Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Insights into N2O decomposition in environmental catalysis: evaluation and an advanced outlook

Khezina Rafiqa, Mamoona Sabira, Iqra Sadiaa, Muhammad Zeeshan Abida, Muhammad Arif Nadeemb and Ejaz Hussain*ab
aInstitute of Chemistry, Inorganic Materials Laboratory 52S, The Islamia University of Bahawalpur, 63100, Pakistan. E-mail: ejaz.hussain@iub.edu.pk; khezina.rafiq@iub.edu.pk
bCatalysis and Nanomaterials Lab 27, Department of Chemistry, Quaid-i-Azam University, Islamabad 45320, Pakistan

Received 24th June 2025 , Accepted 13th September 2025

First published on 15th September 2025


Abstract

Nitrous oxide (N2O) is a potent greenhouse gas and a significant contributor to stratospheric ozone depletion. Its emission arises from a combination of natural and anthropogenic sources, including microbial processes such as nitrification, denitrification, nitrifier denitrification and abiotic chemical pathways such as chemodenitrification and hydroxylamine oxidation. Effective decomposition of N2O into harmless nitrogen and oxygen is therefore critical for climate mitigation. This review provides a comprehensive overview of the current state of N2O decomposition, with a focus on mechanisms, catalyst composition and material design strategies. Catalysts are systematically classified based on their decomposition pathways i.e. direct catalytic decomposition (DCD), selective catalytic reduction (SCR), and surface-mediated redox processes. In addition, they are further classified on the basis of their compositional attributes, encompassing both precious as well as and non-precious metal-based catalysts systems. Furthermore, a progress-based classification is presented, encompassing classical metal oxides, spinel oxides, layered double hydroxides, MXenes, and metal–organic frameworks (MOFs). Emerging systems such as antenna-reactor catalysts and quantum dots have also been discussed based on their unique properties and potential. Mechanistic insights into N2O activation, including thermal, surface-catalysed, Mars–van Krevelen-type redox, radical and photocatalytic pathways, are explored in depth. This review highlights the significance of catalyst design, oxygen vacancy engineering and atomically dispersed active sites in enhancing activity and selectivity. Future perspectives emphasize the development of low-cost, thermally stable and environmentally benign catalysts, along with gaining a mechanistic understanding through in situ studies and computational modelling. This review aims to guide the rational design of next-generation catalysts for efficient N2O abatement across industrial and environmental systems.


1. Introduction

Climate change has become a challenge to international agencies such as the WHO, EPA, IPCC (Intergovernmental Panel on Climate Change), AFED (Arab Forum for Environment and Development), APAN (Asian Pacific Adaptation Network), CCCCC (Caribbean Community Climate Change Centre), CAN (Climate Action Network), CCL (Citizens’ Climate Lobby), CJA (Climate Justice Alliance), EDF (Environmental Defense Fund, USA), HEA (Health and Environmental Alliance) and NRDC (Natural Resources Defense Council).1 For a long time, nitrous oxide (N2O) was not considered a significant environmental pollutant, and it did not receive much attention from an environmental perspective. In recent decades, N2O has been recognized as a potent greenhouse gas. It contributes to the greenhouse effect by trapping the heat in Earth's atmosphere and leading to global warming. In addition, it is believed to play a role in the destruction of ozone in the stratosphere, contributing to ozone layer depletion.2 The presence of N2O in the atmosphere causes acid rain, which can have detrimental effects on the ecosystem and environment. Moreover, N2O raises a particular concern due to its long atmospheric lifetime, estimated at approximately 120 years. This means that once released into the atmosphere, it persists for a very long time.3 The global warming potential (GWP) of N2O is nearly 310 times that of carbon dioxide (CO2). This high GWP makes it a significant contributor to climate change.4 Chemical industries, especially those involving processes such as nitric acid production and the industrial synthesis of adipic acid, are major sources of anthropogenic N2O emissions. Specifically, in the synthesis of adipic acid, N2O is produced when a cyclohexanone/cyclohexanol mixture reacts with nitric acid. It is noted that N2O emissions from the adipic acid industry contribute a substantial percentage (5–8%) to anthropogenic sources of N2O in the atmosphere.5,6 Both natural and anthropogenic sources of N2O exist, with industrial sources responsible for emitting a significant amount (400–500 kt) annually.7

2. N2O generation mechanisms

N2O is generated through multiple pathways, primarily mediated by microbial processes, and also via abiotic chemical reactions. Key mechanisms include nitrification, where ammonia is oxidized and N2O is formed as a by-product, and denitrification, where nitrate is reduced to nitrogen gases, often realising N2O when the process in incomplete. Nitrifier denitrification also plays a role under oxygen-limiting conditions, while abiotic chemical reactions, such as nitrate reduction by iron or manganese, contribute to additional N2O fluxes. On top of natural processes, human activities significantly amplify these emissions, particularly through intensive agriculture, wastewater management and industrial processes. Different mechanisms are summarized in this section, shown in Fig. 1.
image file: d5ma00668f-f1.tif
Fig. 1 Key mechanisms of N2O generation.

2.1. Nitrification

Nitrification is an aerobic microbial process where ammonium (NH4+) or ammonia (NH3) is oxidized to nitrate (NO3) via nitrite (NO2). This process is mediated by ammonia-oxidizing archaea (AOA), as well as nitrate-oxidizing bacteria (NOB) in a two-step sequence. The oxidation of hydroxylamine (NH2OH), a key intermediate, can inadvertently release N2O as a by-product. N2O production during nitrification is particularly pronounced under suboptimal oxygen conditions, low pH, or high substrate (ammonia) availability, which destabilize the electron transport chain in AOB. Moreover, recent studies have suggested that AOA contribute significantly to N2O emissions in acidic soils, where bacterial nitrification is less efficient. Thus, nitrification represents both a fundamental nitrogen cycle process and a critical source of N2O in agricultural and natural soils.8

2.2. Denitrification

Denitrification is an anaerobic respiratory process in which nitrate (NO3) is sequentially reduced to nitrite (NO2), nitric oxide (NO), nitrous oxide (N2O) and finally dinitrogen gas (N2). This pathway is primarily driven by facultative anaerobic bacteria such as Pseudomonas, Paracoccus, and Bacillus species under low-oxygen or anoxic conditions. The process requires organic carbon as an electron donor, linking it closely to carbon cycling. The accumulation of N2O often occurs when the enzyme nitrous oxide reductase (NosZ), responsible for reducing N2O to N2, is inhibited by low pH, low copper availability, or the presence of oxygen. As a result, soils and sediments under fluctuating oxygen conditions, such as wetlands, wastewater systems, and agricultural fields after irrigation, often exhibit high N2O fluxes. Denitrification is considered the largest single biological source of N2O globally.9

2.3. Nitrifier denitrification

Nitrifier denitrification is a hybrid pathway in which nitrifying organisms, under oxygen-limited or microaerophilic conditions, reduce nitrite (NO2) to nitric oxide (NO), nitrous oxide (N2O) and nitrogen gas (N2). Unlike conventional denitrification carried out by facultative anaerobes, this process is mediated by ammonia-oxidizing bacteria (AOB). The pathway is particularly relevant in soils, sediments, and wastewater treatment systems where oxygen availability fluctuates. Nitrifier denitrification can account for substantial N2O emissions when ammonia and nitrite are both abundant, such as in fertilized agricultural soils. It provides a mechanistic link between nitrification and denitrification, highlighting the versatility of microbial nitrogen metabolism under transitional redox conditions.8

2.4. Abiotic chemical pathways

Although microbial processes dominate, abiotic pathways also contribute to N2O formation. These include: (i) chemodenitrification: it involves the reduction of nitrites by ferrous ions (Fe2+) or other reductants, leading to N2O release. This mechanism is common in waterlogged soils, sediments and mining-impacted environments with high levels of reduced metals. (ii) Oxidation of hydroxylamine can occur through reactions with metal oxides, particularly ferric (Fe3+) and manganese (Mn4+) oxides, producing N2O and NO. Additionally, photochemical reactions in sunlit waters and soils have been reported to drive abiotic N2O formation from nitrite. While biotic pathways may represent a smaller fraction globally, they can dominate under specific geochemical conditions, especially in soils rich in iron or manganese.10

2.5. Anthropogenic sources

Human activities have amplified N2O emissions far beyond natural background levels. These include (i) agriculture: excessive application of nitrogen-based fertilizers (urea, ammonium nitrate) and livestock manure leads to surplus reactive nitrogen in soils, fueling nitrification, denitrification and nitrifier denitrification. Agricultural soils are estimated to contribute more than 60% of global anthropogenic N2O emissions. (ii) Wastewater treatment: biological nitrogen removal processes, particularly in activated sludge and sequencing batch reactors, release N2O due to incomplete nitrification and denitrification under variable oxygen conditions. Emerging evidence shows that wastewater plants may be hotspots of N2O release if not properly managed. (iii) Industrial processes: N2O is also released during the production of nitric acid (used in fertilizers) and adipic acid (used in nylon manufacture). Though mitigation technologies exist, incomplete adoption leads to continued emissions. (iv) Biomass burning and fossil fuel combustion: smaller but notable contributions come from burning crop residues, forests, and fossil fuels, releasing reactive nitrogen that indirectly contributes to N2O formation. Together, these anthropogenic sources have increased atmospheric N2O concentrations to record highs, making it the third most important long-lived greenhouse gas and a major contributor to stratospheric ozone depletion. Fig. 2a shows the distribution of natural versus anthropogenic sources.
image file: d5ma00668f-f2.tif
Fig. 2 (a) Natural and anthropogenic sources of N2O (infamous laughing gas). (b) Thermal and catalytic paths of N2O decomposition. (c) Functioning of Cu-exchanged zeolites in the selective catalytic reduction of NOx from mobile sources.11 Reproduced with permission from the American Chemical Society (Copyright©2013) (d). Advances in zeolite synthesis and characterization techniques for heterogeneous catalysis.

3. Categories of N2O decomposition catalysts based on their mechanisms and compositions

Catalysts for N2O decomposition are generally classified into different categories depending on their catalytic mechanism, composition, and support material. This classification helps us in understanding theory operational principles, activity under industrial conditions, cost-effectiveness and long-term stability. A detailed overview of each category is presented below.

3.1. Based on catalytic mechanisms

3.1.1. Direct catalytic decomposition (DCD). In this mechanism, catalysts decompose nitrous oxide (N2O) directly into N2 and O2 without the need for a reducing agent:
2N2O → 2N2 + O2
Examples of such catalysts include metal oxides [such as cobalt oxides (Co3O4), iron oxide (Fe2O3) and manganese oxide (MnO2)], perovskites (e.g. La1−xSrxCoO3 with tunable oxygen vacancies that improve oxygen mobility and enhance N2O activation) and spinels (e.g. CuFe2O4, MnAl2O4) known for their stable structures and redox properties. In DCD, oxygen vacancies on these catalysts adsorb and activate N2O, leading to bond cleavage and release of N2. The oxygen atoms are either released as O2 or incorporated temporarily into the lattice. These catalysts are widely considered for industrial emission control (e.g. nitric acid plants) where large volumes of N2O are generated.
3.1.2. Selective catalytic reduction (SCR). In SCR, a reducing agent (such as NH3, CO or hydrocarbons like CH4 and C3H6) is introduced to facilitate the reduction of N2O into N2:
N2O + 2CO → N2 + 2CO2

N2O + 2NH3 → 2N2 + 3H2O
Examples of these catalysts include Fe- or Cu-exchanged zeolites (Fe–ZSM-5, Cu–SSZ-13, and Cu–ZSM-5) and vanadium-based catalysts (V2O5/TiO2 and V2O5–WO3/TiO2). The reducing agent reacts with oxygen species from N2O decomposition, enhancing N2 formation. SCR can operate at relatively low temperatures compared to DCD, making it practical for vehicular exhaust systems and stationary sources. SCR is often applied in mixed gas systems such as automotive exhausts, power plants, and mobile sources. Fig. 2c demonstrates the role of Cu-zeolites in ammonia-assisted SCR of N2O, particularly effective in diesel engines.

3.2 Based on catalyst compositions

3.2.1. Precious metal-based catalysts. Precious metals such as Rh, Ru, Pt and Pd are highly active for N2O decomposition. They promote rapid N–O bond cleavage due to their strong adsorption capacity for N2O. Examples include Rh/Al2O3, Ru/CeO2, and Pt/ZrO2. High catalytic activity and efficiency at relatively low temperatures are their key advantages. Their limitations are high cost (that makes them less suitable for large-scale industrial use) and susceptibility to deactivation due to sintering or poisoning (e.g. by sulphur or water vapours). They are often used in niche or specialized applications where high efficiency is required despite high costs (e.g. medical gas purification or research setups).
3.2.2. Non-precious metal-based catalysts. Non-precious metals (mainly transition metals and mixed oxides) are more cost-effective and environmentally sustainable, making them widely used in industrial processes. Examples include transition metal oxides (e.g. Co3O4, MnOx, CuO, and Fe2O3), perovskites (LaCoO3, LaMnO3, and Sr-doped cobaltites with enhanced oxygen mobility), spinels (CuFe2O4 and MnAl2O4) and zeolites (Fe–ZSM-5 and Cu–SSZ-13 with high hydrothermal stability). Their advantages include low cost, abundant raw materials and high thermal stability. They generally require higher operating temperatures than that needed for precious metals, but can be optimized by doping or adding promoters. These are widely employed in industrial emission abatement systems, including nitric acid and adipic acid plants, as well as power generation facilities.

