts for the selective catalytic oxidation of NH3. A summary of Pt-, Pd-based catalysts is presented in Table 1. Pt/Al2O3 is usually applied to provide high low-temperature activity. The main drawback of such catalyst is low N2 selectivity due to the formation of N2O (below 250 °C) and NOx (above 250 °C). The metallic Pt is significantly more active for NH3-SCO than oxidized platinum [26,27], which provide limited sites for O2 dissociation [28]. The operando XANES/EXAFS studies revealed the highest N2 selectivity (ca. 80%) over H2-reduced (2 wt.%)Pt/Al2O3. Nevertheless, under reaction conditions, at least 40% of Pt surface remains oxidized resulting in the formation of N2O [29]. Otherwise, the metal-support interactions of Pt/TiO2 were reported to stabilize Pt in the metallic state (also under reaction conditions) [30]. The Pt0 content can also be manipulated by preparation procedure of Pt/SiO2-Al2O3, i.e., by adding ascorbic acid (vitamin C, vC; n(vC)/n(Pt) = 0.25–1.5) [31]. Ostermaier et al. [26] reported that small Pt particles (2.0 and 2.7 nm) of (1–2.93 wt.%)Pt/Al2O3 demonstrate lower activity in comparison to larger crystallites (15.5 nm). The catalysts with a small size of Pt0 crystallites were characterized by the strongest deactivation during NH3 oxidation due to their oxidation to PtOx, where x depends on the particle size [32]. Later, Sobczyk et al. [33,34] demonstrated with positron emission profiling (PEP) that the catalysts deactivate due to poisoning of the surface mainly by nitrogen species (NH and NH2).
The dispersion of Pt species was increased after the introduction of ethylenediamine (from 2.8 to 2.0 nm) during the preparation of (1 wt.%)Pt/SiO2-Al2O3 [35], and thus lead to higher activity in NH3-SCO in the presence of CO2 and H2O, compared to that of unmodified samples. The N2 selectivity remained nearly unaffected by the Pt particle size. Contrary to these studies, Slavinskaya et al. [36] found that a larger Pt particle size (ca. ~23 nm compared to ~1 nm) of (2 wt.%)Pt/Al2O3 enhanced activity. Additional measurements in the presence of CO2 and H2O did not change the trends of activity and selectivity of Pt/Al2O3 on the Pt dispersion and Pt state. Similar to the above discussed studies over Pt/SiO2-Al2O3, N2 selectivity did not depend on the Pt particle size, while in all cases, N2 selectivity was below 70% (>300 °C). Furthermore, authors [27] showed no deactivation of the catalysts, i.e., the oxidation state of platinum in Pt/Al2O3 did not increase after the catalytic experiments. Also, the hydrothermal aging (in a feed containing O2, H2O, CO2, at 550 °C over 250 h) of a Pt/Al2O3 washcoated monolith did not influence its activity below 250 °C [37]. Above 300 °C the activity significantly decreased with aging time (0, 122, 253 h) but the product selectivity remains the same. Recently, Machida et al. [38] found that a thin-film catalyst, which was prepared by deposing a nanoscale-thickness Pt(111) overlayer on a 50 μm-thick Fe-Cr-Al metal foil (Pt/SUS) achieved more than 180-fold higher TOF compared with the conventional (0.13 wt.%)Pt/Al2O3. The thermal stability of Pt/SUS was enhanced by the insertion of the Zr layer between the Pt and SUS foil. For the Pt surfaces, the NH mechanism was mostly proposed by experimental and DFT simulation studies; e.g., over Pt(100) or Pt(111) [39,40,41]. Also, the so-called NH mechanism (i.e., NH as the intermediates in the imide mechanism) occurred on Pt/Al2O3, while the HNO and N2H4 mechanism (i.e., HNO and N2H4 as the intermediates in the imide and hydrazine mechanism, respectively) coexisted on Pt/CeZrO2 (Figure 2) [42].
Li and Armor [43] studied a series of zeolite ZSM-5 ion-exchanged with (4.07 wt.%)Pd, (2.66 wt.%)Rh, and (2.55 wt.%)Pt as catalysts for NH3-SCO. Among them, relatively high activity and N2 selectivity in the low-temperature range (≤300 °C) were found for the Pd-containing catalysts (i.e., full ammonia conversion with 91% N2 selectivity at 300 °C in the presence of 5 vol.% H2O). 58–61% of N2O selectivity on (2.55 wt.%)Pt-ZSM-5 and 16–25% on (4.01 wt.%)Pd-ZSM-5 was obtained at 250–300 °C. The noble metal-exchanged ZSM-5 materials were less affected by water vapor than the corresponding Al2O3 supported catalysts. Similar results, i.e., 92% NH3 conversion and 73% N2 selectivity at 350 °C over (5.51 wt.%)Pd-ZSM-5 in NH3-SCO (feed without H2O), were also reported by Long and Yang [44]. Furthermore, Jabłońska et al. [45] investigated zeolites HY modified with palladium (0.05–2.5 wt.%). An increase in Pd loading leads to higher catalysts activity together with the drop in N2 selectivity. The palladium oxide species (PdOx) were found to be active sites for ammonia oxidation (based on FT-IR studies). A part of ammonia was stabilized against oxidation over the zeolite framework acid sites (in the form of NH4+), and lead to enhanced N2 selectivity in the higher temperature range. The analysis of the results of temperature-programmed (NH3-SCO with various spaces velocity) and spectroscopic studies lead to the conclusion that the ammonia oxidation over Pd-Y followed the i-SCR mechanism. On the other hand, an appearance of hydrazine species (intermediates in the hydrazine mechanism) on the ammonia pre-adsorbed Pd-Y catalyst at 250 °C during FT-IR studies, suggested that the ammonia oxidation is more complicated and followed different parallel routes. Otherwise, Wells et al. [46,47] identified PdNx under reaction conditions over H2-reduced (1.5 wt.%)Pd/Al2O3 and Pd/Y (n(Si)/n(Al) = 2.6) as the dominant species during N2 formation (based on combined operando spectroscopy and DFT calculations). As stated above, palladium-containing materials appear as promising NH3-SCO catalysts, although, their stability (also in the presence of H2O, SOx, and COx) needs confirmation.
The state-of-the-art NH3-SCO systems include a combination of a noble metal-based catalyst—usually Pt/Al2O3, and an SCR catalyst, e.g., Cu- or Fe-containing zeolite. Thus, a part of ammonia is oxidized over the noble metal-based catalyst to N2 and NOx which is further transformed to N2 over the SCR catalyst. Different arrangements of both metals were reported in the literature, i.e., noble/transition metal deposited on one support in systems such as (0.05 wt.%)Pt/(1 wt.%)Cu/Al2O3 [48], (0.5–4 wt.%)Pt/(20 wt.%)Cu/Al2O3 [49], (1 wt.%)Pt/(20 wt.%)Cu/Al2O3 [50,51,52], (1.5 wt.%)Pt-(5.5 wt.%)Cu/ZSM-5 [53], (0.5 wt.%)Pt-(1.54 wt.%)Fe-ZSM-5 [54], (1.5 wt.%)Pt-(0.5 wt.%)Fe/ZSM-5 [55], (0.21 wt.%)Pt/CuMgAlOx hydrotalcite-derived mixed metal oxides [13], (2 wt.%)PdO/(5 wt.%)CuO/Al2O3 [56], etc. Besides such form, the active components may be present in different configurations—dual-layer configuration, mixed and hybrid layer sample types, that are presented in Figure 3. The dual-layer catalytic systems consist of noble metal-based catalyst as a bottom layer, and transition metal-based catalysts as an upper layer, i.e., Pt/Al2O3 and Cu-ZSM-5, Pt/Al2O3 and Cu-SSZ-13, Pt/Al2O3, and Fe-ZSM-5 investigated by Shrestha et al. [57,58,59], Pt/Al2O3 and Fe-zeolite investigated by Scheuer et al. [60] and Colombo et al. [61]. E.g., Shrestha et al. [57] pointed out the increase of N2 selectivity (with a corresponding decrease in NO selectivity) with increasing copper loading of Cu/ZSM-5 (NH3-SCR layer), e.g., from 58% to 82% at 250 °C for 0.8 and 2.5 wt.% of Cu, respectively. Still, N2O selectivity reached a maximum of about 40% at 260 °C in all cases. Similar N2O selectivity was reported over Pt/Al2O3 and Fe-ZSM-5 arranged as both dual-layer and mixed catalysts [58]. Also, the NH3 oxidation depended on the applied conditions, i.e., space velocity (66,000 versus 265,000 h−1). At higher space velocity mixed catalyst revealed (ca. 7%) higher NH3 conversion, while dual-layer catalyst provided higher N2 selectivity (especially above 350 °C). The hybrid catalyst (bottom layer of mixed Cu-SSZ-13 + Pt/Al2O3, top layer of Cu-SSZ-13) allowed to achieve (ca. 5%) higher NH3 conversion than that of the dual-layer catalyst [59]. Furthermore, Dhillon et al. [62] applied sacrificial agents (yeast or polymer) to generate macropores on a (2.90 wt.%)Cu-SSZ-13 top-layer washcoat supported on (1.47 wt.%)Pt/Al2O3, and thus to enhance NH3 conversion (still below 90% up to 500 °C) without impact on N2 selectivity. Recently, Gosh et al. [63] reported a Pt/Al2O3@Cu-ZSM-5 (0.05 wt.% Pt, 2.93 wt.% Cu) core-shell catalyst that allowed full NH3 conversion at ca. 300 °C and 100% N2 selectivity (up to 275 °C). The NH3 conversion was negligibly affected by the variation in the shell thickness (0.5 µm versus 1.2 µm), while the ticker shell was beneficial in improving N2 selectivity at higher temperatures. The addition of H2O (5 vol.%) to the feed had a minor impact on the catalyst activity, while more tests in the presence of COx and SOx are still required.