3.3 Based on support types

The support material plays a crucial role in dispersing active catalytic sites, enhancing redox behaviour, and providing structural stability under harsh reaction conditions. Common supports include metal oxides such as Al2O3, TiO2 and CeO2 due to their high surface area and redox properties. CeO2, in particular, enhances oxygen storage and mobility, improving the catalytic efficiency. Zeolites include framework structures (such as ZSM-5, SSZ-13 and beta-zeolites) provide high surface area and microporosity. Their ion-exchange ability allows the incorporation of Fe or Cu cations, which significantly enhances SCR activity. SSZ-13 is particularly noted for high hydrothermal stability under automotive exhaust conditions. In carbon-based supports, activated carbon and graphene are sometimes used due to their large surface area and electrical conductivity. However, they are less common for N2O decomposition because of limited thermal stability at high operating temperatures. The choice of support directly influences catalyst dispersion, durability, oxygen mobility and tolerance to poisons (e.g. SO2 and H2O). Hence, support selection is as important as the active catalytic phase itself.

4. Significance of catalytic decomposition of N2O

N2O is also a by-product in various industrial processes, including the catalytic abatement of nitrogen oxides (NOx) in three-way catalytic converters and selective catalytic reduction (SCR) of NOx with NH3. As N2O is a significant greenhouse gas and its emission contributes to the global warming and climate change, there is a pressing need to reduce N2O emissions from anthropogenic sources. One effective and economical method to reduce N2O emissions is the catalytic decomposition of N2O into its elemental components (nitrogen and oxygen), which is known as catalytic decomposition of N2O.12,13 To make it effective, the catalyst must be durable and selective. Durability ensures long-term effectiveness of the catalyst and selectivity ensures that N2O is effectively converted into nitrogen and oxygen without the formation of undesirable by-products.14 Catalysts designed for N2O decomposition should be stable even in the presence of moisture, as they may encounter atmospheric moisture during real-world applications. N2O decomposition is an exothermic reaction, meaning that it releases heat during the process. This can occur either thermally (without any catalyst) or catalytically (Fig. 2b). The choice between these paths depends on the factors such as operating temperature and the amount of catalysts used.15 However, noble metal-based photocatalysts are among reliable options for N2O decomposition. These catalysts utilize noble metals (e.g. Pt and Pd) to facilitate the decomposition of N2O under the influence of light, making them a potential choice for efficient N2O abatement.16

A considerable amount of work has been done in the catalytic decomposition of N2O, and a number of catalyst systems with various compositions such as zinc, platinum, copper oxides and zeolites have been developed.17 Metal surfaces,18 supported metals,19 metal oxides,20,21 supported oxides,22 ternary oxides,23 spinel oxides,24 perovskite-type oxides,25 hydrotalcites,26 and transition metal exchanged zeolites27 are well-known catalysts that can decompose N2O into its elements even at low temperatures. In this review, important categories of N2O decomposition catalysts are covered, which include transition metal-based zeolites, hydrotalcites, spinel oxides, mixed metal oxides, metal/metal oxides in mesoporous silica, metal oxides, metal-doped nanotubes, nanosheets and nanocages, graphene and graphitic oxides/nitrides, metal–organic frameworks (MOFs), and MOF-derived catalysts. This review compares the different synthesis strategies of catalyst fabrication, experimental conditions, catalyst characterization techniques, N2O decomposition temperatures and catalytic efficiencies of various catalysts (Fig. 2d).

5. Progress-based systematic classification of N2O decomposition catalysts

N2O decomposition catalytic materials are classified into five main categories based on their structure, composition and research progress/maturity. Each category includes subtypes ordered by their level of maturity, from well-established (extensively studied) to emerging (recently developed or still under intense research).

5.1. Classical metal oxide-based catalysts

Classical metal oxide-based catalysts such as Fe2O3, Co3O4 and MnO2 are most well-established and extensively studied for decades. They are renowned for their thermal stability, redox flexibility, and cost-effectiveness. These oxides typically possess high oxygen mobility and variable oxidation states, enabling efficient N2O decomposition via lattice oxygen participation. Their simple structures allow for easy synthesis, scalability, and tunability through doping or morphology control. Spinel-type oxides, in particular, exhibit strong metal–oxygen interactions and good resistance to sintering. Despite their maturity, continued research enhances their surface area, dispersion, and durability under reaction conditions. Their robustness and versatility make them fundamental benchmarks in both industrial and environmental catalytic applications. Four categories of these catalysts are particularly important with respect to N2O decomposition.
5.1.1. Metal oxide catalysts. Transition metal oxides such as NiO, Co3O4, MnO2, CuO, Cr2O3 and Fe2O3 exhibit high catalytic activity for the decomposition of N2O. The transition metal oxides are known for their high thermal and chemical stability, making them suitable candidates for catalytic applications.28–30 For example, Gaidei and co-workers synthesized a series of oxide catalysts containing up to 30% active metal components on an alumina carrier. The temperature dependence of N2O decomposition on these catalysts follow a well-defined S-shaped curve, similar to Pt family metals, but these transition metal oxide catalysts are somewhat less active than Pt-based catalysts. The catalytic activity of these metal oxide catalysts was ranked as follows in the increasing order of decomposition temperature: Fe2O3 < Cr2O3 < CuO < MnO2 < NiO < Co3O4. Among the catalysts based on non-noble metals, Co3O4 showed the highest catalytic activity. The calculated activation energy values for these catalysts vary, with CuO and Cr2O3 having the highest Ea (187–176 kJ mol−1) and MnO2, Co3O4 and NiO having lower values (133–143 kJ mol−1). Based on experimental data, Co, Cu, Mn and Cr oxides are recommended as active components for developing deposited catalysts based on non-noble metals.31 Fig. 3 signifies the use of acidic and high valence metal sites on the various metal oxide catalysts based on MnOx, CeO2, Fe2O3, VOx and CuO for the NH3-assited SCR of NOx.32
image file: d5ma00668f-f3.tif
Fig. 3 Selective catalytic reduction of nitrogen oxides with NH3 employing various acidic metal oxide composite catalysts.32 Reproduced with permission from the American Chemical Society (Copyright©2019).

CaO obtained from the domestic limestone has also demonstrated its catalytic effect in N2O decomposition. The effects of CO and CO2 on N2O conversion activity were found antagonistic to each other. On the one hand, CO increases the N2O decomposition activity. On the other hand, higher CO2 concentrations slowly decrease the N2O decomposition activity. As NO was also a hindering factor in N2O decomposition, in order to lessen its concentration, N2O was concurrently passed in the circulating fluidized bed combustion (CFBC) reactor. Selective non-catalytic reduction (SNCR) is a method of choice for the NO removal from flue gas.33 Similarly, various Bi-based semiconductors were fabricated by a hydrothermal strategy,34–36 e.g. Liu group examined monoclinic BiVO4, orthorhombic Bi2MoO6 and Bi2WO6 catalysts for the photocatalytic degradation of N2O into N2 and O2, as shown in Fig. 4a. They found that BiVO4 exhibited the highest degradation activity due to stronger absorption of visible light and higher charge separation.37


image file: d5ma00668f-f4.tif
Fig. 4 (a) Photocatalytic N2O decomposition via Bi-based photocatalysts (i.e. BiVO4, Bi2MoO6 and Bi2WO6) under visible light irradiation.37 Reproduced with permission from the American Chemical Society (Copyright©2018). (b) Mans–van Krevelen mechanism over phosphotungstic acid supported SACs for the reduction of N2O by CO.38 Reproduced with permission from the American Chemical Society (Copyright©2019).

Metal-supported metal oxide catalysts are also popular for N2O decomposition reactions. Komvokis et al. synthesized a series of Ru/Al2O3 nanocatalysts via in situ reduction with ethylene glycol (EG) and compared their N2O reduction activity with the catalysts prepared through the typical incipient wetness procedure. The EG reduction approach resulted in the formation of spherical Ru nanoparticles (particle size: 1–3 nm; dispersion: 70–35%). While impregnated calcined catalysts had larger sized nanoparticles with a very small dispersion degree (particle size: 10–80 nm; dispersion: 10%). The catalysts synthesized via EG reduction showed significantly higher N2O decomposition activity in an O2-rich environment irrespective of CO, CO2, NO, H2O and SO2. The higher surface area of Ru metal contributes to their higher catalytic activity. These catalysts have a low apparent activation energy, which suggests that they require less energy to initiate the chemical reaction. Ru/Al2O3 nanocatalysts can be regenerated effectively that helps to restore their catalytic activity, achieving a conversion rate of over 90%.39 Similarly, Reddya et al. synthesized alumina-supported Pd catalysts (Pd/Al2O3) employing the deposition precipitation (DP) and impregnation (IMP) methods and studied their effect on N2O conversion activity. Higher catalytic efficiency was recorded for DP catalysts as compared to their IMP equivalents due to the formation of partially oxidized Pd2+ species over the surface of the DP–Cl catalyst.40

In a general N2O decomposition reaction, CO is adsorbed over the surface of catalysts to reduce N2O, but this strong adsorption of CO may lead to the catalyst's poisoning. For example, Zhang et al. have utilized the Mans-van-Krevelen mechanism by employing phosphotungstic acid (PTA)-supported single-atom catalysts (SACs), M/PTA, where M = Fe, Co, Mn, Rh, Ru, Ir, Os, Pt and Pd. Adsorbed CO reacts with the surface oxygen atoms of PTA support and create holes on the M/PTA surface. N2O acts as an oxygen donor to regenerate catalyst and releases N2. It was also demonstrated that among all metals, Pd/PTA, Rh/PTA and Pt/PTA are most efficient catalysts. Fig. 4b shows the mechanism for the reduction of N2O by CO over M/PTA catalysts.38 Metals such as Co, Cu or Fe supported on ZrO2 are also active in the decomposition of N2O. The activity pattern is Co > Cu ≫ Fe. ZrO2 catalysts have the advantage of hydrothermal stability. Unlike that of zeolite systems, their catalytic activity is recovered when H2O vapours are eliminated from the feed gas.41

Tuti et al. further investigated N2O adsorption and decomposition on ZrO2. At 25 °C, N2O molecularly adsorbs through the O2 end on Zr4+ sites and dissociative adsorption of N2O occurs on surface defect sites, i.e. Zr3+ sites. It was noted that N–N bond dissociation did not occur and oxygen atoms of the lattice become incorporated into the product O2 molecules.42

In a different study, researchers have used a cluster model represented by M(OH)3(H2O)2, where M represents different transition metals such as Co, Fe and Rh. The results indicated that Rh and Co sites were more active for the N2O decomposition reaction as compared to Fe. The rate-limiting step is the formation of adsorbed O2 molecules via the interaction of adsorbed oxygen atoms with N2O. A correlation was observed between the activation energy (Ea) and the strength of the M–O bond. Weaker metal-adsorbed oxygen bonds for Co and Rh sites facilitated the decomposition of the second N2O molecule into O2, lowering the activation barrier.43 In the case of Fe(OH)3(H2O)2, the transition state analysis suggested that N2O dissociation was achieved through electron density donation from the metal to an N2O molecule.44 Various other examples of N2O decomposition metal oxide catalysts are provided in Table 1.

Table 1 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of metal oxide-based catalysts for N2O decomposition
Catalysts Method of preparation Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for the characterization and assessing the activity of catalysts Ref.
CaO catalysts Hydrothermal 1073–1273 K DFT method, on-line gas flue analyzer, GC 45
γ-Fe2O3 Co-precipitation N2 50 100–400 °C XRD, XPS, FTIR 46
Al-doped MoS2 Ethylene oxide DFT 47
3.0F–Co3O4 Sol–gel Ar 50 250–450 °C TEM, HRTEM 48
Ti3O6@TiO2 DFT 49
Ni > Co > Fe > Cu Impregnation N2 400 200–500 °C DFT 50
Co3O4 Sol–gel UV-visible spectra, H2-TPR, XPS, TEM, EXAFS, EDS, NAP-XPS 51
Sm0.1–Co3O4 Sol–gel Ar 50 300–500[thin space (1/6-em)]°C XRD, XPS, STEM, EELS, H2-TPR 52
N-doped Co3O4 Sol–gel H2/N2 30 150–500 °C FTIR, XRD, TEM, EDX, XPS, O2-TPD 53



5.1.1.1. Shortcomings of metal oxide catalysts. Though pure metal oxides show the highest catalytic efficiencies in the cases of transition metal oxides of group VIII and some rare earth oxides (e.g. La), it is also worthwhile to note that, based upon experimental conditions, metal oxides including MnO2, MnO, Cu2O and CoO are not stable and are partially decomposed.54 As a result, N2O decomposition studies took a turn towards the development of more active and stable catalytic systems. Another important fact about pure metal oxide catalysts is the utilization of high temperature for catalyzing N2O decomposition because of the limited availability of active sites at low temperatures due to adsorbed oxygen recombination. In addition, the presence of H2O exerts a negative effect on catalytic efficiency due to its competitive adsorption on active sites.21
5.1.1.2. Application scenario: nitric acid plants. Metal oxide catalysts such as Co3O4, Fe2O3 and MnO2 are low-cost, stable, and effective for N2O decomposition under high-temperature conditions typical of nitric acid plants. They offer thermal robustness and environmental compatibility but require elevated activation temperatures, show reduced efficiency at intermediate ranges and are vulnerable to deactivation by H2O and SO2 impurities.
5.1.2. Spinel oxides. Spinel oxides, having the structural formula of AB2O4, show high thermal stability and redox activity. In several redox reactions such as the reduction of N2O and oxidation of higher hydrocarbons and CO, cobalt spinel catalysts have been widely investigated.55 It has been demonstrated that Co3O4 is more active in CO oxidation than Fe2O3, Cu2O, NiO, MnO, V2O5, Cr2O, CeO2 and ZnO. Because of its capability to bind quickly with oxygen and a very low M–O bond energy, cobalt spinel is believed to display the highest activity in hydrocarbon oxidation.56 Fig. 5 represents the synthesis and use of various metal spinels in different electronic materials (ORR and OER).57
image file: d5ma00668f-f5.tif
Fig. 5 Synthesis routes and applications of spinel oxides (AB2X4).56 Reproduced with permission from the American Chemical Society (Copyright©2017).

Shen et al. investigated the catalytic conversion of N2O using cobalt oxide catalysts supported on various materials and synthesized using a co-precipitation method. Among the supports tested, alkaline MgO was found to be the most effective due to its strong electron-donating properties. Here cobalt existed in the form of Co3O4 nanoparticles dispersed in the MgO matrix. Under specific conditions, the Co–MgO-15% catalyst demonstrated good activity on decomposing N2O, making it a potential candidate for mitigating greenhouse gas emissions from anthropogenic sources.58 The advantages of using Co oxide-based spinel catalysts lie in their capability to decompose N2O below 400 °C and sufficient catalytic activity. Another benefit of using spinel oxide catalysts containing Rh is that they exhibit unusual catalytic activities in the absence of water but these catalysts rapidly lose their activity upon exposure to water.59,60

Researchers have tried to find whether the catalytic activity of cobalt spinel oxide catalysts is influenced by the addition of gases such as NO2, NO, O2 and H2O vapours or not. It has been found that operating the reaction at 850 °C led to a higher N2O conversion rate than operating the reaction at 800 °C.

This improvement was attributed to the presence of O2, which altered the oxidation state of Co in the catalyst.25 However, at lower temperatures, the presence of O2 reduced N2O conversion. H2O vapours also have a detrimental effect on the catalyst's activity at all temperatures. This was likely due to the competitive chemisorption of H2O vapours on the active sites, hindering N2O decomposition. Unlike O2 and H2O vapours, however NO2 did not decompose over the cobalt catalyst and it did not impact the catalyst's state. It had no effect on N2O conversion. However, the presence of NO in the feed gas mixture at 850 °C led to higher N2O reduction values. This was attributed to NO's influence on the process, possibly affecting the last phase of N2O decomposition. The state of catalyst was also influenced by temperature variations, particularly when exposed to a feed gas mixture containing O2, N2O and Ar at temperatures ranging from 700 °C to 850 °C.61 Extensive examples of the N2O decomposition by spinel oxide catalysts are enlisted in Table 2.

Table 2 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of different spinel oxide catalysts for N2O decomposition
Catalysts Methods Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Co/nHAP Hydrothermal Ar 200–600 °C XRD, TEM, HADDF–STEM, XPS 62
(Co3O4) (CS–N) Co-precipitation 100–400 °C XRD 63
Co3O4 Hydrothermal He 30 100–600 °C O2-TPD, DFT calculations, XRD, FTIR, SEM, XRF, TPSR 55
Co3O4-based catalyst Co-precipitation N2 50 400–600 °C XRF, XRD, H2-TPR, TEM, EDX, SAED 64
Cs-doped Co-spinel catalysts (Cs–Co3O4) Wetness impregnation <200 °C XRF, XRD, SEM, XPS, Raman, TPCR, QMS 65
deNOx and deN2O Precipitation 425–450 °C FTIR 66
CuxCo1−xCo2O4 spinel-oxide catalysts Co-precipitation N2 200 150–500 °C TGA, DTA, FTIR, XRD 67
Co3O4 PVA-assisted precipitation N2 65 300–600 °C DFT, XRD, XPS, TEM, SEM, FTIR, EPR 68
Co3O4 PVA-assisted precipitation 150–450 °C XRD 69
K-doped CuxCo3−xO4 catalysts Thermal decomposition O2 20 250–650 °C XRD, TGA, TPR, N2-physisorption, O2-TPD, ICP-OES, BET 70
Co–Mn–Al mixed oxide catalysts Precipitation H2 50 40–450 °C AAS, XRD, SEM, Raman, FTIR, TPR-H2, TPD-N2O, SEM, EDS, MS 71
Co-based spinel oxides Impregnation 50 25–400 °C XRD, BET, H2-TPR, O2-TPD, XPS 59
AgxCo (x = 0–0.08) oxide catalysts Co-precipitation Ar 30 200–400 °C XRD, TEM, O2-TPD, BET, H2-TPR, GC-TCD, XPS 72
Bi–Co3O4 catalysts Co-precipitation Ar 50 200–500 °C O2-TPD, BET, XRD, GC 73
Alien cations doped-nanocrystalline Co-spinel catalysts Co-precipitation He 30 100–600 °C XRD, Raman, BET, XPS, H2-TPR, TEM 74
LT-deN2O catalysts (K/Zn0.4Co2.6O4/αAl2O3) Co-precipitation 405–445 °C FTIR using on-line analyzer 75
ZnxCo1−xCo2o4 spinel catalyst Co-precipitation O2 10 XRD, EDX 76
Co3O4–Cs Wet impregnation H2 50 100–400 °C XPS, SEM, H2-TPR, XRD 77
Cd–Co3O4 spinel catalysts Co-precipitation TGA, DTA, XRD, FTIR, N2-physisorption, AAS 78
N-doped Co3O4 Sol–gel method H2/N2 30 200–500 °C XRD, FTIR, Raman, N2-physisorption, TEM, EDX, HADDF–STEM, H2-TPR, O2-TPD, EPR, CO2-TPD 53
(CuMgNiZn)1Co2O4 catalysts Co-precipitation H2 50 150–500 °C DFT, TPSR 79
Mn–Fe spinel catalysts Co-precipitation H2 100 100–400 °C NH3-TPD, NO-TPD, DRIFTS, FTIR, DFT 80



5.1.2.1. Shortcomings of spinel oxide catalysts. Spinel oxide catalysts often suffer from limited surface area and poor dispersion of active sites, reducing their catalytic efficiency. Their high-temperature synthesis can lead to particle sintering, decreasing reactivity. Additionally, some spinels exhibit slow oxygen mobility and reduced activity at low temperatures, limiting their applicability under mild reaction conditions. Stability under hydrothermal conditions can also be a concern.
5.1.2.2. Application scenario: industrial flue gas treatment. Spinel oxides are attractive for N2O decomposition in industrial flue gas streams due to their structural stability, redox flexibility and cost-effectiveness. They operate efficiently at high temperatures and tolerate thermal stress, but limitations include moderate activity at lower temperatures and susceptibility to poisoning by SO2 and H2O commonly present in flue gases.
5.1.3. Hydroxyapatites. Hydroxyapatites, having the formula Ca10(PO4)6(OH)2, act as dual catalysts in acid-base and redox catalysis. They enable moderate activity for N2O decomposition. Their flexible structure allows ion substitution (e.g. with transition metals), enhancing catalytic sites. They exhibit good thermal stability and surface hydroxylation but generally require modification for high activity. Their tenability makes them suitable for multifunctional or hybrid catalytic systems under mild conditions.81

In the 1990s, hydroxyapatite catalysts (containing calcium and phosphate ions) were employed for indirect N2O decomposition via partial oxidation of methane.82,83 However, during the last few years, hydroxyapatites have been used as supports for Rh and Ru catalysts in N2O abatement.84,85 Galloni et al. evaluated Cu- and Fe-loaded hydroxyapatite catalysts for N2O decomposition, revealing superior Cu performance due to nanoparticle formation, with detailed structural, stability and resistance analyses supporting catalytic behaviour.86 Wei et al. obtained natural hydroxyapatite from bone, and after doping it with Co ions, Co/nHAP catalysts were prepared and utilized for the investigation of N2O decomposition reaction.62 Fig. 6a shows the mechanistic pathway of N2O reduction over a Co/nHAP catalyst, while Fig. 6b indicates N2O conversion profiles with respect to the increasing temperature. Similarly, Tan et al. synthesized hydroxyapatite-supported bimetallic (Fe, Rh) catalysts for plasma-assisted N2O decomposition studies. In this instance, reaction temperature was lowered to a higher degree as free radicals generated by plasma initiated the catalytic decomposition reaction faster even at low temperatures.87 Fig. 6c and d indicate the mechanism and conversion profiles of N2O. More examples are represented in Table 3.


image file: d5ma00668f-f6.tif
Fig. 6 (a) Break down of N2O over the surface of Co-doped bone-derived hydroxyapatite (Co/nHAP) catalysts. (b) N2O decomposition profile.62 Reproduced with permission from Elsevier (Copyright©2020). (c) Mechanism of plasma-assisted N2O degradation over the surface of bimetallic RhFe/HAP catalysts. (d) Variation in % N2O decomposition w.r.t. temperature on RhFe/HAP and Fe/HAP catalysts.87 Reproduced with permission from Springer-Nature (Copyright©2023).
Table 3 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of hydroxyapatite-based catalysts for N2O decomposition
Catalysts Method Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Hydroxyapatite-supported RhOx catalysts (RhOx-HAP) Wet impregnation O2 30 150–400 °C XRD, TEM, ICP, XPS, BET, CO2-TPD, O2-TPD, GC, FT-IR 88
RuOx/M−P−O catalysts Impregnating H2/He 10 150–400 °C ICP–OES, XRD, BET, TEM, XPS, CO2-TPD, O2-TPD, H2-TPR, GC-TCD, QMS 84
Co/Hydroxyapatite Hydrothermal 100–900 °C XRD, TEM, HRTEM, EDX, STEM, UV-vis 89
Hydroxyapatite supported Rh, Fe, and Rh–Fe catalysts Impregnation He 60 150–200 °C XRD, SEM, TEM, TG, XPS, CO2-TPD 87
Hydroxyapatite (HAP, Ca10(PO4)6(OH)2) Hydrothermal NO 15 400–450 °C N2-physisorption, XRPD, NH3-titration, UV-DRS, Mössbauer, XPS, and EXAFS 90
Calcium hydroxyapatite (HAP, Ca10(PO4)6(OH)2) Co-precipitation 120–800 °C UV-DRS and Mössbauer spectroscopies, NH3 titration, N2-physisorption and XRPD 91
Apatite-supported Co3O4 Co-precipitation H2/N2 30 300–600 °C XRD, XPS, H2-TPR, TEM, Raman, FT-IR 92



5.1.3.1. Shortcomings of hydroxyapatite catalysts. These catalysts face several limitations including relatively low intrinsic activity compared to conventional metal oxides. Their performance heavily depends on metal ion doping, as pure HAP is largely inactive. Additionally, achieving uniform dispersion of active sites can be a challenge. They may also show reduced thermal stability under harsh conditions, and are susceptible to deactivation by sulphur or alkali contaminants without proper modification.
5.1.3.2. Application scenario: wastewater treatment facilities. These catalysts offer tunable surface properties, ion-exchange capacity and good thermal stability, making them promising for N2O decomposition in wastewater treatment off-gases. Their biocompatibility and low costs are additional advantages. However, they often show limited intrinsic activity, require modification with transition metals to enhance performance, and may suffer from deactivation under humid, impurity-rich gas conditions.
5.1.4. Mixed metal oxide catalysts. A lot of work has already been conducted on mixed metal oxide systems, e.g. metal-doped oxides and transition metal ions in inert matrices. Due to the unusual structure, thermal stability, synthesis at high temperature and low SSAs, the mixed metal oxides are famous for their N2O decomposition activity.93 Transition metal ions are very specific in their activity based on their oxidation state, e.g. MgO matrix-dispersed Mn(III) ions showed the highest catalytic efficiency as compared to Mn(II) and Mn(IV).94 However, the supported oxides possess better practical applications than pure and mixed metal oxides because of higher dispersion of metal ions in the large surface areas of different well-known supports including alumina, mesoporous silica, zirconia, ceria and titania. The catalytic efficiency is determined by metal loading, method of synthesis, and temperature.

For the purpose of efficient N2O abatement, mixed metal oxide catalysts have been proved to be very effective catalysts.95–98 Beyer et al. studied the N2O decomposition over different Rh-supported metal and non-metal oxides, e.g. Rh/SiO2, Rh/MgO, Rh/Al2O3, Rh/TiO2 and Rh/CeO2, in both the presence and the absence of O2. Rh/SiO2 and Rh/MgO showed high catalytic activity for N2O conversion, even at low temperatures in the presence of O2. In Rh/Al2O3, Rh/TiO2 and Rh/CeO2, smaller sized Rh nanoparticles were abundant, resulting in lower catalytic activities. The particle size of Rh was a crucial factor affecting the overall catalytic performance.99 The redox properties of active Rh components were affected by the acid-base properties of the support. The reduction abilities of Rh species decreased as the basicity of the support increased, indicating a strong interaction of Rh species with O2. Han et al. employed mesoporous TiO2@Fe2O3@Al2O3 core-shell nanostructures for low-temperature selective catalytic reduction (SCR) of NOx. It effectively prevented the deposition of sulphur compounds and accelerated the SCR of NOx by facilitating electron transfer at the Fe2O3–TiO2 interface. Fig. 7a depicts the comparison of mechanisms for the standard SCR with the fast SCR that these catalysts follow.100


image file: d5ma00668f-f7.tif
Fig. 7 (a) Mechanism of SO2-tolerant selective catalytic reduction of NOx with NH3 employing mesoporous TiO2@Fe2O3@Al2O3 monolith composites at low temperatures.80 Reproduced with permission from the American Chemical Society (Copyright©2019). (b) Mechanism of NO2 release via the combination of S-nitrocysteine with PbS quantum dots decorated TiO2 nanotubes.101 Reproduced with permission from American Chemical Society (Copyright©2011).