Concluding this part, as can be seen from the above presented data, a great variety of Pt- and Pd-based catalytic systems were developed and tested for NH3-SCO. In particular, (1–2 wt.%)Pt/Al2O3 was reported as one of the most active catalyst that allowed to obtain full NH3 conversion around 200–450 °C with N2 selectivity of 15–87% (according to data gathered in Table 1), depending on the applied catalyst preparation procedure as well as pretreatment and reaction conditions. Pt0 serves as the active species responsible for high catalytic activity, and thus most of the catalytic systems were prepared by an impregnation method and subsequently reduced in H2. Great research efforts aimed at determining the role of Pt dispersion in NH3-SCO, but there is no clear consensus about it yet. Platinum (0.05–1.5 wt.%) was also applied as the most active noble metal component in the bifunctional catalysts. Such systems fully oxidized NH3 in a broad temperature range of 195–500 °C with 44–100% N2 selectivity (according to data gathered in Table 1) depending on their architecture, i.e., Pt deposited on one support with transition metal component, as dual-layer catalysts, mixed form, hybrid catalysts or incorporated in the core-shell structure, etc. Among them the (0.46 wt.%)Pt/Al2O3-(2.5 wt.%)Cu-ZSM-5 dual-layer catalyst (full NH3 conversion at 250–500 °C, 82–100% N2 selectivity) and (0.05 wt.%)Pt/Al2O3@(2.93 wt.%)Cu/ZSM-5 core-shell catalyst (full NH3 conversion at 310–500 °C, 91–94% N2 selectivity) appear as the most interesting systems for further catalysts optimization. However, the stability of these catalysts and the influence of the potential catalyst pollutants usually present in the exhausts, i.e., H2O, SOx and COx, were not fully provided within the scope of the studies. A similar conclusion can be given for other Pt- or Pd-containing catalysts (only a few materials were tested in the presence of H2O and CO2). Hence, further studies are required to understand structure-activity relationships and reaction mechanisms under application-relevant reaction conditions.
Figure 3. Schematic diagram representing bifunctional ammonia slip catalyst in three different washcoated structured methodologies. Reprinted from [64] with permission from Science Direct.
Figure 3. Schematic diagram representing bifunctional ammonia slip catalyst in three different washcoated structured methodologies. Reprinted from [64] with permission from Science Direct.
Molecules 26 06461 g003
Table 1. Comparison of full NH3 conversion and N2 selectivity in same temperature range over Pt-based catalysts reported in literature.
Table 1. Comparison of full NH3 conversion and N2 selectivity in same temperature range over Pt-based catalysts reported in literature.
CatalystCatalyst PreparationReaction ConditionsNH3 Conversion_N2 Selectivity/%
(Temperature/°C)
Ref.
(3 wt.%)Pt-Rhwash-coated on Al2O3, calcination, 400 °C, air0.08 vol.% NH3, 4 vol.% O2, He balance, GHSV 92,000 h−1100/62% (400 °C)[65]
(4.4 wt.%)Pt-Rhwash-coated on Al2O3, calcination, 500 °C, air0.1 vol.% NH3, 4 vol.% O2, He balance, GHSV 92,000 h−1100/80% (400 °C)[66]
(1.2 wt.%)Pt/Al2O3 impregnation, calcination, 600 °C, air; reduction conditions not shown)0.1 vol.% NH3, 10 vol.% O2, He balance, 50 mL min−1, mass of the catalyst: 0.1 g, WHSV 30,000 mL h−1 g−1100/75% (200 °C)[67]
1.14 vol.% NH3, 8.21 vol.% O2, He balance, 74.7 mL min−1, mass of the catalyst: 0.2 g, WHSV 22,410 mL h−1 g−1100/87% (200 °C)
(1.73 wt.%)Pt/Al2O3 impregnation, calcination, 400 °C, air; reduction, 250 °C, H20.1 vol.% NH3, 4 vol.% O2, He balance, 500 mL min−1, mass of the catalyst: 0.145 g, GHSV 120,000 h−1100/40–60% (200–400 °C)[27]
(2 wt.%)Pt/Al2O3impregnation, calcination, 400 °C, air; oxidation, 400 °C, O2/He0.1 vol.% NH3, 4 vol.% O2, He balance, 500 mL min−1, mass of the catalyst: 0.145 g, GHSV 120,000 h−1 100/40–50% (325–400 °C) [36]
impregnation, calcination, 400 °C, air; reduction, 350 °C, H2; calcination, 400 °C, Ar100/30–60% (250–400 °C)
(1 wt.%)Pt/Al2O3impregnation, calcination, 550 °C, air0.02 vol.% NH3, 10 vol.% O2, N2 balance, 2.31 l min−1, mass of the catalyst: 0.34 g, WHSV 407,000 mL g−1 h−1100/25–49% (250–450 °C)[68]
(1 wt.%)Pt/CeO2-SiO2 100/15–42% (225–450 °C)
(1 wt.%)Pt/SiO2-Al2O3impregnation, calcination, 550 °C, air; monolithic catalyst0.02 vol.% NH3, 10 vol.% O2, 8 vol.% CO2, 5 vol.% H2O, N2 balance, GHSV 100,000 h−1100/10–40% (300–450 °C)[35]
(1 wt.%)Pt/SiO2-Al2O3impregnation; treatment strategies—conditions not shown; monolithic catalyst; Vc—ascorbic acid0.02 vol.% NH3, 10 vol.% O2, 8 vol.% CO2, 5 vol.% H2O, N2 balance, GHSV 100,000 h−1100/28–60% (240–300 °C)[31]
(1 wt.%)Pt/SiO2-Al2O3-vC 100/22–50% (240–300 °C)
(1 wt.%)Pt/Al2O3impregnation, calcination, 550 °C, air; monolithic catalyst0.02 vol.% NH3, 8 vol.% O2, N2 balance, GHSV 100,000 h−1100/15–30% (300–400 °C)[42]
(1 wt.%)Pt/CeZrO2100/20–45% (325–400 °C)
(1.5 wt.%)Pt/ZrO2impregnation, calcination, 550 °C, air0.018 vol.% NH3, 8 vol.% O2, N2 balance, GHSV 100,000 h−1100/25–60% (350–500 °C)[53]
(1.5 wt.%)Pt-(5 wt.%)W/ZrO2100/28–50% (300–500 °C)
(2.0 wt.%)Pt/TiO2impregnation, calcination, 400 °C, air; oxidation, 400 °C, O2/He0.1 vol.% NH3, 4.0 vol.% O2, He balance, 500 mL min−1, mass of the catalyst: 0.145 g, WHSV 206,897 mL h−1 g−1100/38–55% (200–400 °C)[30]
pulsed laser ablation in liquids, calcination, 400 °C, air oxidation, 400 °C, O2/He100/22–50% (175–400 °C)
(0.1 wt.%)Pt/(2 wt.%)V/TiO2 impregnation, reduction, 600 °C, H2/N20.02 vol.% NH3, 8.0 vol.% O2, 6.0 vol.% H2O, He balance, 500 mL min−1, mass of the catalyst: 0.25 g, GHSV 60,000 h−1100/63–81% (250–350 °C)[69]
(1.2 wt.%)Pd/Al2O3 impregnation, calcination, 600 °C, air; reduction—conditions not shown1.14 vol.% NH3, 8.21 vol.% O2, He balance, 74.7 mL min−1, mass of the catalyst: 0.2 g, WHSV 22,410 mL h−1 g−1100/98% (300 °C)[67]
(4.2 wt.%)PdO/Al2O3impregnation, calcination, 500 °C, air**reduction, 400 °C, H2/He0.1 vol.% NH3, 4.0 vol.% O2, He balance,
* 0.1 vol.% NH3, 4.0 vol.% O2, 5 vol.% H2O, He balance,
100 mL min−1, mass of the catalyst: 0.1 g, WHSV 60,000 mL h−1 g−1
100/67% (350 °C)
** 100/86% (300 °C)
[43]
(4.07 wt.%)Pd-ZSM-5ion-exchange, calcination—conditions not shown* 100/91% (300 °C)
(1.5 wt.%)Pd/Y impregnation, calcination—conditions not shown)0.5 vol.% NH3, 2.5 vol.% O2, He balance, 40 mL min−1, mass of the catalyst: 0.05 g, GHSV 15,400 h−1100/80–90% (250–500 °C)[45]
(0.05 wt.%)Pt/(1 wt.%)CuO/Al2O3impregnation, calcination air, 600 °C; impregnation, calcination air, 500 °C0.5 vol.% NH3, 2.5 vol.% O2, He balance, 40 mL min−1, mass of the catalyst: 0.05 g, GHSV 15,400 h−1100/44–73% (325–500 °C)[48]
(1 wt.%)Pt/(20 wt.%)CuO/Al2O3impregnation, calcination air, 500 °C; impregnation, calcination air, 450 °C0.07 vol.% NH3, 0.5 vol.% O2, He balance
*0.07 vol.% NH3, 8 vol.% O2, He balance
1000 mL min−1, WHSV 180,000 mL h−1 g−1
100/88 (210–230 °C)
* 100/95 (230 °C)
[49]
(4 wt.%)Pt/(20 wt.%)CuO/Al2O3100/83 (195–230 °C)
* 100/90 (220–230 °C)
(1.5 wt.%)Pt-(0.5 wt.%)Fe/ZSM-5 ion-exchange, calcination, air, 500 °C, impregnation calcination, air, 500 °C0.1 vol.% NH3, 2 vol.% O2, He balance, 500 mL min−1, WHSV 500,000 mL h−1 g−1100/61–88% (200–300 °C) [55]
(0.5 wt.%)Pt-(1.54 wt.%)Fe/ZSM-5 ion-exchange, calcination, air, 500 °C, impregnation calcination, air, 500 °C0.1 vol.% NH3, 2 vol.% O2, He balance, 500 mL min−1, GHSV 230,000 h−1100/77–89% (250–400 °C)[54]
(1.5 wt.%)Pt-(5.5 wt.%)Cu/ZSM-5impregnation calcination, air, 550 °C0.018 vol.% NH3, 8 vol.% O2, N2 balance, GHSV 100,000 h−1100/56–73% (275–450 °C)[53]
(0.21 wt.%)Pt/CuMgAlOximpregnation, calcination, air, 500 °C0.5 vol.% NH3, 2.5 vol.% O2, He balance, 40 mL min−1, mass of the catalyst: 0.05 g, GHSV 15,400 h−1100/67–89% (350–500 °C)[13]
(0.21 wt.%)Pd/CuMgAlOx100/71–76% (425–500 °C)
(0.46 wt.%)Pt/Al2O3-(0.8 wt.%)Cu-ZSM-5Pt/Al2O3: impregnation, calcination, 500 °C, air; reduction, 500 °C, H2/Ar; Cu-ZSM-5: preparation not provided; oxidation, 650 °C, O2/Ar; monolithic catalyst
*dual layer catalyst
0.05 vol.% NH3, 5 vol.% O2, Ar balance, GHSV 66,000 h−1 * 100/58–74% (250–500 °C) [57]
(0.46 wt.%)Pt/Al2O3-(2.5 wt.%)Cu-ZSM-5 * 100/82–100% (250–500 °C)
(0.46 wt.%)Pt/Al2O3-Fe-ZSM-5 Pt/Al2O3: impregnation, calcination, 500 °C, air, reduction, 500 °C, H2/Ar, oxidation, 650 °C, O2/Ar; Fe-ZSM-5: commercial; oxidation, 650 °C, O2/Ar; monolithic catalyst
*dual layer catalyst
**mixed catalyst
0.05 vol.% NH3, 5 vol.% O2, Ar balance, GHSV 66,000 h−1 * 100/48–93% (250–500 °C)
** 100/57–82% (250–500 °C)
[58]
(0.05 wt.%)Pt/Al2O3@(2.93 wt.%)Cu/ZSM-5 core-shell catalyst, Pt/Al2O3: impregnation, calcination, 550 °C, air; Cu-ZSM-5: ion-exchange, calcination, 500 °C, air0.05 vol.% NH3, 5 vol.% O2, Ar balance, 100 mL min−1, mass of the catalyst: 0.18 g, GHSV 280,000 h−1 100/91–94% (310–500 °C)[63]