Similarly, Gaidei et al. also synthesized and compared N2O decomposition activities of various Rh-supported metal oxide catalysts, i.e. Rh/ZrO2, Rh/Al2O3 and Rh/Al2O3–SiO2. The catalysts were exposed to an intimal temperature of 450 °C in a reactor. As a result, they exhibited excellent dynamic characteristics to initiate and sustain the reaction. The catalyst's activity decreased over time. Unloaded catalysts achieved 100% decomposition of N2O within the temperature range of 460–480 °C that indicated its effectiveness in promoting decomposition reaction. X-ray analysis revealed the presence of phase transitions in carrier materials and oxidation of Rh during the operational time.102 In yet another study, Ratanatawanate et al. have devised a method that combined PbS QD-decorated TiO2 nanotubes with S-nitrosocysteine that released NO2 that further produced singlet oxygen.101 The whole setup for the attachment of PbS QDs with the surface of TiO2 nanotubes and release of NO2 is shown in Fig. 7b.

Imamura et al. investigated the impact of adding praseodymium (Pr) to CeO2 to investigate the catalytic decomposition of N2O. They prepared a catalyst by incorporating Rh on Pr/CeO2 composites. These catalysts were then subjected to calcination at different temperatures. However, the presence of Pr alone did not significantly affect the catalytic activity. A significant acceleration in the N2O decomposition rate was observed, as the calcination temperature of the catalysts was increased. Notably, the catalytic performance of the Rh-supported composite oxides (Pr–Rh/CeO2) calcined at 800 °C was found to be most effective.103 One of the chief advantages of Rh-doped ceria catalysts is that they retain an exceptionally high catalytic efficiency even in the presence of oxygen and water. Because of the hydrophobic nature of Rh/CeO2 catalysts, they are much less sensitive to moisture than the other catalysts.104 In another investigation, Peck et al. fabricated a series of CeO2-supported Fe2O3 and Co2O3 catalysts along with the increase in the metal loadings to determine monolayer surface coverage on the CeO2 support. It results in higher catalytic activities through the maximization of supporting metal–oxygen bonds. NOx was reduced by CO in the presence of sufficient amounts of O2.105 Fig. 8 represents the overall layout of NOx reduction by CO.


image file: d5ma00668f-f8.tif
Fig. 8 Structure and activity relationship of ceria-supported Fe and Co oxides for NO reduction by CO.105 Reproduced with permission from the American Chemical Society (Copyright©2017).

N2O decomposition was also tested on mixed metal oxides containing noble metals (i.e. Pt and Pd) supported on various oxide materials. The type of support material has a significant impact on the initial decomposition of N2O. Important support materials include SiO2 and Al2O3. At low temperatures (298 K), the Pt/SiO2–Al2O3 catalysts showed low activity compared to the other catalytic systems. However, their activity boosted at higher temperatures (573 K). This suggested that the support material's influence on N2O decomposition depends on the reaction temperature.106 Another important aspect of the support effect is change in the N2O conversion with pulse numbers. In most cases, N2O decomposition exhibited a steep decrease with the increase in pulse number. However, Pt/SiO2–Al2O3 catalysts showed exceptionally high N2O conversion rates at the second and third pulses.107

Rh-, Ru-, and Ir-based mixed metal oxides exhibited the highest activity in decomposing N2O. However these catalysts are susceptible to oxidation at high temperatures due to their oxidation potentials. Ru- and Ir-based metal oxide catalysts tend to form highly volatile oxides at elevated temperatures, leading to the loss of active metal components and a decrease in catalytic activity. Given its stability in the context of N2O decomposition, Rh is recommended as the preferred active component for catalysts used in this reaction.108 Table 4 presents various examples of mixed metal oxide catalysts utilized for N2O abatement studies.

Table 4 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of different mixed metal oxide catalysts for N2O decomposition
Catalysts Methods Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Cu/CeO2 Hydrothermal H2/Ar 25 300–450 °C STEM–EDX, STEM–EELS, H2-TPR, XPS, operando DRS-UV-Vis, DRIFTS, CO-DRIFTS 109
Rh/CeO2 Hydrothermal H2/Ar 40 400 °C CO-DRIFTS, N2O-DRIFTS, MCT detector, IR 110
RuO2/Al2O3 Wet impregnation N2 1000 25–300 °C BET surface area measurement, XRD, FT-IR, DBD reactor 111
Ru/Al2O3 Impregnation He 20 250–500 °C 112
PrBaCoCO3 Hydrothermal He 100 200–600 °C XPS, BET, H2-TPR, O2-TPR 25
Cu–Al–Ox mixed metal oxides Co-precipitation method He 100 300–450 °C XRD, ICP-MS, N2-physisorption, O2-TPD, H2-TPR, in situ FT-IR and XAFS 113
Metal oxides supported Au-NPs (Au/MxOy) (MxOy: Al2O3, CeO2, Fe2O3, TiO2 and ZnO) Deposition precipitation He 10 400–700 °C BET, XRD, SEM, HR-TEM, XPS, H2-TPR, EDS, GC-TCD 114
Co3O4/CeO2 mixed oxide catalysts Hydrothermal He 100 300–600 °C BET, XRD, TEM, H2-TPR, XPS, GC-TCD 115
K-modified Co–Mn–Al mixed oxides Hydrothermal He 50 250–450 °C ICP, XPRD, H2-TPR, CO2-TPD, NO-TPD, XPS, N2-sorption 116
Co3O4/ZrO2 Hydrothermal He 30 200–400 °C XRD, TEM, HRTEM, N2-physisorption, FTIR, XPS, O2-TPD, H2-TPR, GC, Raman 117
K/Y2O3–Co3O4 Co-precipitation He/Ar 20 100–400 °C BET, DTA 118
Nd(Cu,Co)Al–Ox Co-precipitation O2/He 100 50–600 °C XRD, BET, HRTEM, XPS, H2-TPR 119
Co3O4–LaCoO3 Co-precipitation He 40 700–850 °C XRD, BET, FTIR 120
MnxCo1−xCo2O4 Co-precipitation He 40 100–500 °C XRF, XRD 121
LaFeO3 Hydrothermal 250–500 °C XRD, EDX, XPS 122
MnOx/TiO2 Ultrasonic impregnation N2 1000 0–400 °C XRD, EDX, SEM 123
CuFeOx thin-film catalysts Adsorption Ar 20 100–550 °C XRD, EDX, XPS 124
MnO2/MOx (M = Al, Si and Ti) Wet impregnation He 60 120–280 °C XRD, XPS, BET 125
MnOx/TiO2 Wet impregnation 100–300 °C XRD, BET, HRTEM, XPS, H2-TPR and NH3-TPD 126
Pd/CeO2 Hydrothermal NH3/He 20 100–400 °C XRD, XPS, TEM, FEI 127
Pt/Co3O4 Incipient-wetness H2/Ar 35 100–500 °C XRD, SEM, TEM, TPO, TPR 128
MnOx/TiO2 Ultrasonic impregnation O2/Ar/NO/NH3 1000 0–350 °C XRD, SEM, XPS, BET 129
Cs-supported Co3O4 (Cs/Co) Co-precipitation Ar 100 100–300 °C XRD, XPS, H2-TPR 130
Bi0.1NiO1.15 Co-precipitation N2/O2 3100 300–400 °C XRD, XPS, H2-TPR 131
CuO/CeO2 Co-precipitation 300–550 °C STEM, EDX, XRD 132
Pd/γ-Al2O3 Impregnation N2 6 550–350 °C XRD, XPS 18
Rh/PTA     DFT 38
P@SiC Adsorption DFT, PDOS 133
(SACs) M1/PTA (M = Fe, Co, Mn, Ru, Rh, Os, Ir & Pt; PTA = [PW12O40]3−) Ion exchange DFT 134
Fe-substituted La-hexaaluminate supported Ir catalysts Microemulsion He 50 384–450 °C XRD, XPS, DRFTS 135
Rh/CeO2 Incipient wetness impregnation Ar 30 300–500 °C XRD, STEM, EDS, HRTEM, XPS, FTIR, DRIFTS 136



5.1.4.1. Shortcomings of mixed metal oxide catalysts. Mixed metal oxide catalysts, despite their tunable composition and synergistic effects, often suffer from limited control over active site uniformity and metal dispersion. Their synthesis can lead to phase separation or undesired crystallinity, reducing the catalytic efficiency. Additionally, some systems exhibit poor low-temperature activity and deactivation under long-term operation, especially in the presence of moisture or contaminants. Scalability and reproducibility also remain a challenge for complex multi-metal formulations.
5.1.4.2. Application scenario: nitric acid plant emissions. Mixed metal oxides are highly effective for N2O abatement in nitric acid plant emissions due to their synergistic redox properties, abundant oxygen vacancies and strong thermal stability. They are cost-efficient and tunable, but challenges include maintaining long-term stability under fluctuating gas conditions and vulnerability to poisoning by SO2 and water vapour.

5.2. Layered and framework structures

Layered and framework structure-based catalysts such as zeolites, hydrotalcites and mesoporous silica-supported systems offer unique structural advantages for N2O decomposition. Their well-defined pores and channels facilitate controlled diffusion and confinement of reactants, enhancing the selectivity and activity. Transition metal ions incorporated into frameworks (e.g. Fe–ZSM-5) act as isolated redox centres, promoting effective N–O bond cleavage.

Hydrotalcites provide tunable acid-base sites and high dispersion of metal species upon calcination. These materials also support ion-exchange capabilities, allowing precise modifications of active sites. However, their catalytic performance can depend heavily on pore architecture, metal loading and thermal stability under reaction conditions. Three classes of this category are given in this section.

5.2.1. Hydrotalcites (layered double hydroxides). With the structural formula of [M1−x2+Mx3+(OH)2]x+(An)x·yH2O, hydrotalcites (HTs) possess excellent anion exchange capacity and basicity.75 Their advantages include improved stability, control and efficiency compared to other reaction systems. Because of their flexible structure, hydrotalcite-derived transition metal oxides are considered very active and selective catalysts, even far more efficient than zeolitic catalysts. They are also thermally very stable. When using hydrotalcite catalysts, N2O decomposes below 500 K. Another benefit of calcined hydrotalcites is that they do not degrade at temperatures above 900 K, thus the high stability makes the hydrotalcite-derived catalysts quite promising for practical applications. Higher calcination temperatures combined with alkali promoters (i.e. K) have a great enhancing impact on the N2O decomposition activity. However, water and oxygen are strong inhibitors of the N2O conversion performance.137

The overall rate of N2O decomposition reaction depends on the adsorption of N2O molecules on the catalyst surface during the reaction. This adsorption is a crucial step in the decomposition process. The rapid formation of O2 led to the fast desorption of adsorbed oxygen, making active sites on the catalyst surface available for further adsorption of N2O. As a result of fast adsorption and desorption processes, the overall rate of N2O decomposition is higher. This led to the establishment of first-order rate equation w.r.t the concentration of N2O.138

Layered double hydroxides (LDHs) are successfully employed as starting materials for metal-supported catalysts. Generally, they are synthesized via three distinct routes: (1) with elements of redox properties present in between the layer spacing, direct calcination and/or reduction of LDH precursors is required, (2) within the LDH sheets, anionic exchange with the desired metal precursors is crucial, followed by calcination and/or reduction and (3) pre-calcined LDH precursors are impregnated with different inorganic materials followed by calcination and/or reduction.139 Fig. 9a summarizes all three routes for the synthesis of LDH-derived metal-supported catalysts.


image file: d5ma00668f-f9.tif
Fig. 9 (a) Representation of main routes leading to the synthesis of metal-supported catalysts from LDH precursors.139 Reproduced with permission from Springer (Copyright©2003). (b) N2O decomposition to N2 and O2 in the mirror plane phase of perovskite-type La-hexaaluminates (LaFexAl12−xO19).140 Reproduced with permission from the American Chemical Society (Copyright©2014). (c) Hydrotalcite-derived CuxMg3–xAlO oxides for pollutant-destruction mechanism.141 Reproduced with permission from the American Chemical Society (Copyright©2019).

Oxygen molecules tend to adsorb on the catalyst's surface and occupy the active sites, reducing the availability of free active sites for the N2O decomposition reaction. As a result, the overall N2O decomposition rate decreases. The Eley–Rideal reaction, that involves the desorption of oxygen, is not active enough to drive the overall N2O decomposition reaction within the temperature range of 250 °C and 500 °C. This behaviour of oxygen in N2O decomposition process is different from selective catalytic reduction (SCR) processes where oxygen plays a key role in breaking down organic molecules to produce N2 from nitrogen oxides.142 A first-order reaction rate model can be employed to describe the decomposition of N2O, but it is only applicable when the concentration of N2O is less than 10[thin space (1/6-em)]000 ppm. Beyond this range, the reaction kinetics deviate from the first-order behaviour. The presence of impurities in the feed gas can get adsorbed on the active sites of the catalysts, resulting in free defect sites on the surface, which can slowly decrease the rate of N2O decomposition.21 De Stefanis et al. used catalysts consisting of alumina-pillared smectites, which are a type of layered clay material. These clay materials were exchanged with transition metals.143 The main focus of this study was to reduce N2O via an eco-friendly decomposition reaction as follows:

N2O → N2 + 1/2O2
This reaction converts N2O into less harmful nitrogen (N2) and oxygen (O2). The results of this study showed that when pillared clays were exchanged with transition metals, the yield of decomposition reaction increased, suggesting that catalysts with transition metals are effective in promoting the N2O decomposition rate. It was noted that double-exchanged alumina-pillared montmorillonite/beidellite catalysts, exchanged with Ca and Co ions, exhibited the highest decomposition activity for N2O. Overall, the mechanism suggested that N2O is first adsorbed on the catalyst surface. Then, it underwent decomposition through catalytic oxidation and reduction.143 Similarly, Zhang et al. synthesized La-hexaaluminates (LaFexAl12−xO19) having abundant Al sites in the mirror plane (MP) phase, so that the degradation of N2O can be achieved efficiently. Fig. 9b demonstrates that Fe(III) ions prefer to occupy octahedral Al(III) sites in the MP of aluminate crystals confirming a high activity for N2O decomposition.140 Kiss et al. reported AlFe–PILC-type catalysts having different metal/OH ratios and employed them for N2O decomposition reactions.144 N2O was 100% converted into N2 and H2O on treatment with NH3 below 500 °C. However, in the direct conversion of N2O below 500 °C, the activity of catalysts did not exceed 40%. For the N2O abatement reaction, the catalytic activities of AlFe–PILC-type catalysts prepared with higher metal/OH ratios are higher than the apparent activity of AlFe–PILC-type catalysts having lower metal/OH ratios. It is attributed to the presence of free FeO–Fe2O3 particles.145

Hydrotalcites have been effectively utilized for dual decomposition of volatile organic compounds (VOCs) and gaseous pollutants. For instance, fabricated and calcined CuxMg3−xAlO were examined for 100% primary catalytic oxidation of n-butylamine and 83% secondary catalytic reduction of NOx at a lower temperature of 350 °C, as represented in Fig. 9c.141 In another study, catalytic decomposition of N2O was estimated at Co/Al hydrotalcites, [Co1−xAlx(OH)2[(CO3)]x/2·H2O] (where x ≈ 0.25–0.33), in a static glass reactor at a pressure of ≈50 torr within the temperature range of 150–280 °C. Catalysts exhibited a first-order dependence for the N2O decomposition activity. An increase in Co concentration led to an increased decomposition activity. A higher activity was shown by the catalyst precursor synthesized under a low saturation (LS) method than that synthesized by a sequential precipitation (SP) method.146 Table 5 presents various examples of hydrotalcites that have been applied for N2O decomposition.

Table 5 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of different hydrotalcite catalysts for N2O decomposition
Catalysts Preparation method Reaction atmosphere Air speed (mL min−1) Decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Mn(Fe)CoAlOx hydrotalcite Co-precipitation N2 100 100–500 °C XRD, H2-TPR, TGA, XPS, XAFS, BET 147
Ru/MexOy Hydrothermal 300–550 °C CO2-TPD, TEM 148
Rh/Mg/Al HT-derived film Co-precipitation H2/Ar 30 50–500 °C SEM, EDX, TEM, SAED, BET, Raman, H2-TPR, XPS 149
Hydrotalcite-derived CuxCoy (CuO–Co3O4 mixed oxides) Co-precipitation N2 100 350–500 °C XRD, XPS 150
Hydrotalcite-derived mixed oxides Ion exchange H2 200 300–500 °C XRD, XPS, BET 151
Hydrotalcite-derived Co3O4, Co3O4–P Sol–gel method 150–400 °C TEM, TGA, XRD and BET 152
Hydrotalcite-derived [y = Cu/(Cu + Ni)] Co-precipitation H2/Ar 41 600–900 °C XRD, XPS, TGA–DTA, BET, XPS 153
CoCuAl mixed oxides derived from hydrotalcites Solid phase H2/Ar 60 150–350 °C XRD, N2-physisorption, H2-TPR 154
Al@Cu-based hydrotalcite CuAl-HT-c and Cu-P Co-precipitation NO/He 8 450–600 K XRD, TPR and in situ XANES 155
CO/He 8



5.2.1.1. Shortcomings of hydrotalcite catalysts. Hydrotalcite catalysts, while offering tunable composition and high dispersion of active sites, suffer from several drawbacks. Their thermal stability is limited, as structural collapse can occur upon high-temperature calcination. The resulting mixed metal oxides may exhibit low surface area and poor crystallinity. Additionally, their activity in N2O decomposition is generally moderate, requiring further modifications. Sensitivity to water and contaminants can also affect the long-term performance and catalyst durability.
5.2.1.2. Application scenario: automotive exhaust control. These catalysts with layered double hydroxide structures offer high surface area, tunable composition, and excellent dispersion of active metal sites, making them promising for N2O decomposition in automotive exhaust systems. They are low cost and versatile but disadvantages include limited intrinsic activity, dependence on metal modification for efficiency and reduced stability under high-temperature, water and sulphur-rich exhaust conditions.
5.2.2. Zeolites with transition metal ions. Zeolite catalysts are among the few early known catalysts for N2O decomposition, while some show catalytic activities below 600 K.156 The activity of transition metal ion-exchanged zeolite was determined by the type of transition metal ion (TMI) and zeolite. This is the reason that the activity order for different transition metals in zeolite catalysts differs considerably from other metal oxide catalysts. One major advantage associated with zeolite catalysts is that the higher oxygen concentration shows a very small inhibitory effect on their catalytic efficiency, but one disadvantage associated with Rh-exchanged zeolites is their extreme sensitivity to water.60

Transition metal ions in zeolites, supported on oxides, or dispersed in solid solutions, are very active for N2O decomposition.157,158 Numerous zeolites with transition metal ions are extensively studied in heterogeneous catalysis. A zeolite's ion-exchange capacity (IEC) depends on how it is chemically composed. A higher IEC is perceived in the zeolites having low SiO2/Al2O3 ratios. Specific ion-exchange capacity of a zeolite varies by varying the structure of zeolites and cations exchanged. When a wet ion exchange (WIE) method was used for the preparation of Fe–ZSM-5, Lobree et al. were able to achieve only half of the complete IEC without getting small iron oxide particles.159

Natural zeolites show low activity for the catalytic reduction of N2O using NH3 due to the complex structures of zeolites.160 However, when natural zeolites were modified with Fe, their activity in the SCR of N2O became comparable to synthetic zeolites. It was found that the highest activity of N2O decomposition was achieved when natural zeolites contained 6.5 wt% Fe.

Natural zeolites prepared with FeCl2 as the precursor exhibited higher activity for the reduction of N2O due to the formation of more reducible Fe species with FeCl2.160 Zeolites prepared with FeSO4 as the precursor showed a considerably lower activity for the decomposition of N2O as surface sulphate ions were identified as inhibitors, and their presence shifted the reaction temperature to higher values. It is suggested that natural zeolites, especially when modified with Fe, can serve as a cost-effective alternative for catalysing the SCR of N2O, as they demonstrated comparable activity to synthetic zeolites.161,162

Most active catalysts for N2O decomposition are usually inhibited by NO. This suggests that NO can interfere with the process of breaking down N2O into less harmful components. However, interestingly enough, NO can significantly enhance the rate of N2O decomposition over a specific catalyst, Fe–ZSM-5. Pérez-Ramírez et al. were first to report this positive effect of NO on N2O decomposition over Fe–ZSM-5. They proposed a mechanism where NO in the gas phase combines with both N2O and adsorbed O2, leading to the regeneration of active sites and the production of nitrogen oxide. This mechanism explains the increase in the N2O decomposition rate due to NO.61 However, Li and Armor reported that zeolites exchanged with Co or Cu ions were effective catalysts for N2O decomposition, and that metal ions (particularly Co, Cu, and Fe) in ZSM-5 were far more active than the same ions on Al2O3.163 On Cu–ZSM-5, Fe–ZSM-5 and Co–ZSM-5, the N2O decomposition rate was proportional to the exchange extent. The existence of excess O2 left the catalytic activity of Co–ZSM-5 unchanged and slightly decreased Cu–ZSM-5 activity, but enhanced the catalytic activity of Fe–ZSM-5.164 Conversely, H2O present in the feed strongly affected the catalytic activity of zeolites and led to irreversible deactivation because of its heterolytic adsorption on active sites. In the presence of water vapours, Co–ZSM-5 was more stable than Cu–ZSM-5 and Fe–ZSM-5.165 For the decomposition of N2O into its elements, TMI-exchanged pentasil-zeolites are recognized as most appropriate catalysts. Among the samples investigated, the highest N2O decomposition activity was displayed by Cu–ZSM-5 catalysts.166 Therefore, Deka et al. have carried out an extensive investigation on the use of Cu-exchanged zeolite-based catalysts (that include zeolite Y, ZSM-5, SSZ-13 and zeolite beta) for the abatement of NOx present in the exhaust emissions of light and heavy-duty vehicle engines.11 However, Cu–ZSM-5 has a tendency to lose activity in the presence of H2O vapours limiting its use in practical applications. However, Fe-containing zeolites are less sensitive towards H2O vapours in comparison to Cu-containing zeolite.167 Liu et al. studied N2O formation using Cu–SSZ-13 zeolites via the NH3-asssited SCR of NOx. All reactions were performed in a flow reactor system, as shown in Fig. 10. It was found out that N2O formation occurred through two entirely different mechanisms at low and high temperatures. At low temperatures, N2O was released due to the decomposition of NH4NO3 on Cu–SSZ-13 zeolite, while high temperatures favoured the NH3 oxidation route for N2O formation.168


image file: d5ma00668f-f10.tif
Fig. 10 Experimental setup of N2O formation during the selective catalytic reduction of NOx with NH3 over Cu–SSZ-13.168 Reproduced with permission from the American Chemical Society (Copyright©2019).

Yasumura et al. demonstrated an efficient strategy for the dispersion of metal species to design single-atom catalysts. Chabazite (CHA) zeolite precursors were utilized as templates to disperse bulk Pd atoms into their small pores.169 At 600 °C, the flow of feed gas was regulated in such a way that Pd atoms present on the outside of CHA zeolites can get dispersed introducing Pd(II) ions into the pores of zeolite forming Pd–CHA zeolites along with the formation of N2O. Fig. 11a illustrates the conversion of bulk Pd and CHA zeolites into Pd(II)-intercalated zeolites.


image file: d5ma00668f-f11.tif
Fig. 11 (a) Transformation of bulk Pd to Pd cations in small-pore CHA zeolites facilitated by NO.169 Reproduced with permission from the American Chemical Society (Copyright©2021). (b) Decomposition of N2O (%) via selective catalytic reduction (SCR) through NH3 on Fe-MOR catalysts. (c) Decomposition of NH3 (%) by oxidation of NH3 with and without N2O. (d) Decomposition of N2O (%) via NH3 as a function of P(NH3) with and without O2 and e. TPSR profiles of oxidation of NH3 (conditions: 100 mg catalyst, 0 or 4000 ppm NH3, 0 or 5% O2, 5000 ppm N2O, and balance He. GHSV = 30[thin space (1/6-em)]000 h−1).170 Reproduced with permission from the American Chemical Society (Copyright©2012).

It has been observed that adsorbed NH3 leaves the active sites by reaction with the surface O2. As NH3, as a reducing agent, preferably reacts with O2, the N2O decomposition process becomes suppressed due to the presence of surface O2. As a result, a vast majority of studies on the SCR of N2O have been devoted to Fe-exchanged zeolite catalytic systems, more specifically, the Fe–BEA catalysts.171,172 The N2O abatement by NH3 follows the Mars and Van Krevelen mechanism that involves an oxidation–reduction cycle of interchange between two oxidation states of Fe (Fe3+↔ Fe2+). In this scenario, N2O reduction into oxygen surface species occurs over some definite Fe-sites, which lead to simultaneous N2 release. Consequently, O2 has no inhibitory influence on N2O decomposition through NH3.173 In Fig. 11b, it can be noted that the presence of O2 does not hinder the conversion of N2O. In fact, there was small improvement in N2O conversion when O2 was present. This suggested that O2 did not negatively impact the conversion of N2O. Fig. 11d demonstrates that the enhanced reduction of N2O was associated with the oxidation of NH3, which could change the NH3/N2O stoichiometry. In the absence of O2, the highest N2O conversion occurred at around 3500 ppm of NH3 concentration. In contrast, in the excess of O2, the highest N2O removal occurred at around 4000 ppm of NH3 concentration. This also suggested that O2 did not significantly interfere with N2O decomposition and did not efficiently react with NH3. Fig. 11c indicates that the oxidation of NH3 by O2 was much less significant when N2O was present compared to when N2O was absent. This suggested that in the reaction system involving N2O, NH3 and O2, the selective catalytic reduction (SCR) of N2O by NH3 was more dominant than the oxidation of NH3 by O2. Fig. 11e shows that when NH3 is oxidized by O2, it produces a large amount of NO, NO2 and N2O. However, in the N2O + NH3 + O2 reaction system, only N2 is detected as the main product. This indicated that NH3 can be oxidized by both O2 and N2O, but N2O is much more reactive than O2, and their reaction pathways are distinct.170

Ju et al. have found that UV irradiation of the Pb2+–ZSM-5 catalyst led to the decomposition of N2O at 298 K in a linear manner with the irradiation time.174 It can be confirmed by observing the yield of N2 that increases as a function of the UV-exposure time, while under dark conditions, the formation could not be noticed. The yield of N2 also increases when the amount of Pb2+ loading becomes higher. It was reported that the photocatalytic decomposition of N2O proceeds much more efficiently in the presence of propane.130 In addition to the N2 evolution, this reaction also led to the formation of oxygen-containing compounds such as ethanol and acetone. It was indicated that the efficient decomposition of N2O proceeds photocatalytically in the presence of propane on the Pb2+–ZSM-5 catalysts. The efficiency of the photocatalytic reduction of N2O is found to be strongly dependent on the type of hydrocarbons used, e.g. methane or ethane, and among the hydrocarbons, the highest enhancement in the reaction rate is shown by propane. These results can be attributed to the observation that during the reaction, oxygen atoms formed by the photodecomposition of N2O detach more easily from the isolated Pb2+ ions in the presence of propane.174 Similarly, Co-based BEA zeolite catalysts were also used to study the oxidative dehydrogenation of propane (C3H8) with N2O or a mixture of N2O and oxygen (O2) as the oxidants. N2O is more selective but less active compared to O2 as an oxidizing agent. Increasing the concentration of N2O in the reaction mixture with C3H8 and O2 led to a substantial increase in propane conversion while maintaining a constant selectivity for propene. The combination of O2 and N2O had a synergistic effect, resulting in a three-fold higher yield of propene compared using O2 and N2O alone.175 Table 6 summarizes various examples of catalysts used along with the techniques applied during the course of analysis for the purpose of N2O decomposition.