3. Ag-Based Catalysts

Il’chenko et al. [20,70] reported that the specific activity of metal Ag at 300 °C was lower than that of Pt and Pd. Among silver-based catalysts, γ-Al2O3 impregnated with Ag species (mainly 10 wt.%) was widely investigated. Depending on the applied conditions, i.e., catalyst (its preparation, pre-treatment strategies, etc.) and reaction conditions, the full NH3 conversion can be reached in the range of 150–400 °C with 45–95% N2 selectivity over (10 wt.%)Ag/Al2O3 (according to data gathered in Table 2). However, at temperatures above 300 °C N2 selectivity dropped due to the large NO production. For NH3-SCO, the Ag/Al2O3 catalysts are mainly applied after H2 pretreatment. E.g., Gang et al. [71,72] reported extremely high activity of Ag/Al2O3 at 160 °C (full NH3 conversion with N2 selectivity of about 82%), which was even superior to H2-reduced Ir/Al2O3 or Pt/Al2O3. The activity of Ag/Al2O3 was also higher than over silver powder and Ag/SiO2 [72], indicating that the applied support influenced the Ag particle dispersion. However, the difference in the Ag particle size of the Ag/Al2O3 (8.2 nm) and Ag/SiO2 (24 nm) catalysts was not discussed in these studies. The authors correlated the NH3 oxidation activity at low temperatures to the catalysts’ ability to promote dissociative or nondissociative adsorption of O2. However, again the role of different oxygen species (i.e., adsorbed molecular oxygen, adsorbed atomic oxygen, subsurface oxygen, and bulk dissolved oxygen) in the activity and the reaction mechanisms was not fully explored. Zhang and He [73] reported that the dissociation of O2 is a rate-determining step for NH3-SCO. They concluded that molecular O2 can be dissociatively chemisorbed on the surface of H2-reduced Ag/Al2O3, i.e., metallic species (in contrast to fresh material) to form O species, and thus enhance NH3-SCO activity [74]. Furthermore, the modification of Ag/Al2O3 with CeO2 improved catalysts’ ability in the adsorption and activation of O2 to form O species [75]. However, Wang et al. [76] claimed that the recovery of Brønsted acid sites via H2 reduction (i.e., break of Ag-O bonds on the Ag/Al2O3 surface and formation of Ag clusters in the metallic state—Agn0, based on EXAFS analysis) is also responsible for the improved activity of H2-reduced Ag/Al2O3. Highly dispersed particles of Ag0 (3.5–25 nm) were found to enhance the catalytic activity below 140 °C, whereas large particles (12–50 nm) of Ag0 were responsible for improved N2 selectivity [74].
Furthermore, N2 selectivity of about 85% above 300 °C over (2.2 wt.%)Ag/Al2O3 (c(NH3):c(O2) = 1:1–1:25), was assigned to the small Ag particle size (<5 nm, based on XRD analysis) [77]. Ag+ cations are the main active species in NH3-SCO at temperatures above 140 °C. The adsorbed NH3 mainly reacts with the gaseous O2 over fresh Ag/Al2O3. Although besides Ag0 and Ag+, also Agnδ+ species were evidenced by DR UV-Vis analysis, the authors did not specify their role in ammonia oxidation. Also, Qu et al. [78] obtained highly dispersed Ag0 particles with a size of 5–6 nm (based on XRD and DR UV-Vis analyses) on the H2-reduced (10 wt.%) Ag/Al2O3 catalyst, which reached full conversion and 89% of N2 selectivity at 180 °C. Yang et al. [79] and Jabłońska et al. [80] studied (1–10 wt.%)Ag/Al2O3 and claimed that the low N2 selectivity above 200–300 °C over Ag/Al2O3 was ascribed to the formation of Ag2O crystals.
Besides the influence of the Ag particle size, Wang et al. [17] investigated the effect of the different support particle size (micro-Al2O3 versus nano-Al2O3) on the activity of the final Ag/Al2O3 catalysts in NH3-SCO. The catalyst characterization indicated that nano-Al2O3 was beneficial for Ag dispersion (the average Ag particle size of 3.7 nm, based on HRTEM analysis). The catalyst abundant acid sites (based on NH3-TPD analysis) facilitated the adsorption and dissociation of NH3, therefore resulting in an enhanced activity. Furthermore, the same research group [81] studied Ag supported on TiO2, SiO2 as well as their mixture—Ag/SiO2-TiO2. Although the (10 wt.%)Ag/SiO2-TiO2 catalyst reached full NH3 conversion at 140 °C, N2 selectivity in the whole studied range of 100–240 °C was below 70%. Significantly higher N2 selectivity was obtained over Ag/TiO2 (91–99% at 180–240 °C). Jabłońska et al. [82] compared commercial TiO2 with mesoporous TiO2 (prepared by evaporation induced self-assembly (EISA)) as support in NH3-SCO. The activity and N2 selectivity were favored over (1.5 wt.%)Ag-doped mesoporous TiO2 (calcined at 600 °C, with predominant anatase phase). The easily reducible highly dispersed oxidized silver species were converted into Ag0 and possibly Agnδ+ clusters through in situ H2-pretreatment of catalyst. The metallic silver decomposed N2O into N2 and surface oxygen species, leading to higher N2 selectivity. Further studies, concerning the stability tests, revealed that these materials are unstable, especially in the higher temperature range (>600 °C). However, temperature of the full conversion of NH3 over (9.8 wt.%)Ag/Al2O3 also gradually increased after ca. 4 reaction cycles (from 150 to 250 °C). A higher stability in subsequent catalytic runs revealed (9.9 wt.%)Ag/ZSM-5 with the postsynthetic modified support. The micro-/mesoporous structure could prevent sintering and/or leaching of Ag particles during NH3-SCO [83].
As stated above, while a broad number of studies examined NH3-SCO over Ag-based catalysts, the mechanism of NH3 oxidation and N2 formation is still uncertain, and the studies are mainly based on temperature-programmed (TPD) or in situ DRIFTS studies. E.g., Gang et al. [84] investigated NH3-SCO over powder silver by TPD and in situ FT-Raman spectroscopy. They found NO as the main reaction intermediate yielding N2O and/or N2 (Figure 4a). The dissociation of oxygen was believed to be the rate-controlling step for ammonia oxidation, while low surface coverage favors N2 formation. Similar conclusions were given by Karatok et al. [85] The exposure of ozone on Ag(111) surfaces at −133 °C led to a disordered surface atomic oxygen overlayer (confirmed by LEED). Such oxygen species selectively catalyzed N-H bond cleavage, yielding mostly N2 and minor amounts of by-products (NO and N2O). Higher coverage O/Ag(111) surfaces at −133 °C led to bulk-like amorphous silver oxide species, forming NO and N2O (Figure 4b). The ordered oxide surfaces—obtained through annealing of atomic oxygen-covered Ag(111) surface at 200 °C in UHV, showed only limited reactivity toward ammonia. Suppression of the N2 formation at high oxygen coverages was also reported over Ir(510) and Ir(110) surfaces [86].
Zhang and He [73] investigated the reaction mechanisms over Ag/Al2O3 based on in situ DRIFTS studies and found that at low temperatures (<140 °C), NH3 oxidation follows the -NH (imide) mechanism (Ag0 as the main active species), while above 140 °C, NH3 oxidation follows an in situ selective catalytic reduction (i-SCR) mechanism (Ag+ as the main active species). Furthermore, they claimed that NH3-SCO over Ag/nano-Al2O3 follows a reaction pathway called the N2 mechanism (based on in situ DRIFTS, kinetic measurements, and DFT calculation results) [17]. The intermediate N2 species appear from the combination of two -NH2/NH species (considered to be the rate-determining step). In the next step, the N2 species are converted into N2 and/or N2O in the presence of O2.
The activity and N2 selectivity strongly depend on the loading of noble metal and can be steered into the desired direction by the introduction of transition metal. E.g., Yang et al. [79] studied Ag-Cu/Al2O3 with 5–5 or 10–10 wt.% of metal and indicated the material with the first composition as a highly efficient catalyst (full NH3 conversion below 320 °C with N2 selectivity of more than 95%). Unfortunately, the authors did not present results of catalytic tests above 350 °C. Gang et al. [71] investigated (7.5 wt.%)Ag-(2.5 wt.%)Cu/Al2O3 and found the full conversion of ammonia at 200–300 °C with 95% N2 selectivity. Above 300 °C appeared significant amounts of by-products—NO and N2O. A mechanical mixture of (10 wt.%)Ag/Al2O3 and (10 wt.%)Cu/Al2O3—applied for comparative purposes, showed comparable activity and N2 selectivity to a silver-based catalyst. The same oxidation state for bimetallic (Ag-Cu/Al2O3 (5–5 wt.%, 7.5–2.5 wt.%) and monometallic (10 wt.% Ag/Al2O3 or 10 wt.% Cu/Al2O3) catalysts were approved (based on XPS analysis). Additionally, LEIS analysis over (10–2.5 wt.%, 2.5–7.5 wt.%, 5–5 wt.%, 9–1 wt.%) Ag-Cu/Al2O3 excluded formation of any Ag-Cu phases. Jabłońska et al. [80] found among all tested combinations—1–1, 1–10, 1.5–10, 5–5 wt.% of silver and copper, respectively, the (1.5 wt.%)Ag-(10 wt.%)Cu/Al2O3 catalyst with an optimum activity, N2 selectivity (full ammonia conversion and 94% N2 selectivity at 375 °C) and stability in NH3-SCO under wet conditions and time-on-stream tests. (0.59–2.34 wt.%)Ag-promoted CuMgAlOx hydrotalcite derived mixed metal oxides [87] with noble metal deposited inside the structure revealed relatively low NH3 conversion below 350 °C. Silver loading of 2.34 wt.% (n(Ag)/n(Cu)/n(Mg)/n(Al) = 1/5/65/29) led to the formation of CuOx and Ag2O—that caused higher catalytic activity and the observed drop in N2 selectivity. Significantly higher catalytic activity was reported for the Ag-Cu alloy nanoparticles (Figure 5a–d) synthesized by a solventless mix-bake-wash method. Ag-Cu (n(Ag)/n(Cu = 2/1, 77.25 wt.% of Ag) revealed full NH3 conversion at 200 °C. The AgCu alloy structure maintains the metallic state of Ag and Cu as well as structure stability, which enhanced activity and thermal stability in NH3 oxidation. The calcination of precursors of noble metal and transition metal did not form the alloy structure (Figure 5e–h). Besides, above-mentioned catalyst architectures, the (7.2 wt.%)Ag-(12.2 wt.%)Cu species were deposited onto wire-mesh honeycomb (WMH; characterized by open frontal area: 74.1%; geometric surface area: 16.2 cm2 cm−3; pressure drop: 2.58 × 10−2 Pa) [88]. Such catalyst revealed enhanced N2 selectivity (above 89% at 180–320 °C) compared to Ag/WHM as well as stability in the presence of H2 and CH4.