Table 6 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of different zeolite catalysts for N2O decomposition
Catalysts Preparation method Reaction atmosphere Air speed (mL min−1) Decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Co–, Cu– and Fe–SSZ-13 zeolites Hydrothermal N2 10 300–530 K XRD, FT-IR, H2-TPR, N2O-TPR 27
H2/Ar 30
Rh–ZSM catalysts Incipient wetness He/Ar 70 250–450 °C STEM, XPS, TPR, N2O-DRIFTS 104
CoMnAl HTC catalysts Co-precipitation He 25 330–460 °C SEM, EDX, XRD, NH3-TPD, CO2-TPD, H2-TPR, BET 157
Cu–ZSM-5 catalysts Ion exchange NH3/He 30 200–600 °C BET, XRD, TEM, CO2-TPR, NH3-TPR, GC–TCD 176
Fe–ZSM-5 Hydrothermal He/Ar 15 550–750 K N2O-TPD, O2-TPD, NO–TPD, AAS 177
Co–ZSM-5 Hydrothermal He/O2 25 380–500 °C SEM, XRD, NH3-TPD, H2-TPR, ICP–AES, UV-Vis-DRS, DFT 178
Co/zeolites (Co/Beta, Co/mordenite, Co/ZSM-5, Co/MCM-49, Co/ZSM-23, Co/SSZ-13) Wet impregnation N2 30 200–450 °C XRD, BET, NH3-TPD, H2-TPR, XPS UV-Vis-DRS, 179
Fe–, Co–, Ni– and Cu–ZSM-5 catalysts Ion exchange Ar 4 400–800 °C DFT, TPR 158
Fe, Co and Ni-exchanged Na-MOR catalysts (Fe-MOR, Co-MOR and Ni-MOR catalysts) Ion exchange O2/He 100 293–773 K GC, TCD, FID, operando FTIR equipped with MCT detector 180
Cu/ZSM-5 catalysts Adsorption N2 40 250–500 °C ICP–AES, UV-Vis-DRS, H2-TPR, N2O-TPD 181
Cu–ZSM-5 catalysts Adsorption DFT 182
Fe-zeolites (Fe–PST-7, Fe–LTA, Fe–RTH, Fe–SSZ-13, Fe–FER, Fe–ZSM-5) Hydrothermal He 55 600–850 °C XRD, SEM, TEM, BET, UV-vis, ESR, FTIR, NMR 183
Cu–ZSM-5 catalysts Hydrothermal 600–800 K H2-TPR, IR 184
Cu–ZSM-5 catalysts Ion exchange N2O/N2 300 350–450 °C H2-TPR, IR, BET, H2-TPR, TPD, FESEM, EDX 151
Cu–, Ni–, Zn–, Co–, Fe– and Mn–ZSM-5 catalysts Ion exchange N2 300 50–300 °C DFT, N2O-TPD, GC–TCD, FT-IR, BET, ICP–OES 185
Fe–ZSM-5 catalysts Wet impregnation N2 30 250–550 °C XRD, XPS, BET, HRTEM, UV-Vis-DRS, XPS, H2-TPR, TPO, DRIFTS 186
Cu/HZSM-5 catalysts Wet impregnation N2 60 620 K FTIR, XPS, XPRD 187
Co/HZSM-5 catalysts Citric acid impregnation O2/Ar 80 300–450 °C XRD, TG–DSC, Raman, SEM, H2-TPR, XPS, EPR, FTIR, DRIFTS 188
Fe–ZSM-5 catalysts Hydrothermal N2 20 300–600 °C XRD, SEM, BET, UV-vis, FTIR 189
Fe–ZSM-5, Fe-Beta, Fe–FER catalysts Wet impregnation He 70 200–400 °C FTIR, BET, H2-TPR, TGA–MS 190
Fe-beta, Fe–ZSM-5 and Fe–SAPO-34 catalysts Ion exchange H2/Ar 30 200–400 °C UV-vis-DRS, H2-TPR 191
Fe–SSZ-13 catalysts Ion exchange He 103 400–600 °C ICP–AES, BET, UV-vis-DRS, XRD, TEM, On-line GC–TCD 192
Fe/CHA catalysts Ion exchange N2 80 160–550 °C XRD, ICP–OES, DFT, TEM, EDX, XPS, EPR, FTIR, H2-TPR 193
Fe-modified MCM-22 zeolite and its derivative ITQ-2 catalysts Ion exchange Ar 10 200–600 °C XRD, ICP–OES, UV-vis-DRS, H2-TPR, HRTEM 161
(Co, Fe)–ZSM-5 and (Co, Fe)-beta catalysts Ion exchange Ar 30 25–600 °C XRD, H2-TPR, UV-vis, TCD 165
Cu/SSZ-13, Cu/SSZ-5, Cu/BEA Ion exchange O2 2 100–500 °C XRD, BET, DRIFTS, FTIR, H2-TPR, NH3-TPD 194
Co(x)/PRSA catalysts Impregnation O2 20 348–438 °C XRD, BET, H2-TPR, NH3-TPD, XPS, XRF, DRIFTS 195
Fe-FER catalysts impregnation NO/O2/H2O 350 200–900 °C XRD, BET, SEM, UV-vis-NIR, FTIR, Mössbauer spectroscopy 196
Fe zeolite catalysts Diffusion impregnation He 20 UV-Vis-NIR, FTIR, Mössbauer spectroscopy 197
Zeolite 5A, 13X and ZSM-5 catalysts Wet impregnation N2 300 298–358 K N2O-TPD, BET, isosteric heat of adsorption 198
Cu/SSZ-13 catalysts Hydrothermal NO/NH3/CO2/H2O/NO2/N2O 800–1500 200–600 °C XRD, BET, ICP–OES, UV-Vis-DRS, H2-TPR, HRTEM 199
Fe/BEA Ion exchange N2 300 350–380 °C XRF 200
Fe-MOR and Fe-FER Hydrothermal O2/He 100 0–350 °C FTIR, UV-vis 201
Fe–FER Hydrothermal 500–900 °C XRD, SEM, FTIR 202
Zeolite-based Fe catalysts Ion exchange NO/O2/H2O/He 350 400–800 °C XRD, FTIR, TPD 203



5.2.2.1. Shortcomings of zeolite catalysts. Zeolite catalysts, despite their well-defined microporous structures and strong metal–support interactions, have notable limitations in N2O decomposition. Their narrow pore sizes can restrict the diffusion of reactants and products, especially at higher conversions. Metal ion exchange may lead to non-uniform dispersion or clustering, reducing active site accessibility. Thermal stability under prolonged high-temperature reactions can be a concern, leading to framework degradation. Additionally, zeolites are sensitive to water vapours and contaminants such as sulphur, which can cause deactivation and limit their industrial applicability.
5.2.2.2. Application scenario: diesel vehicle exhaust mechanism. These catalysts are highly effective for N2O decomposition in diesel exhaust due to their high surface area, tunable acidity and excellent hydrothermal stability. They enable selective catalytic reduction with NH3, offering high efficiency at medium temperatures. However, they are sensitive to sulphur poisoning, may deactivate under prolonged hydrothermal stress, and require careful formulation for durability.
5.2.3. Mixed metal oxides in mesoporous silica. After the discovery of well-organized mesoporous molecular sieves, scientific community has developed interest in their catalytic applications. Discovered by Mobil researchers, these mesoporous silica-supported catalysts possess a higher surface area than the conventional zeolite-based catalysts. Owing to their fascinating properties, these materials play an important role to support the other co-catalysts in many catalytic reactions.204

For example, SBA-15 is a mesoporous silica material with a fibre-like morphology that is several tens of micrometers long. It possesses a 2D porous structure and unique properties.205,206 It has various applications including its use as a support material for Rh catalysts in N2O abatement with Rh being one of the most active metals in N2O decomposition. Different mesoporous silica-supported metal ions are crucial with regard to their N2O abatement activity and age resistance, but among them Rh/SBA-15–S is found to be most active and durable. It is well known because of its superior performance in the presence of O2 which is typically an inhibiting agent. This is attributed to the large pore size of Rh/SBA-15–S that enhances catalyst's access, diffusion and dispersion. The smaller particle size of Rh in Rh/SBA-15–S results in the higher dispersion of metals on the catalyst surface. It favours the formation of more Rh+ ions than Rh0 or Rh3+, which, in turn, leads to higher N2O decomposition activity. Rh/SBA-15–S is found to be durable, maintaining its performance at a temperature of 430 °C. This suggests its suitability as a support material in long-term applications and further research.207 Similarly, Wei et al. reported the synthesis of a highly dispersed and well-homogenized bimetallic Au–Pd alloy supported on amine-functionalized SBA-15. The catalytic activities, when tested from 150 °C to 600 °C, revealed a decreasing trend with the increase in the percentage of Au in bimetallic alloy catalysts (Fig. 12a). It led to the conclusion that the addition of Au negatively affects the activity of Pd catalysts. Though on the one hand, the Au–Pd alloy facilitates the desorption of O2 molecules, on the other hand, it affects other elementary steps that change the rate-determining step. Fig. 12b depicts the energy profiles of N2O degradation over Pd, Pd3Au/Pd and PdAu3/Pd catalyst's surface, which has been split into three processes, i.e. (a) decomposition of N2O into N2 and O atoms, (b) diffusion of O atoms on the surface of catalysts and (c) recombination of two O atoms to form O2. The respective mechanisms of degradation are shown in Fig. 12c–k.208 Various other examples of such catalysts are summarized in Table 7.


image file: d5ma00668f-f12.tif
Fig. 12 (a) N2O conversion (%) as a function of reaction temperature onto AuxPd1−x/SBA-15 catalysts (where x = atomic ratio of Au to total metal atoms; metal loading = 2 wt%). (b) Energy profiles for N2O decomposition on Pd, AuPd3 and Au3Pd catalysts. (c)–(k) Mechanism of N2O degradation into N2 and O2 over (c)–(e) Pd, (f)–(h) Pd3Au/Pd and (i)–(k) PdAu3/Pd catalysts.209 Reproduced with permission from the American Chemical Society (Copyright©2012).
Table 7 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of mesoporous silica-supported metal/metal oxide catalysts for N2O decomposition
Catalysts Method Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Al2O3 + SiO2 Wet impregnation N2 100 300–450 °C AAS, XRD, TPR-H2, FTIR, Raman, XPS, SEM, N2 adsorption 210
Pt, Pd, Rh/Al2O3–SiO2 Hydrothermal He 100 600–900 °C XRD, N2 physisorption and FESEM 211
Pt/SiO2 Impregnation Ar 10 350–600 °C BET, XRD, ICP 212
Cu-supported on hollow silica–alumina composite Impregnation N2 50 450–800 K EDX, UV-Vis-DRS, SEM, TEM 213
Rh@S-1 Hydrothermal 200–400 °C STEM, XPS, DRIFTS, TEM 214
NiCoAl-based monolithic catalysts Wet impregnation N2 10 300–600[thin space (1/6-em)]°C SEM, EDX, BET 215
TiO2@Fe2O3@Al2O3 monolith catalysts Co-precipitation 220–420 °C XRD, XPS 100



5.2.3.1. Shortcomings of mesoporous silica-supported metal oxide catalysts. Mesoporous silica-supported metal oxide catalysts often suffer from weak metal-support interactions, leading to metal sintering or leaching under harsh conditions. Their thermal and hydrothermal stability is generally lower than that of crystalline supports like zeolites. Additionally, achieving uniform metal dispersion can be a challenge, and their catalytic activity may decline over time due to the structural degradation or active site deactivation.
5.2.3.2. Application scenario: chemical plant emissions. Mixed metal oxides dispersed in mesoporous silica supports provide high surface area, uniform metal dispersion and enhanced redox properties, making them effective for N2O decomposition in chemical plant emissions. They offer thermal stability and tenability, but disadvantages include susceptibility to pore blockage, reduced activity under humid or sulphur-rich conditions, and higher synthesis complexity.