Concluding this part, Ag-based catalysts (especially Ag/Al2O3) received extensive concerns in NH3-SCO. As can be seen from the above examples, the research focuses mainly on the influence of the valance state of Ag species and particle size on the catalytic properties. As mentioned above, (10 wt.%)Ag/Al2O3 was suited for full NH3 conversion at about 150–400 °C with 45–95% N2 selectivity. Based on the presented above studies, the dispersed Ag0 with an average particle size in the range between 3.5–6.0 nm was found as the active species for NH3-SCO below 200 °C. On the other hand, there are limited studies that discuss the stability (especially concerning oxidation state) of the Ag-based catalysts, i.e., in the consecutive reaction cycles or the presence of H2O, SOx, and COx. Furthermore, despite their high potential, only a few studies addressed Ag-containing bifunctional catalysts for NH3-SCO. A rather high content of Ag species (compared to Pt-based bifunctional catalysts, e.g., (7.5 wt.%)Ag-(2.5 wt.%)Cu/Al2O3 or (7.2 wt.%)Ag-(12.2 wt.%)Cu/WHM) was necessary to reach high activity and N2 selectivity (i.e., full NH3 conversion with 81–95% N2 selectivity at 200–320 °C, according to data gathered in Table 2). The catalysts containing lower content of Ag species, i.e., 0.59–1.5 wt.% required a higher temperature of 375–500 °C to fully oxidize NH3. Another important aspect of NH3-SCO over Ag-based catalysts is the investigation of the reaction mechanisms, which were explored mainly by the application of in situ DRIFTS and the indication of the characteristic intermediates of the imide, hydrazine, or i-SCR (internal) mechanism. Overall, FT-IR investigations suggest that NH3-SCO may follow different parallel pathways. Thus, the combination of the advantages of in situ DRIFTS with the advantages of other (e.g., temperature-programmed and/or transient) techniques will be strategic to clarify the pathways of NH3-SCO. Furthermore, the exploration of the reaction mechanisms should be ongoing in the realistic catalytic mixture containing besides NH3 and O2 in the inert gas also H2O, SOx, and COx.
Table 2. Comparison of full NH3 conversion and N2 selectivity in same temperature range over Ag-based catalysts reported.
Table 2. Comparison of full NH3 conversion and N2 selectivity in same temperature range over Ag-based catalysts reported.
CatalystCatalyst PreparationReaction ConditionsNH3 Conversion_N2 Selectivity/%
(Temperature/°C)
Ref.
Ag powderAg2O, triple reduction, 400 °C, H2/He; oxidation, 400 °C, O2/He0.1 vol.% NH3, 10 vol.% O2, 50 mL min−1, mass of the catalyst: 0.1 g, WHSV 30,000 mL g−1 h−1100/33–77% (185–400 °C)[72]
(10 wt.%)Ag/Al2O3 impregnation, calcination, 500 °C, air; reduction, 400 °C, H2/He0.1 vol.% NH3, 10 vol.% O2, 50 mL min−1, mass of the catalyst: 0.1 g, WHSV 30,000 mL h−1 g−1 100/70–88% (160–400 °C)[72]
(10 wt.%)Ag/Al2O3 impregnation, calcination, 500 °C, air; *reduction, 400 °C, H2/N20.05 vol.% NH3, 10 vol.% O2, N2 balance, 100 mL min−1, GHSV 28,000 h−1100/93–95% (180 °C)
*100/80–82% (160–180 °C)
[76]
(10 wt.%)Ag/Al2O3 calcination, 600 °C, air; reduction, 400 °C, H2/N2
impregnation
0.05 vol.% NH3, 10 vol.% O2, N2 balance, 200 mL min−1, mass of the catalyst: 0.2 g, WHSV 1000 mL h−1 g−1 100/45–55% (150–200 °C)[74]
incipient wetness impregnation100/55–60% (150–200 °C)
sol-gel100/96% (300 °C)
(10 wt.%)Ag/Al2O3impregnation, calcination, 600 °C, air; reduction, 300 °C, H2/N20.1 vol.% NH3, 10 vol.% O2, N2 balance, 400 mL min−1, mass of the catalyst: 0.4 g, GHSV 50,000 h−1100/89 (180–260 °C)[78]
(5 wt.%)Ag/Al2O3 impregnation, calcination, 600 °C, air 0.5 vol.% NH3, 2.5 vol.% O2, Ar balance, 40 mL min−1, mass of the catalyst: 0.1 g, WHSV 24,000 mL h−1 g−1100/58–83% (275–500 °C)[80]
(10 wt.%)Ag/Al2O3 impregnation, calcination, 600 °C, air1 vol.% NH3, 10 vol.% O2, He balance, 400 mL min−1, mass of the catalyst: 0.8 g, WHSV 30,000 mL h−1 g−1100/70–83% (200–250 °C)[79]
(10 wt.%)Ag/Al2O3 impregnation, calcination, 500 °C, air1.14 vol.% NH3, 8.21 vol.% O2, 74.7 mL min−1, mass of the catalyst: 0.2 g, WHSV 22,410 mL h−1 g−1100/83% (300 °C)[71]
(1.5 wt.%)Ag/Al2O3impregnation, calcination, 600 °C, air0.5 vol.% NH3, 2.5 vol.% O2, Ar balance, 40 mL min−1, mass of the catalyst: 0.1 g,
WHSV 24,000 mL h−1 g−1
100/78% (325 °C)[82]
(10 wt.%)Ag/Al2O3100/94% (225 °C)
(2.2 wt.%)Ag/Al2O3homogenous deposition precipitation, reduction, 400 °C, H22 vol.% NH3, 2 vol.% O2, Ar balance, 40 mL min−1, GHSV 2,500 h−1*0.15 vol.% NH3, 3.85 vol.% O2, Ar balance, 40 mL min−1, GHSV 2,500 h−1100/100% (368–400 °C)
*100/85–100% (342–400 °C)
[77]
(1.6 wt.%)Ag/CeOx/Li2O/Al2O3100/95–100% (275–400 °C)
(10 wt.%)Ag/micro-Al2O3impregnation, calcination, 500 °C, air0.05 vol.% NH3, 10 vol.% O2, He balance, 100 mL min−1, GHSV 28,000–115,000 h−1
*0.05 vol.% NH3, 10 vol.% O2, 5 vol.% H2O, He balance, 100 mL min−1, GHSV 136,000 h−1
100/94–96% (160–180 °C)[17]
(10 wt.%)Ag/nano-Al2O3 100/66–76% (120–180 °C)
*100/74–90% (250–400 °C)
(9.8 wt.%)Ag/Al2O3rotary evaporator, calcination, 500 °C, air; reduction, 400 °C, H2/Ar0.1 vol.% NH3, 10 vol.% O2, He balance, 100 mL min−1, GHSV 35,000 h−1100/70–74% (140–190 °C)[83]
(9.9 wt.%)Ag/ZSM-5 100/82–89% (150–190 °C)
(10 wt.%)Ag/SiO2 impregnation, calcination, 500 °C, air; reduction, 400 °C, H2/He0.1 vol.% NH3, 10 vol.% O2, 50 mL min−1, mass of the catalyst: 0.1 g, WHSV 30,000 mL h−1 g−1100/40–73% (200–400 °C)[72]
(10 wt.%)Ag/SiO2impregnation, calcination, 600 °C, air; reduction, 300 °C, H2/N20.1 vol.% NH3, 10 vol.% O2, N2 balance, 400 mL min−1, mass of the catalyst: 0.4 g, GHSV 50,000 h−1100/50–63% (220–260 °C)[78]
(10 wt.%)Ag/TiO2100/64% (260 °C)
(10 wt.%)Ag/SiO2impregnation, calcination, 500 °C, air0.05 vol.% NH3, 10 vol.% O2, N2 balance, 100 mL min−1, GHSV 28,000 h−1100/60–62% (180–240 °C)[81]
(10 wt.%)Ag/TiO2100/91–99% (180–240 °C)
(10 wt.%)Ag/SiO2-TiO2100/60–70% (140–240 °C)
(1.5 wt.%)Ag/mesoTiO2impregnation, calcination, 600 °C, air; *reduction, 600 °C, H2/Ar0.5 vol.% NH3, 2.5 vol.% O2, Ar balance, 40 mL min−1, mass of the catalyst: 0.1 g, WHSV 24,000 mL h−1 g−1100/74–76% (375–400 °C)
*100/81–87% (350 °C)
[82]
(10 wt.%)Ag/mesoTiO2 100/81–87% (350–400 °C)
*100/72–78(275–400 °C)
(7.3 wt.%)Ag/MnO2multi-step process, calcination, 400 °C, air0.005 vol.% NH3, 20 vol.% O2, Ar balance
*0.005 vol.% NH3, 20 vol.% O2, 0.057 vol.% H2O, Ar balance
100 mL min−1, mass of the catalyst: 0.15 g, WHSV 40,000 mL h−1 g−1
100/98–99% (90–120 °C)
*100/95–96% (115–130 °C)
[90]
(10 wt.%)Ag-Y impregnation, calcination, 600 °C, air; reduction, 300 °C, H2/N20.1 vol.% NH3, 10 vol.% O2, N2 balance, 400 mL min−1, mass of the catalyst: 0.4 g, GHSV 50,000 h−1100/32–50% (220–260 °C)[78]
(21 wt.%)Ag-Yion-exchange, calcination, 400–500 °C, air 0.5 vol.% NH3, 2.5 vol.% O2, He balance
*0.5 vol.% NH3, 2.5 vol.% O2, 3.2 vol.% H2O, He balance
**0.5 vol.% NH3, 2.5 vol.% O2, 4.8 vol.% CO2, He balance
40 mL min−1, mass of the catalyst: 0.05 g, WHSV 48,000 mL h−1 g−1
100/90% (300 °C)[91]
(21 wt.%)Ag-USY 100/92–95% (200–300 °C)
*100/98% (200–300 °C)
**100/90–95% (200–300 °C)
(33 wt.%)Ag-Yion-exchange, reduction, 400 °C, H20.05 vol.% NH3, 7 vol.% O2, N2 balance, 800 mL min−1, mass of the catalyst: 0.25 g, WHSV 192,000 mL h−1 g−1100/70–80% (300–400 °C)[92]
(5 wt.%)Ag-(5 wt.%)Cu/Al2O3impregnation, calcination, 600 °C, air1 vol.% NH3, 10 vol.% O2, He balance, 400 mL min−1, mass of the catalyst: 0.8 g, WHSV 30,000 mL h−1 g−1100/95% (320 °C)[79]
(7.5 wt.%)Ag-(2.5 wt.%)Cu/Al2O3impregnation, calcination, 500 °C, air0.1 vol.% NH3, 10 vol.% O2, He balance, 50 mL min−1, mass of the catalyst: 0.1 g, WHSV 30,000 mL h−1 g−1100/95% (200–300 °C)[71]
(10 wt.%)Ag/Al2O3+(10 wt.%)Cu/Al2O3
(mixture 3:1)
impregnation, calcination, 500 °C, air1.14 vol.% NH3, 8.21 vol.% O2, 74.7 mL min−1, mass of the catalyst: 0.2 g, WHSV 22,410 mL h−1 g−1100/82% (300 °C)[71]
(1.5 wt.%)Ag-(10 wt.%)Cu/Al2O3impregnation, calcination, 600 °C, air0.5 vol.% NH3, 2.5 vol.% O2, Ar balance
*0.5 vol.% NH3, 2.5 vol.% O2, 3.2 vol.% H2O, Ar balance
40 mL min−1, mass of the catalyst: 0.1 g, WHSV 24,000 mL h−1 g−1
100/83–94% (375–500 °C)
*100/83–94% (375–500 °C)
[80]
(0.59 wt.%)AgCuMgAlOxcoprecipitation, calcination, 600 °C, air0.5 vol.% NH3, 2.5 vol.% O2, Ar balance, 40 mL min−1, mass of the catalyst: 0.1 g, WHSV 24,000 mL h−1 g−1 100/88% (425–500 °C)[87]
(2.34 wt.%)AgCuMgAlOx100/78–92% (400–500 °C)
(77.25 wt.%)Ag-Cu nanoalloysolventless mix-bake-wash method, 300 °C, air0.1 vol.% NH3, 10 vol.% O2, N2 balance, 100 mL min−1, GHSV 12,000 h−1 100/72–85% (210–240 °C)[89]
(67.57 wt.%)AgCuOxcalcination, 500 °C, air100/64–72% (300–340 °C)
(7.4 wt.%)Ag/WMHWMH—wire-mesh honeycomb; impregnation, calcination, 500 °C, air0.1 vol.% NH3, 10 vol.% O2, He balance, 300 mL min−1, GHSV 2,250 h−1 100/63–73% (210–320 °C)[88]
(7.2 wt.%)Ag-(12.2 wt.%)Cu/WMH100/81–89% (220–320 °C)

4. Au- and Ru-Based Catalysts

Gold-based catalysts are well-known for their high catalytic activity at low temperatures, e.g., in the oxidation of CO [93,94]. However, such catalysts were rarely investigated in NH3-SCO (Table 3). Interestingly, Lin et al. [95] investigated NH3 oxidation over the in situ H2-reduced (5 wt.%)Au/MOx/Al2O3 (M = Cu, Fe, Ce, Li, and Ti) catalysts in the temperature range from 200 to 400 °C. Among all investigated catalysts, Au/Cu-Al2O3 was the most active and N2 selective (full NH3 conversion at 300 °C with 95% N2 selectivity). The H2-reduced Au/Al2O3 catalyst revealed significantly lower NH3 conversion (ca. 30% above 400 °C). However, the NH3 conversion increases with an increasing O2 in the feed (c(NH3):c(O2) = 1:10; NH3 conversion of 45% at 400 °C) [77]. Gong et al. [96] and Liu et al. [97] reported from experimental and theoretical studies that NH3 did not dissociate on the Au(111) surface until it was precovered with oxygen atoms or hydroxyl groups. Thus, Lippits et al. [77] observed the enhanced NH3 conversion (full NH3 conversion at 338 °C with N2 selectivity below 50%) over Au/Al2O3 after its doping with CeOx (able to provide and store oxygen) and Li2O (responsible for decreasing the catalyst acidity and thus improved oxygen adsorption). Recently, Lin et al. [12] reported the acidic metal-oxide-supported gold catalyst (Au/Nb2O5) with improved N2 selectivity compared to other metal-oxide-supported gold catalysts (e.g., Au/SiO2, Au/Al2O3, Au/Fe2O3, etc., Figure 6a). Specifically, Au/Nb2O5 contains both Brønsted and Lewis acid sites that allowed NH3 oxidation according to hydrazine mechanism (N2H4 as the intermediate) and imide mechanism (HNO as the intermediate), respectively (Figure 6b). Overall, further detailed property-activity studies over Au-containing materials could path the way for further catalysts design and their optimization.
Some studies were carried out on RuO2(110) surface characterized by two types of atoms with unsaturated bonds along [1] direction: a) the twofold coordinated oxygen atoms (O-bridge; O-br) and b) the fivefold coordinated Ru atoms (Ru-cus; the adsorption site for ammonia) [98]. NH3 decomposes (to NH2) at −183 °C, while successive annealing to −23–27 °C produces N [98,99]. N2 is predominantly formed over polycrystalline RuO2 in a direct combination of Ru-coordinated N atoms (at ambient pressure, 6% of NO at c(O2):c(NH3) = 2:1; 65% of NO at c(O2):c(NH3) = 140:1) [100]. The selectivity to NO increases with increasing temperature (100% around 257 °C, in UHV, p(NH3) = 10−7 mbar, and c(O2):c(NH3) = 20:1) because of the high desorption temperature for NO (227 °C). At lower temperatures, NO-formation is hindered by surface water molecules [98,99]. Seitsonen et al. [99] estimated energy barriers to the elementary H-abstraction steps and the recombination of N and O atoms on RuO2(110) surface by using DFT calculations and high-resolution core-level shift spectroscopy (Figure 7). The high activity of RuO2(110) arose from low activation energies from the successive H-abstraction.
Also, RuO2-supported catalysts are active for ammonia oxidation (Table 3). E.g., Cui et al. investigated the RuO2-CuO/Al-ZrO2 [101] and CuO/RuO2 [102] catalysts with 5–30 wt.% and 70–95 wt.% of Ru loading, respectively. The catalysts possessed excellent activity and N2 selectivity at low temperatures, i.e., for (20 wt.%)RuO2-CuO/Al-ZrO2 full NH3 conversion at 195 °C with 100% N2 selectivity [101], or for (10 wt.%)CuO/RuO2 full NH3 conversion at 180 °C and N2 selectivity above 95% [102] were achieved. However, these catalysts are relatively expensive and therefore their commercialization is hindered. Thus, Chakrobaty et al. [103] investigated the Cu/Ru catalysts with varying overlaying thickness of Cu film (with the optimum of 0.8 monolayers) deposited by physical vapor deposition on (5 nm)Ru/TiO2 (111). The synergistic interaction between Cu and Ru species led to a threefold higher ammonia conversion rate than was achieved over Ru-based catalyst. Concerning the powder materials, Wang et al. [104] studied a series of WO3-modified RuO2-Fe2O3 catalysts with a lower cost, i.e., with 1 wt.% of ruthenium. The introduction of 5 wt.% of WO3 (among 1–9 wt.% of WO3) tuned the surface acidity, and thus, enhanced activity and N2 selectivity of RuO2-Fe2O3 (full NH3 conversion at 250–400 °C and 93–97% N2 selectivity). In situ DRIFTS results indicated that NH3-SCO over (1 wt.%)RuO2-(5 wt.%)WO3-Fe2O3 proceeds according to the i-SCR mechanism. The -NH2 intermediate reacted with the in situ-generated NOx ad-species with the formation of N2. Furthermore, Chen et al. [105,106] studied (0.2 wt.%)Ru/Ce0.6Zr0.4O2(PVP) or (0.2 wt.%)IrO2/Ce0.6Zr0.4O2(PVP) (PVP, polyvinylpyrrolidone) and claimed that -HNO appeared as an intermediate in the i-SCR mechanism of NH3-SCO (Figure 8). The formed -HNO interacted with atomic oxygen with the formation of NO, which furthermore reacts with -NHx (-NH2 and -NH) species with the formation of N2 and N2O (minor by-product). The presence of SO2 in the feed gas effectively inhibits the production of N2O, i.e., the reactions between gaseous NO and -NH2 will be enhanced (more adsorbed ammonia on the sulfated (acidic) surface). SO2 can also inhibit NH3 oxidation resulting in higher N2 selectivity (up to 100%) in the absence of NOx. Similar conclusions were given for RuOx/TiO2-SO42−, however, the time of the sulfated treatment (0.5–6 h) of the support varied activity and N2 selectivity of the final catalysts (with an optimum at 2 h) [107].
Concluding this part, both Au- and Ru-based catalysts were significantly less investigated in NH3-SCO. A highly loaded Ru-containing materials (10–20 wt.% Ru; (10 wt.%)CuO-RuO2 or (20 wt.%)RuO2-CuO/Al-ZrO2) present a class of highly active and N2 selective catalysts at relatively low temperatures (i.e., full NH3 conversion at 180–350 °C with 95–100% N2 selectivity; according to data gathered in Table 3). Otherwise, the materials with significantly lower content of Ru species (0.5–3 wt.%) were less active (full NH3 conversion at 175–400 °C) and N2 selective (43–99%). However, again, concerning the influence of the preparation variables (e.g., the different total amount of ruthenium, variety of applied metal promoters and supports) as well as pretreatment and reaction conditions, the comparison of activity and N2 selectivity over Ru-based catalysts each other or even with other noble metal-based catalysts is limited. Furthermore, for both Au- and Ru-based catalysts, oxidized metal species (i.e., Au+/Au3+, Ru4+) ensure enhanced activity and N2 selectivity (in contrast to the Pt- or Ag-based catalysts). Nevertheless, an in-depth understanding of the role of active species in NH3-SCO is still lacking and needs to be demonstrated clearly in further studies (especially over Au-based catalysts). Furthermore, the presented catalytic systems were mainly investigated under ideal conditions (only NH3 and O2 diluted in inert gas) also concerning the investigation of the reaction mechanisms, i.e., through in situ DRIFTS experiments (NH3 adsorption/desorption in inert gas or oxygen). Otherwise, surface reactions are fast (residence time in a range of seconds) and the reaction mechanism involves a series of parallel and consecutive reactions. Thus, the reaction intermediates and conversion of the substrate molecules on the catalyst surface should be followed with more detailed ex situ, in situ, and operando spectroscopic studies as well as transient kinetic investigations under-applied reaction conditions.
Table 3. Comparison of full NH3 conversion and N2 selectivity in same temperature range over Au- and Ru-based catalysts reported in literature.
Table 3. Comparison of full NH3 conversion and N2 selectivity in same temperature range over Au- and Ru-based catalysts reported in literature.
CatalystCatalyst PreparationReaction ConditionsNH3 Conversion_N2 Selectivity/%
(Temperature/°C)
Ref.
(5 wt.%)Au/CuO/Al2O3impregnation, calcination, 300 °C, air; reduction, 300 °C, H22 vol.% NH3, 2 vol.% O2, He balance, 30 mL min−1, mass of the catalyst: 0.15 g, WHSV 12,000 mL h−1 g−1100/95% (300 °C)[95]
(4 wt.%)Au/CeOx/Li2O/Al2O3 homogenous deposition precipitation, reduction, 400 °C, H22 vol.% NH3, 2 vol.% O2, Ar balance, 40 mL min−1, GHSV 2500 h−1 100/34–49% (338–400 °C)[77]
(10 wt.%)CuO-RuO2 conanocasting-replication method, calcination, 500 °C, air0.1 vol.% NH3, 2 vol.% O2, Ar balance, 100 mL min−1, mass of the catalyst: 0.08 g, WHSV 75,000 mL h−1 g−1100/95–97% (180–350 °C)[102]
(20 wt.%)RuO2-CuO/ZrO2impregnation, calcination, 550 °C, air0.04 vol.% NH3, 5 vol.% O2, 6 vol.% H2O, Ar balance, 200 mL min−1, mass of the catalyst: 0.1 g, WHSV 120,000 mL h−1 g−1100/68–98% (250–275 °C)[101]
(20 wt.%)RuO2/Al-ZrO2100/70–82% (248–325 °C)
(20 wt.%)RuO2-CuO/Al-ZrO2 100/100% (195–280 °C)
(1 wt.%)RuO2-Fe2O3sol-gel route, calcination, 500 °C, air0.08 vol.% NH3, 5 vol.% O2, Ar balance, 400 mL min−1, GHSV 60,000 h−1100/67–90% (250–400 °C)[104]
(1 wt.%)RuO2-WO3-Fe2O3100/93–97% (250–400 °C)
(1.13 wt.%)Ru/Cu-SSZ-13impregnation, 500 °C, air; pretreatment, 300 °C, O2/N20.05 vol.% NH3, 0.5 vol.% CO, 5 vol.% O2, N2 balance, GHSV 300,000 h−1100/94–96% (220–300 °C)[108]
(3 wt.%)RuO2/TiO2 impregnation, calcination, 400 °C, air0.1 vol.% NH3, 5 vol.% O2, 3 vol.% H2O, N2 balance, 150 mL min−1, mass of the catalyst: 0.1 g, WHSV 90,000 mL h−1 g−1100/47–72% (350–400 °C)[109]
(3 wt.% Ru)RuO2/Na-Y ion-exchange, pretreatment, 450 °C, O2/H2O/N2 100/51–99% (175–400 °C)
(1 wt.%)Ru/TiO2 impregnation, 400 °C, air; pretreatment, 400 °C, O2/N20.02 vol.% NH3, 10 vol.% O2, 6 vol.% H2O, N2 balance, 500 mL min−1, mass of the catalyst: 0.3 g, GHSV 60,000 h−1100/54–63% (275–300 °C)[110]
(0.5 wt.%)Ru/TiO2impregnation, 450 °C, air; pretreatment, 400 °C, O2/N20.08 vol.% NH3, 5 vol.% O2, N2 balance, 400 mL min−1, GHSV 60,000 h−1100/43–85% (200–400 °C)[107]
(0.5 wt.%)Ru/TiO2-SO42− 100/65–93% (225–400 °C)

5. Conclusions and Outlook

NH3-SCO is the most efficient method for ammonia removal from oxygen-containing exhausts. The number of publications related to this process successively increases with the main researchers’ interest in the development of the catalyst with high activity, N2 selectivity, and stability in the broad temperature range. The present mini-review provides a broad picture of the property-activity correlations of noble metal-based catalysts investigated for NH3-SCO. Among presented Pt-, Pd-, Ag- and Au-, Ru-based catalytic systems, mainly H2-reduced (1–2 wt.%)Pt/Al2O3 and (10 wt.%)Ag/Al2O3 were recognized as the most active NH3-SCO systems in the low temperatures (<300 °C). Unfortunately, they caused significant formation of N2O and NOx. Moreover, as can be seen from the above examples, although Pt- and Ag-based catalysts were more intensively investigated regarding their property-activity correlations—compared to that of the Pd-, Au-, and Ru-containing catalysts—there is still a lack of systematic studies concerning the nature and role of active species as well as the influence of (a) the particle size of active components and their aggregation state; (b) the catalyst supports (i.e., inorganic oxides versus zeolites); (c) the preparation methods (i.e., catalysts in the structured forms, e.g., monolith), and (d) feed composition (i.e., various c(NH3):c(O2) ratios (1:1–25—an excess of oxygen together with minor NH3 slip, presence of H2O, SOx, COx, etc.) on activity, N2 selectivity, and stability in catalysis. Contrary to the transition metal-based catalysts only a few examples concern catalytic studies over noble metal-based zeolites. Otherwise, concerning the discussed material requirements (i.e., enhanced ammonia conversion, N2 selectivity and stability in the presence of typical components of exhaust gases and the broad temperature range up to 600–700 °C (in the cycle of diesel particulate filter regeneration)), the zeolite-based catalysts present a class of highly promising materials. For instance, ion-exchanged zeolites showed higher activity also in the presence of H2O, compared to that of alumina-supported oxides with the same metal loading due to high dispersion of metal species and acid sites of high strength. Furthermore, the studies on the reaction mechanisms are rather scarce. The reaction mechanisms must be clarified to rationally develop a process for NH3 oxidation to N2 over applied catalysts. These problems highlight the importance of more detailed ex situ and in situ methods (i.e., temperature-programmed, spectroscopic, and/or transient methods) in studying the catalysts under real working environments.
Furthermore, a relatively narrow operating temperature window of full NH3 conversion (Table 1, Table 2 and Table 3), high selectivity to N2O and NOx, and high costs of noble metals motivated researchers to develop suitable bifunctional systems. Cu-ZSM-5 or even Cu-chabazite (SSZ-13, SAPO-34) are already recognized (also discussed in the previous review articles concerning NH3-SCO, Jabłońska et al., 2016, Jabłońska et al., 2020) as the active and N2 selective catalysts for NH3-SCR. Despite presented here developments in bifunctional catalysts (mainly Pt-Cu or Ag-Cu catalysts systems), challenges remain in achieving enhanced NH3 conversion, N2 selectivity, and stability (in the presence of real flue gases, such as H2O, SOx, and COx). Further investigations concerning bifunctional catalysts with a low number of noble metals, e.g., Au, Pt-Au, Pt-Rh, etc., constitute a promising research direction. Nevertheless, the present findings, indications, and thoughts given in the mini-review form a solid basis for further developments of structured catalysts (i.e., in the form of dual-layer or core-shell structure) and their optimization.

Funding

This research received no external funding.