5.3. Nanostructured metal-doped materials

Nanostructured metal-doped materials exhibit high surface area, quantum size effects, and unique electronic properties. Doping enhances the catalytic activity by modifying charge distribution and creating active sites. Their tunable morphology – nanosheets, nanotubes or cages – improves reactant accessibility. These materials often show superior thermal stability, reactivity, and selectivity, making them promising candidates for advanced catalytic applications like N2O decomposition.
5.3.1. Metal-doped nanotubes, nanosheets and nanocages. Nanotubes, nanosheets and nanocages are elongated, 2D and cage-like structures, respectively, possessing large surface area, crystallinity, high porosity, hollow structure and astonishing electrical and mechanical properties, rendering them novel and excellent choice of catalyst for N2O decomposition due to a shorter diffusion length and easier mass/heat transport.209,216 Recently, the decomposition of N2O supported by metal-doped nanotubes (NTs), nanosheets (NSs) and nanocages (NCs)-has emerged. Researchers have predicted the roles of metal-confined titania nanotubes,217,218

carbon nanotubes (CNTs),219,220 boron nitride nanotubes, nanosheets,221,222 and nanocages223 for N2O decomposition by employing DFT studies.

Metal-doped nanocages are currently among very popular materials that can capture greenhouse gases and protect the environment from their harmful effects. Gao et al. have successfully doped four Ca atoms in a C60 molecule and examined its energy profile for CO2 capture and N2O adsorption. In comparison, the adsorption of N2O on pristine C60 molecules was very weak as compared with Ca-decorated fullerene C60.224 Fig. 13a represents the doping of Ca atoms in a C60 molecule, while Fig. 13b and c show the adsorption of one and two N2O molecules on the surface of CaC60, respectively. Table 8 demonstrates the examples of metal-doped nanotubes, nanosheets and nanocages utilized for N2O mitigation studies.


image file: d5ma00668f-f13.tif
Fig. 13 (a) Four Ca atoms doped C60 nanocage; adsorption profiles and corresponding energies of (b) one and (c) two N2O molecules on fullerene (CaC60).225 Reproduced with permission from the American Chemical Society (Copyright©2011); (d) mechanism of methane-to-methanol conversion by transition-metal-atom-embedded N-doped graphene (TM–N4/C) catalysts with N2O and O2 oxidants and a comparison of their reactivity.226 Reproduced with permission from the American Chemical Society (Copyright©2023). (e) Energy bands of TiO2/MXene/g-C3N4 displaying energy bending.227 Reproduced with permission form Elsevier (Copyright©2023).
Table 8 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of metal-doped nanotubes, nanosheets and nanocages for N2O decomposition
Catalysts Method Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Metal oxide-confined interweaved TiO2 nanotubes M/TNT (M = Mn, Cu, Ce, Fe, V, Cr and Co) Wet impregnation 100–300 °C BET, HRTEM, XRD, XPS, H2-TPR, NH3–TPD 228
C-doped BN nanotubes (C-BNNTs) Adsorption DFT calculations for N2O adsorption and reduction reaction 229
C-doped BN nanosheets Adsorption DFT calculations, computational calculations 222
Si-doped C3N (Si–C3N) nanosheets Adsorption First-principle (DFT) calculations 230
Fullerene-like boron nitride nanocages (C-doped B12N12) Adsorption DFT calculation of adsorption behaviors of N2O and CO 231
Boron-doped C3N monolayers Adsorption DFT 232
A0.5Co2.5O4 (A = Mg, Ca, Sr, Ba) Hydrothermal O2/N2 80 DFT 233
Fe-embedded C2N monolayer DFT 234
SACs supported on defective boron nitride nanotubes (BNNT) DFT 235
Potassium promoted Gd0.06Co catalysts Co-precipitation Ar 50 150–500 °C XPS, H2-TPR, O2–TPD, XRD, SEM, TEM 236
Gd to Co3O4 Co-precipitation N2 350 DFT 237
Ni–Co mixed oxides (NixCo1−xCo2O4) 250–600 °C SEM, EDX, XRD, 238
CuO–Co3O4 Co-precipitation NH3/O2/He 100 400–700 °C XAFS, DRIFTS, DFT, XANES 239
Si (Si-doped haeck-BNNT) Adsorption O2 550 DFT 240
MgO(100) in SO2 and CO Adsorption DFT 241
Rh(0.2)/NC–MnFe Adsorption H2 60 100–150 °C XRD 242
MC23 clusters (M[thin space (1/6-em)] = [thin space (1/6-em)]Ru, Mn, V, Rh, and Pd) clusters DFT 243
Al- or P-doped hexagonal boron nitride nanosheets (h-BNNS) Adsorption DFT 244
Si@BN-yne DFT, PDOS 245
Co-doped Fe–Mn@CNTs Co-precipitation 170–250 °C XRD, XPS 246
Sm0.06Ni Adsorption Ar 50 100–600 °C SEM, TEM, XRD, XPS, HRTEM 247
Co3O4@CoMn2O4 Hydrothermal NH3/O2/SO2/H2O/Ar 300 90–270 °C SEM, TEM, HRTEM, XPS, DRIFTS 248
M13@Cu42 (M = Cu, Co, Ni, Zn, Ru, Rh, Pd, Pt) core–shell clusters Adsorption DFT 249
Si@C24N24 DFT 250



5.3.1.1. Shortcomings of metal-doped nanotubes, nanosheets and nanocages. Controlled synthesis of metal-confined nanotubes remains a big issue. They face challenges like structural instability at high temperatures, metal agglomeration and limited long-term durability under reaction conditions. Their synthesis can be complex and costly and scalability remains a barrier. In low-temperature zones, these catalysts suffer from serious disadvantages at the industrial level.251
5.3.1.2. Application scenario: power plant emissions. Metal-doped nanostructures offer exceptional surface area, abundant active sites, and tunable electronic properties, making them highly promising for N2O decomposition in power plant emissions. They provide excellent dispersion of doped metals and improved redox performance. However, challenges include complex synthesis, potential agglomeration, limited large-scale stability, and higher costs compared to conventional catalyst systems.
5.3.2. Metal-embedded graphene and graphitic nitrides. Graphene is considered one of the most promising candidates for next-generation electronic materials of this decade.225,252 Extensive research is carried out on graphene because of its enormous surface-to-volume ratio (i.e. large area) for catalytic reactions. However, modified graphene sheets display high catalytic performance for N2O decomposition, and several DFT studies have been conducted on this catalytic behaviour, e.g. transition metal-embedded graphene,253–255 N-doped graphene,256,257 Si-doped graphene,245,258 metal oxide-doped graphene,255 and metal-decorated graphene oxides.253,259 Among them, transition metal-embedded graphene structures possess good thermal stabilities and stronger bonds between metals and neighbouring C-atoms.

Like graphene, graphitic carbon nitride (g-C3N4) is composed of carbon and nitrogen atoms arranged in a 2D, hexagonal lattice. However, in g-C3N4, carbon and nitrogen atoms are bonded together in a polymeric, layered structural form. This polymeric arrangement gives it a unique set of properties compared to graphene. However, graphene is a zero-bandgap material while g-C3N4 is a semiconductor with moderate bandgap that makes it suitable for electronics and photocatalytic applications especially under visible light. Furthermore, its properties can be tuned by modifying its structure, composition and doping that optimize its performance for specific applications.260–262 For example, Troppovà et al. modified its structure by making its nanocomposites with TiO2 at different weight ratios. The primary focus of the study was to assess the photocatalytic activity of these TiO2/g-C3N4 nanocomposites. They investigated the decomposition of N2O under UV treatment with a wavelength (λ) of 365 nm. The results indicated that all the synthesized nanocomposites exhibited higher photocatalytic activity than the individual parent materials (g-C3N4 and TiO2).263 The enhanced photocatalytic activity of TiO2/g-C3N4 nanocomposites was attributed to the coupling of TiO2 with g-C3N4. This coupling was found to improve the separation of electron–hole pairs and expand the range of spectral light absorption. Similarly, N-doped graphene embedded with single transition metal atoms produces single-atom catalysts (SACs), and enable the conversion of CH4 into CH3OH with the help of N2O and O2 oxidants (Fig. 13d). The relationship between the adsorption energy of oxygen molecules (ΔEAds) and the catalytic reactivity in the conversion of methane to methanol can be seen in the inset.264 Many other examples of use of such catalysts in N2O abatement are provided in Table 9.

Table 9 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of metal-embedded graphene, graphitic oxides and nitrides for N2O decomposition
Catalysts Method Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
BCmNn-doped graphene sheets Adsorption DFT calculations, Computational studies on adsorption configurations, electronic structures, and catalytic activity by quantum chemical calculations 257
Single Si or P atom incorporated N-doped graphene(SiN4-Gr/PN4-Gr) Comparative DFT study, computational calculations for adsorption energies and reduction mechanism by first-principle 265
Single Ge or Ga atom-doped graphene Adsorption DFT study, computational estimation of adsorption energies and reaction mechanisms 266
ZnO-doped graphene [ZnO(001)@G] DFT calculation for adsorption and activation energies, ab initio model, OAT, ER and LH pathway 267
Al- and Ti-doped graphene (Al-/Ti-graphene) DFT calculations for reaction barriers and thermodynamic parameters 268
M-decorated graphene oxide (M = Mg, Cu or Ag) Adsorption DFT computational method 259
Graphitic carbon nitride (g-C3N4) Melamine condensation N2 15 400–700 °C UV-Vis-DRS, PL, FTIR, Raman, SSA, XRD, SEM, TEM 269
TiO2/g-C3N4 photocatalysts Adsorption N2 30 XRPD, UV-Vis-DRS, Raman, FTIR, SEM, TEM, BET, PL, XPS, GC-BID 270
TiO2/g-C3N4 photocatalysts Hydrothermal O2 3.5–4.5 N2-physisorption, XRD, UV-Vis-DRS, FTIR, Raman, PL, TEM, XPS 263
g-C3N4/WO3 photocatalysts Calcination XRD, XPS, HR–TEM, STEM, N2-adsorption, UV-Vis-DRS, PL, EDS 271
g-C3N4/ZnO nanocomposite photocatalysts Co-precipitation XRD, UV-Vis, FTIR, Raman Spectroscopy, TEM and AAS 272
g-C3N4/BiVO4 composite photocatalysts Sol–gel N2 500 XRD, N2-physisorption, UV-Vis-DRS, PL, FTIR, XPS, TEM 273
BiVO4/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]1) composite photocatalysts Adsorption XPS, UV-Vis-DRS, PL, FTIR, Raman, BET, TEM 274
g-C3N4 and BiOIO3 Hydrothermal 300–800 °C FTIR, XRPD, XPS, HRTEM, DRS, TEM 275
SiN4G DFT 276
Cu-Embedded Graphene Adsorption DFT 253
Si@Graphene Adsorption   100–350 °C DFT 258
Au/C2N and Cu/C2N Adsorption DFT 277
Si-doped C3N (Si–C3N) nanosheets Co-adsorption DFT 278
Co3O4/g-CN Wet impregnation N2 5 100–700 °C XRD, HR–TEM, N2-physisorption, Raman, FT-IR, XPS, H2-TPR, CO2-TPD, EC 279



5.3.2.1. Shortcomings of metal-embedded graphene and graphitic oxides/nitrides. Major disadvantages associated with these types of catalysts are their low surface area and a rapid recombination between photoexcited electrons and holes. However, this problem is overcome by impregnation or doping of a noble metal or co-catalyst into these semiconductor materials.226,280
5.3.2.2. Application scenario: industrial flue gas treatment. Metal-embedded graphene and graphitic oxides/nitrides provide high conductivity, large surface area, and strong metal-support interactions, enhancing N2O decomposition in industrial flue gas treatment. They enable efficient electron transfer and dispersion of active sites. However, their disadvantages include high synthesis costs, sensitivity to oxidation at elevated temperatures, and scalability issues for long-term industrial deployment.
5.3.3. MXene-based materials. MXene-based catalysts have emerged as promising materials for N2O decomposition due to their unique 2D structure, high electrical conductivity and tunable surface chemistry. Doping transition metals (e.g. Fe, Co, and Cu) into MXenes enhances redox activity and introduces abundant active sites for N2O adsorption and activation. Their layered morphology facilitates gas diffusion and charge transfer, which are critical for catalytic efficiency. However, stability under high-temperature and oxidative conditions remains a concern. Additionally, scalable synthesis and control over metal dispersion require further development. Despite these challenges, MXene-based catalysts offer a novel platform for designing efficient and selective N2O abatement systems.281

As of now, there are very limited experimental studies reporting the use of MXene-based catalysts for N2O abatement. However, a few studies have been conducted for their N2O decomposition efficiency by mixing MXenes with other semiconductor photocatalysts. For example, Zhao et al. prepared triphase heterostructures via the self-assembly of titania, MXene and graphitic carbon nitride (TiO2/MXene/g-C3N4) that showed an improved photocatalytic efficiency for NO removal. Z-scheme heterojunctions enhance photocatalysis by preserving strong redox carriers and minimizing recombination of low-activity electrons and holes, as shown in Fig. 13e.282 Other studies on photocatalytic decomposition of NOx using MXenes coupled with other materials are listed in Table 10.