Acknowledgments

The author acknowledges support from the German Research Foundation (DFG) and Leipzig University within the program of Open Access Publishing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Park, S.-J.; Jin, S.-Y. Effect of ozone treatment on ammonia removal of activated carbons. J. Colloid Interface Sci. 2005, 286, 417–419. [Google Scholar] [CrossRef]
  2. Abe, Y.; Miyata, N.; Yasuda, T. Comparison between direct-dontact HfO2/Ge and HfO2/GeO2/Ge stack structures: Physical and electrical properties. ECS Trans. 2008, 16, 375. [Google Scholar] [CrossRef]
  3. Karri, R.R.; Sahu, J.N.; Chimmiri, V. Critical review of abatement of ammonia from wastewater. J. Mol. Liq. 2018, 261, 21–31. [Google Scholar] [CrossRef]
  4. Boardman, G.D.; Starbuck, S.M.; Hudgins, D.B.; Li, X.; Kuhn, D.D. Toxicity of ammonia to three marine fish and three marine invertebrates. Environ. Toxicol. Int. J. 2004, 19, 134–142. [Google Scholar] [CrossRef] [PubMed]
  5. Camargo, J.A.; Alonso, Á. Ecological and toxicological effects of inorganic nitrogen pollution in aquatic ecosystems: A global assessment. Environ. Int. 2006, 32, 831–849. [Google Scholar] [CrossRef] [PubMed]
  6. Dammers, E.; McLinden, C.A.; Griffin, D.; Shephard, M.W.; Graaf, S.V.D.; Lutsch, E.; Schaap, M.; Gainairu-Matz, Y.; Fioletov, V.; Damme, M.V.; et al. NH3 emissions from large point sources derived from CrIS and IASI satellite observations. Atmos. Chem. Phys. 2019, 19, 12261–12293. [Google Scholar] [CrossRef] [Green Version]
  7. Jabłońska, M.; Palkovits, R. Copper based catalysts for the selective ammonia oxidation into nitrogen and water vapour-Recent trends and open challenges. Appl. Catal. B Environ. 2016, 181, 332–351. [Google Scholar] [CrossRef]
  8. European Union. Regulation (EC) No 595/2009 of the European Parliament and of the Council of 18 June 2009 on Type-Approval of Motor Vehicles and Engines with Respect to Emissions from Heavy Duty Vehicles (Euro VI) and on Access to Vehicle Repair and Maintenance Informati. Off. J. Eur. Communities 50 2009, 43, 164–176. [Google Scholar]
  9. Suarez-Bertoa, R.; Mendoza-Villafuerte, P.; Riccobono, F.; Vojtisek, M.; Pechout, M.; Perujo, A.; Astorga, C. On-road measurement of NH3 emissions from gasoline and diesel passenger cars during real world driving conditions. Atmos. Environ. 2017, 166, 488–497. [Google Scholar] [CrossRef]
  10. Suarez-Bertoa, R.; Pechout, M.; Vojtíšek, M.; Astorga, C. Regulated and non-regulated emissions from Euro 6 diesel, gasoline and CNG vehicles under real-world driving conditions. Atmosphere 2020, 11, 204. [Google Scholar] [CrossRef] [Green Version]
  11. Carabineiro, S.A.C.; Matveev, A.V.; Gorodetskii, V.V.; Nieuwenhuys, B.E. Selective oxidation of ammonia over Ru (0 0 0 1). Surf. Sci. 2004, 555, 83–93. [Google Scholar] [CrossRef]
  12. Lin, M.; An, B.; Niimi, N.; Jikihara, Y.; Nakayama, T.; Honma, T.; Takei, T.; Shishido, T.; Ishida, T.; Haruta, M.; et al. Role of the acid site for selective catalytic oxidation of NH3 over Au/Nb2O5. ACS Catal. 2019, 9, 1753–1756. [Google Scholar] [CrossRef]
  13. Chmielarz, L.; Jabłońska, M.; Strumiński, A.; Piwowarska, Z.; Węgrzyn, A.; Witkowski, S.; Michalik, M. Selective catalytic oxidation of ammonia to nitrogen over Mg-Al, Cu-Mg-Al and Fe-Mg-Al mixed metal oxides doped with noble metals. Appl. Catal. B Environ. 2013, 130, 152–162. [Google Scholar] [CrossRef]
  14. Chmielarz, L.; Jabłońska, M. Advances in selective catalytic oxidation of ammonia to dinitrogen: A review. RSC Adv. 2015, 5, 43408–43431. [Google Scholar] [CrossRef]
  15. Gao, F.; Liu, Y.; Sani, Z.; Tang, X.; Yi, H.; Zhao, S.; Yu, Q.; Zhou, Y. Advances in selective catalytic oxidation of ammonia (NH3-SCO) to dinitrogen in excess oxygen: A review on typical catalysts, catalytic performances and reaction mechanisms. J. Environ. Chem. Eng. 2020, 9, 104575. [Google Scholar]
  16. Lan, T.; Zhao, Y.; Deng, J.; Zhang, J.; Shi, L.; Zhang, D. Selective catalytic oxidation of NH3 over noble metal-based catalysts: State of the art and future prospects. Catal. Sci. Technol. 2020, 10, 5792–5810. [Google Scholar] [CrossRef]
  17. Wang, F.; Ma, J.; He, G.; Chen, M.; Zhang, C.; He, H. Nanosize effect of Al2O3 in Ag/Al2O3 catalyst for the selective catalytic oxidation of ammonia. ACS Catal. 2018, 8, 2670–2682. [Google Scholar] [CrossRef]
  18. Li, P.; Zhang, R.; Liu, N.; Royer, S. Efficiency of Cu and Pd substitution in Fe-based perovskites to promote N2 formation during NH3 selective catalytic oxidation (NH3-SCO). Appl. Catal. B Environ. 2017, 203, 174–188. [Google Scholar] [CrossRef]
  19. Jabłońska, M. Progress on selective catalytic ammonia oxidation (NH3-SCO) over Cu-containing zeolite-based catalysts. ChemCatChem 2020, 12, 4490–4500. [Google Scholar] [CrossRef]
  20. Il’chenko, N.I. Catalytic oxidation of ammonia. Russ. Chem. Rev. 1976, 45, 1119. [Google Scholar] [CrossRef]
  21. Zeng, Y.F.; Imbihl, R. Structure sensitivity of ammonia oxidation over platinum. J. Catal. 2009, 261, 129–136. [Google Scholar] [CrossRef]
  22. Novell-Leruth, G.; Valcarcel, A.; Pérez-Ramírez, J.; Ricart, J.M. Ammonia dehydrogenation over platinum-group metal surfaces. Structure, stability, and reactivity of adsorbed NHx species. J. Phys. Chem. C 2007, 111, 860–868. [Google Scholar] [CrossRef]
  23. Gland, J.L.; Korchak, V.N. Ammonia oxidation on a stepped platinum single-crystal surface. J. Catal. 1978, 53, 9–23. [Google Scholar] [CrossRef]
  24. Baerns, M.; Imbihl, R.; Kondratenko, V.A.; Kraehnert, R.; Offermans, W.K.; Van Santen, R.A.; Scheibe, A. Bridging the pressure and material gap in the catalytic ammonia oxidation: Structural and catalytic properties of different platinum catalysts. J. Catal. 2005, 232, 226–238. [Google Scholar] [CrossRef]
  25. Pérez-Ramírez, J.; Kondratenko, E.V.; Novell-Leruth, G.; Ricart, J.M. Mechanism of ammonia oxidation over PGM (Pt, Pd, Rh) wires by temporal analysis of products and density functional theory. J. Catal. 2009, 261, 217–223. [Google Scholar] [CrossRef]
  26. Ostermaier, J.J.; Katzer, J.R.; Manogue, W.H. Crystallite size effects in the low-temperature oxidation of ammonia over supported platinum. J. Catal. 1974, 33, 457–473. [Google Scholar] [CrossRef]
  27. Svintsitskiy, D.A.; Slavinskaya, E.M.; Stonkus, O.A.; Romanenko, A.V.; Stadnichenko, A.I.; Kibis, L.S.; Derevyannikova, E.A.; Evtushkova, A.A.; Boronin, A.I. The state of platinum and structural features of Pt/Al2O3 catalysts in the reaction of NH3 oxidation. J. Struct. Chem. 2019, 60, 919–931. [Google Scholar] [CrossRef]
  28. Bahrami, B.; Komvokis, V.G.; Ziebarth, M.S.; Alexeev, O.S.; Amiridis, M.D. NH3 decomposition and oxidation over noble metal-based FCC CO combustion promoters. Appl. Catal. B Environ. 2013, 130, 25–35. [Google Scholar] [CrossRef]
  29. Svintsitskiy, D.A.; Kibis, L.S.; Stadnichenko, A.I.; Slavinskaya, E.M.; Romanenko, A.V.; Fedorova, E.A.; Stonkus, O.A.; Doronkin, D.E.; Marchuk, V.; Zimina, A.; et al. Insight into the nature of active species of Pt/Al2O3 catalysts for low temperature NH3 oxidation. ChemCatChem 2020, 12, 867–880. [Google Scholar] [CrossRef] [Green Version]
  30. Kibis, L.S.; Svintsitskiy, D.A.; Stadnichenko, A.I.; Slavinskaya, E.M.; Romanenko, A.V.; Fedorova, E.A.; Stonkus, O.A.; Svetlichnyi, V.A.; Fakhrutdinova, E.D.; Vorokhta, M.; et al. In situ probing of Pt/TiO2 activity in low-temperature ammonia oxidation. Catal. Sci. Technol. 2021, 11, 250–263. [Google Scholar] [CrossRef]
  31. Liu, J.; Lin, Q.; Liu, S.; Xu, S.; Xu, H.; Chen, Y. Promotional effects of ascorbic acid on the low-temperature catalytic activity of selective catalytic oxidation of ammonia over Pt/SA: Effect of Pt0 content. New J. Chem. 2020, 44, 4108–4113. [Google Scholar] [CrossRef]
  32. Ostermaier, J.J.; Katzer, J.R.; Manogue, W.H. Platinum catalyst deactivation in low-temperature ammonia oxidation reactions: I. Oxidation of ammonia by molecular oxygen. J. Catal. 1976, 41, 277–292. [Google Scholar] [CrossRef]
  33. Sobczyk, D.P.; Van Grondelle, J.; Thüne, P.C.; Kieft, I.E.; De Jong, A.M.; Van Santen, R.A. Low-temperature ammonia oxidation on platinum sponge studied with positron emission profiling. J. Catal. 2004, 225, 466–478. [Google Scholar] [CrossRef]
  34. Sobczyk, D.P.; Hensen, E.J.M.; De Jong, A.M.; Van Santen, R.A. Low-temperature ammonia oxidation over Pt/γ-alumina: The influence of the alumina support. Top. Catal. 2003, 23, 109–117. [Google Scholar] [CrossRef]
  35. Liu, J.; Sun, M.; Lin, Q.; Liu, S.; Xu, H.; Chen, Y. Promotional effects of ethylenediamine on the low-temperature catalytic activity of selective catalytic oxidation of ammonia over Pt/SiAlOx: States and particle sizes of Pt. Appl. Surf. Sci. 2019, 481, 1344–1351. [Google Scholar] [CrossRef]
  36. Slavinskaya, E.M.; Kibis, L.S.; Stonkus, O.A.; Svintsitskiy, D.A.; Stadnichenko, A.I.; Fedorova, E.A.; Romanenko, A.V.; Marchuk, V.; Doronkin, D.E.; Boronin, A.I. The effects of platinum dispersion and Pt state on catalytic properties of Pt/Al2O3 in NH3 oxidation. ChemCatChem 2020, 13, 313–327. [Google Scholar] [CrossRef]
  37. Dhillon, P.S.; Harold, M.P.; Wang, D.; Kumar, A.; Joshi, S. Hydrothermal aging of Pt/Al2O3 monolith: Washcoat morphology degradation effects studied using ammonia and propylene oxidation. Catal. Today 2019, 320, 20–29. [Google Scholar] [CrossRef]
  38. Machida, M.; Tokudome, Y.; Maeda, A.; Kuzuhara, Y.; Hirakawa, T.; Sato, T.; Yoshida, H.; Ohyama, J.; Fujii, K.; Ishikawa, N. Nanometric platinum overlayer to catalyze NH3 oxidation with high turnover frequency. ACS Catal. 2020, 10, 4677–4685. [Google Scholar] [CrossRef]
  39. Mieher, W.D.; Ho, W. Thermally activated oxidation of NHs on Pt (111): Intermediate species and reaction mechanisms. Surf. Sci. 1995, 322, 151–167. [Google Scholar] [CrossRef]
  40. Offermans, W.K.; Jansen, A.P.J.; Van Santen, R.A. Ammonia activation on platinum {111}: A density functional theory study. Surf. Sci. 2006, 600, 1714–1734. [Google Scholar] [CrossRef]
  41. Novell-Leruth, G.; Ricart, J.M.; Pérez-Ramírez, J. Pt (100)-catalyzed ammonia oxidation studied by DFT: Mechanism and microkinetics. J. Phys. Chem. C 2008, 112, 13554–13562. [Google Scholar] [CrossRef]
  42. Sun, M.; Liu, J.; Song, C.; Ogata, Y.; Rao, H.; Zhao, X.; Xu, H.; Chen, Y. Different reaction mechanisms of ammonia oxidation reaction on Pt/Al2O3 and Pt/CeZrO2 with various Pt states. ACS Appl. Mater. Interfaces 2019, 11, 23102–23111. [Google Scholar] [CrossRef] [PubMed]
  43. Li, Y.; Armor, J.N. Selective NH3 oxidation to N2 in a wet stream. Appl. Catal. B Environ. 1997, 13, 131–139. [Google Scholar] [CrossRef]
  44. Long, R.Q.; Yang, R.T. Superior ion-exchanged ZSM-5 catalysts for selective catalytic oxidation of ammonia to nitrogen. Chem. Commun. 2000, 17, 1651–1652. [Google Scholar] [CrossRef]
  45. Jabłońska, M.; Król, A.; Kukulska-Zając, E.; Tarach, K.; Chmielarz, L.; Góra-Marek, K. Zeolite Y modified with palladium as effective catalyst for selective catalytic oxidation of ammonia to nitrogen. J. Catal. 2014, 316, 36–46. [Google Scholar] [CrossRef]
  46. Decarolis, D.; Clark, A.H.; Pellegrinelli, T.; Nachtegaal, M.; Lynch, E.W.; Catlow, C.R.A.; Gibson, E.K.; Goguet, A.; Wells, P.P. Spatial profiling of a Pd/Al2O3 Cctalyst during selective ammonia oxidation. ACS Catal. 2021, 11, 2141–2149. [Google Scholar] [CrossRef]
  47. Dann, E.K.; Gibson, E.K.; Blackmore, R.H.; Catlow, C.R.A.; Collier, P.; Chutia, A.; Erden, T.E.; Hardacre, C.; Kroner, A.; Nachtegaal, M.; et al. Structural selectivity of supported Pd nanoparticles for catalytic NH3 oxidation resolved using combined operando spectroscopy. Nat. Catal. 2019, 2, 157–163. [Google Scholar] [CrossRef]
  48. Jabłońska, M. TPR study and catalytic performance of noble metals modified Al2O3, TiO2 and ZrO2 for low-temperature NH3-SCO. Catal. Commun. 2015, 70, 66–71. [Google Scholar] [CrossRef]
  49. Olofsson, G.; Wallenberg, L.R.; Andersson, A. Selective catalytic oxidation of ammonia to nitrogen at low temperature on Pt/CuO/Al2O3. J. Catal. 2005, 230, 1–13. [Google Scholar] [CrossRef]
  50. Burch, R.; Southward, B.W.L. Low-temperature, clean catalytic combustion of N-bearing gasified biomass using a novel NH3 trapping catalyst. Chem. Commun. 2000, 13, 1115–1116. [Google Scholar] [CrossRef]
  51. Kušar, H.M.J.; Ersson, A.G.; Vosecký, M.; Järås, S.G. Selective catalytic oxidation of NH3 to N2 for catalytic combustion of low heating value gas under lean/rich conditions. Appl. Catal. B Environ. 2005, 58, 25–32. [Google Scholar] [CrossRef]
  52. Burch, R.; Southward, B.W.L. A novel application of trapping catalysts for the selective low-temperature oxidation of NH3 to N2 in simulated biogas. J. Catal. 2000, 195, 217–226. [Google Scholar] [CrossRef]
  53. Sun, M.; Wang, S.; Li, Y.; Xu, H.; Chen, Y. Promotion of catalytic performance by adding W into Pt/ZrO2 catalyst for selective catalytic oxidation of ammonia. Appl. Surf. Sci. 2017, 402, 323–329. [Google Scholar] [CrossRef]
  54. Long, R.Q.; Yang, R.T. Noble metal (Pt, Rh, Pd) promoted Fe-ZSM-5 for selective catalytic oxidation of ammonia to N2 at low temperatures. Catal. Lett. 2002, 78, 353–357. [Google Scholar] [CrossRef]
  55. Kim, M.-S.; Lee, D.-W.; Chung, S.-H.; Hong, Y.-K.; Lee, S.H.; Oh, S.-H.; Cho, I.-H.; Lee, K.-Y. Oxidation of ammonia to nitrogen over Pt/Fe/ZSM5 catalyst: Influence of catalyst support on the low temperature activity. J. Hazard. Mater. 2012, 237, 153–160. [Google Scholar] [CrossRef]
  56. Abbas-Ghaleb, R.; Chlala, D. Selective catalytic oxidation of NH3 into N2 during biogas combustion over 2 wt% PdO/5 wt% CuO/γ-Al2O3. SN Appl. Sci. 2020, 2, 592–599. [Google Scholar] [CrossRef] [Green Version]
  57. Shrestha, S.; Harold, M.P.; Kamasamudram, K.; Yezerets, A. Ammonia oxidation on structured composite catalysts. Top. Catal. 2013, 56, 182–186. [Google Scholar] [CrossRef]
  58. Shrestha, S.; Harold, M.P.; Kamasamudram, K.; Yezerets, A. Selective oxidation of ammonia on mixed and dual-layer Fe-ZSM-5+Pt/Al2O3 monolithic catalysts. Catal. Today 2014, 231, 105–115. [Google Scholar] [CrossRef]
  59. Shrestha, S.; Harold, M.P.; Kamasamudram, K.; Kumar, A.; Olsson, L.; Leistner, K. Selective oxidation of ammonia to nitrogen on bi-functional Cu-SSZ-13 and Pt/Al2O3 monolith catalyst. Catal. Today 2016, 267, 130–144. [Google Scholar] [CrossRef]
  60. Scheuer, A.; Hauptmann, W.; Drochner, A.; Gieshoff, J.; Vogel, H.; Votsmeier, M. Dual layer automotive ammonia oxidation catalysts: Experiments and computer simulation. Appl. Catal. B Environ. 2012, 111, 445–455. [Google Scholar] [CrossRef]
  61. Colombo, M.; Nova, I.; Tronconi, E.; Schmeißer, V.; Bandl-Konrad, B.; Zimmermann, L. Experimental and modeling study of a dual-layer (SCR+ PGM) NH3 slip monolith catalyst (ASC) for automotive SCR aftertreatment systems. Part 1. Kinetics for the PGM component and analysis of SCR/PGM interactions. Appl. Catal. B Environ. 2013, 142, 861–876. [Google Scholar] [CrossRef]
  62. Dhillon, P.S.; Harold, M.P.; Wang, D.; Kumar, A.; Joshi, S.Y. Enhanced transport in washcoated monoliths: Application to selective lean NOx reduction and ammonia oxidation. Chem. Eng. J. 2019, 377, 119734. [Google Scholar] [CrossRef]
  63. Ghosh, R.S.; Le, T.T.; Terlier, T.; Rimer, J.D.; Harold, M.P.; Wang, D. Enhanced selective oxidation of ammonia in a Pt/Al2O3@ Cu/ZSM-5 core-shell catalyst. ACS Catal. 2020, 10, 3604–3617. [Google Scholar] [CrossRef]
  64. Dhillon, P.S.; Harold, M.P.; Wang, D.; Kumar, A.; Joshi, S.Y. Optimizing the dual-layer Pt/Al2O3+Cu/SSZ-13 washcoated monolith: Selective oxidation of NH3 to N2. Catal. Today 2021, 360, 426–434. [Google Scholar] [CrossRef]
  65. Hung, C.-M.; Lai, W.-L.; Lin, J.-L. Removal of gaseous ammonia in Pt-Rh binary catalytic oxidation. Aerosol Air Qual. Res. 2012, 12, 583–591. [Google Scholar] [CrossRef] [Green Version]
  66. Hung, C.-M. Fabrication, characterization, and evaluation of the cytotoxicity of platinum-rhodium nanocomposite materials for use in ammonia treatment. Powder Technol. 2011, 209, 29–34. [Google Scholar] [CrossRef]
  67. Gang, L.; Anderson, B.G.; Van Grondelle, J.; Van Santen, R.A. NH3 oxidation to nitrogen and water at low temperatures using supported transition metal catalysts. Catal. Today 2000, 61, 179–185. [Google Scholar] [CrossRef]
  68. Chang, S.; Harle, G.; Ma, J.; Yi, J. The effect of textural Properties of CeO2-SiO2 mixed oxides on NH3-SCO activity of Pt/CeO2-SiO2 catalyst. Appl. Catal. A Gen. 2020, 604, 117775. [Google Scholar] [CrossRef]
  69. Kim, G.J.; Kwon, D.W.; Shin, J.H.; Kim, K.W.; Hong, S.C. Influence of the addition of vanadium to Pt/TiO2 catalyst on the selective catalytic oxidation of NH3 to N2. Environ. Technol. 2019, 40, 2588–2600. [Google Scholar] [CrossRef] [PubMed]
  70. Il’chenko, N.I.; Golodets, G.I.; Avilova, I.M. Oxidation of ammonia on metals. [Catalytic activity of Pt, Pd, Cu, Ag, Ni, Au, Fe, W, and Ti]. Kinet. Catal. (USSR) (Engl. Transl.) (USA) 1975, 16, 1455–1460. [Google Scholar]
  71. Gang, L.; Anderson, B.G.; Van Grondelle, J.; Van Santen, R.A.; Van Gennip, W.J.H.; Niemantsverdriet, J.W.; Kooyman, P.J.; Knoester, A.; Brongersma, H.H. Alumina-supported Cu-Ag catalysts for ammonia oxidation to nitrogen at low temperature. J. Catal. 2002, 206, 60–70. [Google Scholar] [CrossRef]
  72. Gang, L.; Anderson, B.G.; Van Grondelle, J.; Van Santen, R.A. Low temperature selective oxidation of ammonia to nitrogen on silver-based catalysts. Appl. Catal. B Environ. 2003, 40, 101–110. [Google Scholar] [CrossRef]
  73. Zhang, L.; He, H. Mechanism of selective catalytic oxidation of ammonia to nitrogen over Ag/Al2O3. J. Catal. 2009, 268, 18–25. [Google Scholar] [CrossRef]
  74. Zhang, L.; Zhang, C.; He, H. The role of silver species on Ag/Al2O3 catalysts for the selective catalytic oxidation of ammonia to nitrogen. J. Catal. 2009, 261, 101–109. [Google Scholar] [CrossRef]
  75. Zhang, L.; Fudong, L.I.U.; Yunbo, Y.U.; Yongchun, L.I.U.; Zhang, C.; Hong, H.E. Effects of adding CeO2 to Ag/Al2O3 catalyst for ammonia oxidation at low temperatures. Chin. J. Catal. 2011, 32, 727–735. [Google Scholar] [CrossRef]
  76. Wang, F.; He, G.; Zhang, B.; Chen, M.; Chen, X.; Zhang, C.; He, H. Insights into the activation effect of H2 pretreatment on Ag/Al2O3 catalyst for the selective oxidation of ammonia. ACS Catal. 2019, 9, 1437–1445. [Google Scholar] [CrossRef]
  77. Lippits, M.J.; Gluhoi, A.C.; Nieuwenhuys, B.E. A comparative study of the selective oxidation of NH3 to N2 over gold, silver and copper catalysts and the effect of addition of Li2O and CeOx. Catal. Today 2008, 137, 446–452. [Google Scholar] [CrossRef]
  78. Qu, Z.; Wang, H.; Wang, S.; Cheng, H.; Qin, Y.; Wang, Z. Role of the support on the behavior of Ag-based catalysts for NH3 selective catalytic oxidation (NH3-SCO). Appl. Surf. Sci. 2014, 316, 373–379. [Google Scholar] [CrossRef]
  79. Yang, M.; Wu, C.; Zhang, C.; He, H. Selective oxidation of ammonia over copper-silver-based catalysts. Catal. Today 2004, 90, 263–267. [Google Scholar] [CrossRef]
  80. Jabłońska, M.; Beale, A.M.; Nocuń, M.; Palkovits, R. Ag-Cu based catalysts for the selective ammonia oxidation into nitrogen and water vapour. Appl. Catal. B Environ. 2018, 232, 275–287. [Google Scholar] [CrossRef] [Green Version]
  81. Wang, F.; Ma, J.; He, G.; Chen, M.; Wang, S.; Zhang, C.; He, H. Synergistic effect of TiO2-SiO2 in Ag/Si-Ti catalyst for the selective catalytic oxidation of ammonia. Ind. Eng. Chem. Res. 2018, 57, 11903–11910. [Google Scholar] [CrossRef]
  82. Jabłońska, M.; Ciptonugroho, W.; Góra-Marek, K.; Al-Shaal, M.G.; Palkovits, R. Preparation, characterization and catalytic performance of Ag-modified mesoporous TiO2 in low-temperature selective ammonia oxidation into nitrogen and water vapour. Microporous Mesoporous Mater. 2017, 245, 31–44. [Google Scholar] [CrossRef]
  83. Wang, Z.; Sun, Q.; Wang, D.; Hong, Z.; Qu, Z.; Li, X. Hollow ZSM-5 zeolite encapsulated Ag nanoparticles for SO2-resistant selective catalytic oxidation of ammonia to nitrogen. Sep. Purif. Technol. 2019, 209, 1016–1026. [Google Scholar] [CrossRef]
  84. Gang, L.; Anderson, B.G.; Van Grondelle, J.; Van Santen, R.A. Intermediate species and reaction pathways for the oxidation of ammonia on powdered catalysts. J. Catal. 2001, 199, 107–114. [Google Scholar] [CrossRef]
  85. Karatok, M.; Vovk, E.I.; Koc, A.V.; Ozensoy, E. Selective catalytic ammonia oxidation to nitrogen by atomic oxygen species on Ag (111). J. Phys. Chem. C 2017, 121, 22985–22994. [Google Scholar] [CrossRef]
  86. Carabineiro, S.A.C.; Nieuwenhuys, B.E. Selective oxidation of ammonia over Ir (510). Comparison with Ir (110). Surf. Sci. 2003, 532, 87–95. [Google Scholar] [CrossRef]
  87. Jabłońska, M.; Nothdurft, K.; Nocuń, M.; Girman, V.; Palkovits, R. Redox-performance correlations in Ag–Cu–Mg–Al, Ce–Cu–Mg–Al, and Ga–Cu–Mg–Al hydrotalcite derived mixed metal oxides. Appl. Catal. B Environ. 2017, 207, 385–396. [Google Scholar] [CrossRef]
  88. Qu, Z.; Wang, Z.; Quan, X.; Wang, H.; Shu, Y. Selective catalytic oxidation of ammonia to N2 over wire-mesh honeycomb catalyst in simulated synthetic ammonia stream. Chem. Eng. J. 2013, 233, 233–241. [Google Scholar] [CrossRef]
  89. Zhou, M.; Wang, Z.; Sun, Q.; Wang, J.; Zhang, C.; Chen, D.; Li, X. High-performance Ag-Cu nanoalloy catalyst for the selective catalytic oxidation of ammonia. ACS Appl. Mater. Interfaces 2019, 11, 46875–46885. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, H.; Lin, M.; Murayama, T.; Feng, S.; Haruta, M.; Miura, H.; Shishido, T. Ag size/structure-dependent effect on low-temperature selective catalytic oxidation of NH3 over Ag/MnO2. ACS Catal. 2021, 11, 8576–8584. [Google Scholar] [CrossRef]
  91. Góra-Marek, K.; Tarach, K.A.; Piwowarska, Z.; Łaniecki, M.; Chmielarz, L. Ag-loaded zeolites Y and USY as catalysts for selective ammonia oxidation. Catal. Sci. Technol. 2016, 6, 1651–1660. [Google Scholar] [CrossRef]
  92. Martinez-Ortigosa, J.; Lopes, C.W.; Agostini, G.; Palomares, A.E.; Blasco, T.; Rey, F. AgY zeolite as catalyst for the selective catalytic oxidation of NH3. Microporous Mesoporous Mater. 2021, 323, 111230. [Google Scholar] [CrossRef]
  93. Guzman, J.; Gates, B.C. Catalysis by supported gold: Correlation between catalytic activity for CO oxidation and oxidation states of gold. J. Am. Chem. Soc. 2004, 126, 2672–2673. [Google Scholar] [CrossRef] [PubMed]
  94. Kaskow, I.; Sobczak, I.; Ziolek, M.; Corberán, V.C. The effect of support properties on n-octanol oxidation performed on gold-silver catalysts supported on MgO, ZnO and Nb2O5. Mol. Catal. 2020, 482, 110674. [Google Scholar] [CrossRef]
  95. Lin, S.D.; Gluhoi, A.C.; Nieuwenhuys, B.E. Ammonia oxidation over Au/MOx/γ-Al2O3—Activity, selectivity and FTIR measurements. Catal. Today 2004, 90, 3–14. [Google Scholar] [CrossRef]
  96. Gong, J.; Ojifinni, R.A.; Kim, T.S.; White, J.M.; Mullins, C.B. Selective catalytic oxidation of ammonia to nitrogen on atomic oxygen precovered Au (111). J. Am. Chem. Soc. 2006, 128, 9012–9013. [Google Scholar] [CrossRef]
  97. Liu, R.; Shen, W.; Zhang, J.; Li, M. Adsorption and dissociation of ammonia on Au (111) surface: A density functional theory study. Appl. Surf. Sci. 2008, 254, 5706–5710. [Google Scholar] [CrossRef]
  98. Wang, Y.; Jacobi, K.; Schöne, W.-D.; Ertl, G. Catalytic oxidation of ammonia on RuO2 (110) surfaces: Mechanism and selectivity. J. Phys. Chem. B 2005, 109, 7883–7893. [Google Scholar] [CrossRef]
  99. Seitsonen, A.P.; Crihan, D.; Knapp, M.; Resta, A.; Lundgren, E.; Andersen, J.N.; Over, H. Reaction mechanism of ammonia oxidation over RuO2 (110): A combined theory/experiment approach. Surf. Sci. 2009, 603, L113–L116. [Google Scholar] [CrossRef]
  100. Perez-Ramirez, J.; López, N.; Kondratenko, E.V. Pressure and materials effects on the selectivity of RuO2 in NH3 oxidation. J. Phys. Chem. C 2010, 114, 16660–16668. [Google Scholar] [CrossRef]
  101. Cui, X.; Chen, L.; Wang, Y.; Chen, H.; Zhao, W.; Li, Y.; Shi, J. Fabrication of hierarchically porous RuO2-CuO/Al-ZrO2 composite as highly efficient catalyst for ammonia-selective catalytic oxidation. ACS Catal. 2014, 4, 2195–2206. [Google Scholar] [CrossRef]
  102. Cui, X.; Zhou, J.; Ye, Z.; Chen, H.; Li, L.; Ruan, M.; Shi, J. Selective catalytic oxidation of ammonia to nitrogen over mesoporous CuO/RuO2 synthesized by co-nanocasting-replication method. J. Catal. 2010, 270, 310–317. [Google Scholar] [CrossRef]
  103. Chakraborty, D.; Damsgaard, C.D.; Silva, H.; Conradsen, C.; Olsen, J.L.; Carvalho, H.W.P.; Mutz, B.; Bligaard, T.; Hoffmann, M.J.; Grunwaldt, J.-D.; et al. Bottom-up design of a copper-ruthenium nanoparticulate catalyst for low-temperature ammonia oxidation. Angew. Chem. 2017, 129, 8837–8841. [Google Scholar] [CrossRef]
  104. Wang, H.; Ning, P.; Zhang, Q.; Liu, X.; Zhang, T.; Fan, J.; Wang, J.; Long, K. Promotional mechanism of WO3 over RuO2-Fe2O3 catalyst for NH3-SCO reaction. Appl. Catal. A Gen. 2018, 561, 158–167. [Google Scholar] [CrossRef]
  105. Chen, W.; Ma, Y.; Qu, Z.; Liu, Q.; Huang, W.; Hu, X.; Yan, N. Mechanism of the selective catalytic oxidation of slip ammonia over Ru-modified Ce-Zr complexes determined by in situ diffuse reflectance infrared Fourier transform spectroscopy. Environ. Sci. Technol. 2014, 48, 12199–12205. [Google Scholar] [CrossRef]
  106. Chen, W.; Qu, Z.; Huang, W.; Hu, X.; Yan, N. Novel effect of SO2 on selective catalytic oxidation of slip ammonia from coal-fired flue gas over IrO2 modified Ce-Zr solid solution and the mechanism investigation. Fuel 2016, 166, 179–187. [Google Scholar] [CrossRef]
  107. Zhang, Q.; Zhang, T.; Xia, F.; Zhang, Y.; Wang, H.; Ning, P. Promoting effects of acid enhancing on N2 selectivity for selectivity catalytic oxidation of NH3 over RuOx/TiO2: The mechanism study. Appl. Surf. Sci. 2020, 500, 144044. [Google Scholar] [CrossRef]
  108. Liao, Y.; Xu, H.; Li, Z.; Ji, L.; Wang, L.; Gao, G.; Huang, W.; Qu, Z.; Yan, N. Boosting RuO2 surface reactivity by Cu active sites over Ru/Cu-SSZ-13 for simultaneous catalytic oxidation of CO and NH3. J. Phys. Chem. C 2021, 125, 17031–17041. [Google Scholar] [CrossRef]
  109. Heylen, S.; Delcour, N.; Kirschhock, C.E.A.; Martens, J.A. Selective catalytic oxidation of ammonia into dinitrogen over a zeolite-supported ruthenium dioxide catalyst. ChemCatChem 2012, 4, 1162–1166. [Google Scholar] [CrossRef]
  110. Shin, J.H.; Kim, G.J.; Hong, S.C. Reaction properties of ruthenium over Ru/TiO2 for selective catalytic oxidation of ammonia to nitrogen. Appl. Surf. Sci. 2020, 506, 144906. [Google Scholar] [CrossRef]
Figure 1. Conversion of NO over the PGMs on pulsing of c(15NH3):c(NO):c(Ne) = 1:0.2:1 (solid bars) and c(15NH3):c(NO):c(Ne) = 1:2:1 (open bars) at 800 °C. Reprinted from [25] with permission from Elsevier.
Figure 1. Conversion of NO over the PGMs on pulsing of c(15NH3):c(NO):c(Ne) = 1:0.2:1 (solid bars) and c(15NH3):c(NO):c(Ne) = 1:2:1 (open bars) at 800 °C. Reprinted from [25] with permission from Elsevier.
Molecules 26 06461 g001
Figure 2. (a) Reaction mechanisms and Pt states over Pt/Al2O3 and (b) Pt/CeZrO2. Reprinted from [42] with permission from ACS Publications.
Figure 2. (a) Reaction mechanisms and Pt states over Pt/Al2O3 and (b) Pt/CeZrO2. Reprinted from [42] with permission from ACS Publications.
Molecules 26 06461 g002
Figure 4. (a) Overall NH3 reaction pathway on silver powder. Reprinted from [84] with permission from Elsevier; (b) reactivity and selectivity trends of NH3-SCO on O/Ag(111) as a function of the oxygen coverage and temperature. Reprinted from [85] with permission from ACS Publications.
Figure 4. (a) Overall NH3 reaction pathway on silver powder. Reprinted from [84] with permission from Elsevier; (b) reactivity and selectivity trends of NH3-SCO on O/Ag(111) as a function of the oxygen coverage and temperature. Reprinted from [85] with permission from ACS Publications.
Molecules 26 06461 g004
Figure 5. Energy-dispersive X-ray mapping for corresponding elemental distribution of Ag, Cu, and O on Ag2Cu1 (ad) and AgCuOx NPs (eh). Reprinted from [89] with permission from ACS Publications.
Figure 5. Energy-dispersive X-ray mapping for corresponding elemental distribution of Ag, Cu, and O on Ag2Cu1 (ad) and AgCuOx NPs (eh). Reprinted from [89] with permission from ACS Publications.
Molecules 26 06461 g005
Figure 6. (a) NH3 conversion and N2 selectivity over Au/MOx at 150°C. Gold loading amount was 1 wt.%. Reaction conditions: a mass of catalyst, 0.15 g; 0.005 vol.% NH3 and 20 vol.% O2, Ar balance, GHSV 40 000 mL h−1 g−1; (b) Suggested NH3-SCO mechanisms of Au/Nb2O5. Reprinted from [12] with permission of ACS Publications.
Figure 6. (a) NH3 conversion and N2 selectivity over Au/MOx at 150°C. Gold loading amount was 1 wt.%. Reaction conditions: a mass of catalyst, 0.15 g; 0.005 vol.% NH3 and 20 vol.% O2, Ar balance, GHSV 40 000 mL h−1 g−1; (b) Suggested NH3-SCO mechanisms of Au/Nb2O5. Reprinted from [12] with permission of ACS Publications.
Molecules 26 06461 g006
Figure 7. Microscopic reaction steps in NH3 oxidation over RuO2(110). Activation energies (red) and total adsorption energies (black) are determined by DFT calculations and are given in kJ mol−1. -Hdiff means that abstracted H from NHx is removed from its direct neighborhood by diffusion along with various O species on the surface. Reprinted from [99] with permission from Elsevier.
Figure 7. Microscopic reaction steps in NH3 oxidation over RuO2(110). Activation energies (red) and total adsorption energies (black) are determined by DFT calculations and are given in kJ mol−1. -Hdiff means that abstracted H from NHx is removed from its direct neighborhood by diffusion along with various O species on the surface. Reprinted from [99] with permission from Elsevier.
Molecules 26 06461 g007
Figure 8. Mechanism of NH3-SCO and effect of SO2. Reprinted from [105] with permission of ACS Publications.
Figure 8. Mechanism of NH3-SCO and effect of SO2. Reprinted from [105] with permission of ACS Publications.
Molecules 26 06461 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jabłońska, M. Progress on Noble Metal-Based Catalysts Dedicated to the Selective Catalytic Ammonia Oxidation into Nitrogen and Water Vapor (NH3-SCO). Molecules 2021, 26, 6461. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26216461

AMA Style

Jabłońska M. Progress on Noble Metal-Based Catalysts Dedicated to the Selective Catalytic Ammonia Oxidation into Nitrogen and Water Vapor (NH3-SCO). Molecules. 2021; 26(21):6461. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26216461

Chicago/Turabian Style

Jabłońska, Magdalena. 2021. "Progress on Noble Metal-Based Catalysts Dedicated to the Selective Catalytic Ammonia Oxidation into Nitrogen and Water Vapor (NH3-SCO)" Molecules 26, no. 21: 6461. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26216461

Article Metrics

Back to TopTop