Table 10 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of MXene-based catalysts for N2O decomposition
Catalysts Method Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
UO2@MXnene (e.g. Ti2CO2, V2CO2, Cr2CO2, Zr2CO2, Nb2CO2, Mo2CO2) ab initio calculations 283
Y-doped Ti2CO2 (MXene) monolayer DFT and ab initio calculations 284
P@Cr2CO2 and P@MoCO2 DFT 285
g-C3N4/TiO2/Ti3C2 MXene Air oxidation and ultrasonic processing O2 3000 XRD, SEM, TEM, HRTEM, EDS, XPS, PL, BET, EPR 286
TiO2–N–Ti3C2Tx Co-precipitation TEM, SEM, FTIR, Raman, XPS, PL, EPR 287
(Ti,C)-BiOBr/Ti3C2Tx Solvothermal N2 200 XRD, FESEM, STEM, DRS, XPS, BET, Raman 288
Ti3C2–OH with TiO2 Co-precipitation N2 SEM, TEM, XRD, XPS 289
Cu/Ti3C2O2-V0 and Ni/Ti3C2O2–Ti1 DFT 290


5.4. Porous MOFs and MOF-derived catalysts

MOFs act as porous solid material catalysts possessing coordinatively unsaturated sites (CUS) for gas adsorption, separation and catalysis.291 Their well-defined composition, hollow structures, small particle size, and exposed surface impart considerable efficiency and stability to these catalysts. MOFs are also utilized as precursors for the synthesis of metal oxides that can prevent the nanoparticle's clustering during the calcination.227

Metal–organic frameworks are composed of metal atoms or metal oxide clusters coordinated to organic linkers. This arrangement creates a highly ordered, porous network with a large surface area. The choice of metal and linkers can be tailored to achieve specific applications. MOFs are well known for their structural flexibility, meaning that their frameworks can be tuned or modified to suit particular needs. This flexibility allows for the design of customized surface and porous structures. Coordinative unsaturated metal sites are of particular interest, as these sites have metal atoms with unoccupied coordination, making them highly reactive. CUS can be crucial in gas adsorption and separation processes as well as catalytic reactions, which is why MOFs are excellent candidates for gas adsorption and separation due to their high surface area and tunable porosity. The tunable nature of MOFs generate catalytic sites with precise geometries, enhancing the catalytic efficiency and selectivity.292,293 For example, nanoscale MOFs may exhibit different properties compared to their bulk counterparts, making them suitable for their specific dynamic applications.294,295

Despite the huge number of reports on gas sorption, the N2O abatement studies by MOFs are still scarce. The common route of N2O abatement is through catalytic decomposition, but adsorptive removal of N2O by MOFs has also been reported in cases when catalytic decomposition is not economically feasible due to very low N2O concentrations.156 In some studies, MOFs also offer an important means for dissociative adsorption of N2O,296–300 as displayed in Fig. 14.


image file: d5ma00668f-f14.tif
Fig. 14 Layout of the removal of atmospheric pollutants (that affect respiratory tract) via adsorption through MOFs.

Vogiatzis et al. conducted a computational study on high-spin Fe(IV)–OXO complexes that have well-known activating effects on the strong C–H bonds.301 Although it is difficult to stabilize a high-spin electronic configuration for homogeneous catalysis in molecular species, the results suggested that it is attainable in MOF-based catalysts, e.g. Fe2(dobdc) and its Mg-substituted analogues. Using a screening approach, different frameworks were identified that could form Fe(IV)–OXO complexes on N2O adsorption. It was concluded from the study that among different materials, Fe–BTT selectively oxidizes C2H6 to C2H5OH by N2O adsorption. The follow-up spectroscopic techniques suggested that the observed reactivity was attributed to the existing defect sites.301 Further examples from the recent studies are given in Table 11.

Table 11 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of MOFs and MOF-derived catalysts for N2O decomposition
Catalysts Method Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
M3(BTC)2 MOF [M = Cr, Fe, Co, Ni, Cu and Zn] Wet impregnation DFT calculations with M06-L functional, EXAFS, computational screening 302
MOF-derived Ni/CeO2 HPOC Wet impregnation Ar 10 250–500 °C SEM, TEM, N2-adsorption, EDS, XRD, H2-TPR, Raman, XPS, ICP–MS, GC–TCD 297
Fe-supported MOF Zr-NU-1000 Wet impregnation DFT and CASPT2 calculations with M06-L functional 303
MnO2 decorated Ru-MOF (Ru-NH2-UiO-66/MnO2) Hydrothermal Not clear XRD, XPS, BET, SEM, TEM, N2O–TPD, MS, EXAFS, IC, DFT 304
MFU-4l MOFs (MFU-4l, CuI-MFU-4l and Li-MFU-4l) (BTDD2− ligands and [Zn5Cl4]6+ building units) Wet impregnation N2 3 Not clear DFT calculations, PXRD, FTIR, TGA, ICP–OES, BET 305
Fe(II)(Me3TACN)(S2SiMe2) Hydrothermal 100–135 °C UV-vis, resonance Raman, EPR, Mössbauer, XAS, DFT 306
ZIF-67 derived Co/CoOx@C Hydrothermal N2 30 305–650 °C TPD, DRIFTS 307
Mn@Cu3(BTC)2 Ion exchange 230–260 °C SEM, TEM, XRD, XPS, FESEM, ZEISS-ΣIGMA HD, BET 308


5.4.1. Shortcomings of MOFs and MOF-derived catalysts. MOFs and MOF-derived catalysts face challenges such as limited thermal stability, structural degradation under reaction conditions, and low conductivity, which can hinder electron transfer during N2O decomposition. Additionally, their synthesis can be complex and costly and active site accessibility may be restricted, affecting the overall catalytic efficiency.
5.4.2. Application scenario: waste gas purification. MOFs and MOF-derived catalysts offer ultra-high surface area, tunable porosity, and flexible metal-ligand coordination, making them highly effective for N2O decomposition in waste gas purification. Their derived metal oxides or carbons show strong redox activity and stability. However, drawbacks include complex synthesis, limited hydrothermal and thermal stability and challenges in scaling up for industrial applications.

5.5. Emerging or novel materials

Antenna-reactor catalysts and quantum dot-based catalysts exhibit unique light-harvesting and charge separation abilities, crucial for N2O decomposition. Further details are provided in the coming sections.
5.5.1. Antenna-reactor catalysts. Very recently, the concept of “Antenna–Reactor” plasmonic photocatalysts is introduced, and it is rapidly gaining interest in the field of catalysis. In a few studies conducted to date, N2O decomposition was selected as a model reaction.138,309 The use of antenna-reactor photocatalysts holds special significance in controlling the rate and products of various chemical reactions. Being poor absorbers of light, many traditional catalysts lack the ability to catalyze efficiently because of the restricted surface chemistry. By combining a catalytic nanoparticle with a plasmonic nanoantenna in a single complex, the light-harvesting properties of plasmonic nanoparticles can be fine-tuned to increase light absorption.310 Table 12 summarizes few examples of these catalysts.
Table 12 Overview of synthesis methods, reaction conditions and various techniques employed during the course of analysis of antenna-reactor catalysts for N2O decomposition
Catalysts Method Reaction atmosphere Air speed (mL min−1) N2O decomposition temp-range (°C K−1) Techniques used for characterization and activity of catalysts Ref.
Al−Ir antenna−reactor photocatalysts Ion exchange 250–350 °C HAADF–STEM, MS, XPS and UV-Vis-NIR 138
Antenna-reactor plasmonic photocatalysts (Pd decorated Al NCs, Al@Cu2O and Cu–Ru surface alloy@Cu) Co-precipitation XPS, ICP–MS, UV-Vis-DRS, TEM, HAADF,GC-MS, HR–TEM 309


5.5.2. Quantum dots. Quantum dots (QDs) can play a significant role in N2O decomposition due to their tunable bandgaps, high surface-to-volume ratios, and excellent light absorption properties. These features enable efficient electron–hole pair generation and separation under visible light, promoting redox reactions. QDs can be engineered to provide active sites for N2O adsorption and activation, potentially lowering the activation energy and enhancing catalytic efficiency under mild, sustainable reaction conditions.311 Currently, there are no experimental studies specifically reporting the use of QDs for N2O decomposition. However, theoretical investigations have predicted their potential.

6. Performance-based summary of N2O decomposition catalysts

Performance-based comparisons are critical for identifying optimal catalysts and guiding their future research directions. Table 13 provides a comparative summary of the benchmark performance metrics for different classes of catalysts used in N2O decomposition, based on their catalytic activity, temperature range, hydrothermal stability, and resistance to inhibitors. This provides a clear performance-based comparison of N2O decomposition catalysts.
Table 13 Comparison of the benchmark performances of N2O decomposition catalysts
Catalyst class Typical active sites T50a (°C) Hydrothermal stability Resistance to H2O/O2b Remarks
a T50 (°C): temperature at which 50% N2O conversion is achieved.b Resistance to H2O/O2: resistance to inhibition by water vapour or oxygen.
Metal oxides CuO, Co3O4, MnO2 300–500 Moderate Low moderate Inexpensive; deactivated by moisture
Spinel oxides CoFe2O4, MnFe2O4 300–450 High Moderate Good stability; better than simple oxides
Hydroxyapatites Ca10(PO4)6(OH)2 doped with Fe, Cu >450 Moderate Poor High thermal stability; lower activity
Mixed metal oxides Cu–Mn, Fe–Al oxides 250–400 Moderate Low – moderate Synergistic effects; still limited by deactivation
Hydrotalcites Mg–Al–Fe–Cu layered structures 300–500 Moderate Moderate Tunable; activity depends on calcination
Zeolites Fe, Co, Cu in ZSM-5, SSZ-13 200–400 High High Fe–SSZ-13 shows best performance under wet/oxidizing conditions
Mixed metal oxides in mesoporous silica Fe/Cu in SBA-15, MCM-41 250–450 Moderate – high Moderate Good dispersion and surface area
Metal-doped nanotubes/nanosheets Fe, Co, Cu on CNTs, g-C3N4, BN 300–500 Moderate Low – moderate High surface are; expensive synthesis
Graphene and graphitic nitrides Single atoms or clusters (Fe, Co) 250–450 Low – moderate Low – moderate Tunable electronic structure; prone to sintering
MXene-based materials Ti3C2, Mo2TiC2 doped with metals 250–400 Moderate Moderate Emerging materials with potential for hybrid photocatalysis
MOFs and MOF-derived catalysts Fe-MIL-100, ZIF-8-derived oxides 300–500 Low – moderate Poor High porosity; often thermally unstable
Antenna-reactor catalysts Dual-site systems (e.g. Ce, Fe) 200–350 Moderate Moderate – high Emerging concept; synergistic enhancement
Quantum dots-based catalysts CdS, Fe3O4 QDs supported on hosts <300 (predicted) Unknown Unknown Theoretical potential; lacks direct experimental validation


7. Mechanisms of N2O decomposition

There are five primary mechanisms of N2O decomposition in the context of fundamental steps and catalyst surface interactions. These mechanisms are key to understanding how various catalysts function in breaking down N2O into harmless components (Fig. 15).
image file: d5ma00668f-f15.tif
Fig. 15 Different mechanisms of N2O decomposition.

7.1. Direct thermal decomposition


N2O → N2 + ½ O2
Direct thermal decomposition of nitrous oxide occurs at temperatures above 800 °C in the absence of catalysts. Although the reaction is thermodynamically favourable, the extreme energy demand makes this route impractical for most industrial applications. However, in certain specialized high-concentration processes such as semiconductor manufacturing, direct pyrolysis of N2O has been investigated under nitrogen-rich atmospheres at temperatures approaching 1000 °C. These studies demonstrate that while N2O can indeed be decomposed under such conditions, the process requires significant thermal input and careful gas-phase control, which limits its scalability for broader emission abatement. Nonetheless, this pathway remains relevant in niche applications where high-purity nitrogen or oxygen streams are simultaneously desired.312

7.2. Surface-catalysed decomposition

Most practical catalysts (e.g. Fe-, Co- and Cu-based) accelerate this reaction by activating N2O on their surfaces. It requires active oxygen-vacancy sites. Catalysts can be used to lower the activation energy for N–O bond cleavage. Overall, it is summarized in three steps:313,314

1. Adsorption

N2O(g) → N2O(ads)

2. N–O bond cleavage

N2O → N2(g) + O(ads)

3. Recombination of surface oxygen

O(ads) + O(ads) → O2(g)

7.3. Redox mechanism (Mars–van Krevelen type)

It is commonly observed in transition metal oxide catalysts. Here, the catalyst surface is first reduced and then deoxidized, enabling a catalytic cycle:315,316

1. In the first step, N2O reacts with a lattice oxygen.

N2O + [M]–O → N2 + [M]–O–O

2. Oxygen vacancy is replenished by O2 from the gas phase.

[M] + ½O2 → [M]–O

7.4. Single atom or atomically dispersed catalysis

This type of catalysis is common in Fe–ZSM-5, Fe–SSZ-13 and Cu-CHA materials. During this catalysis, first N2O coordinates to a single metal centre. Then, the N–O bond is cleaved through electron transfer from the metal. Oxygen binds to the metal site or adjacent framework oxygen. It shows high site selectivity and stability, especially in the presence of steam or O2.

7.5. Photocatalytic decomposition

It involves photon-excitation of catalysts such as TiO2, g-C3N4 or MXene hybrids. In this case, light generates electron–hole pairs in the photocatalysts. Electrons are responsible for the reduction of N2O to N2. The holes may oxidize adsorbed species or water to maintain charge neutrality. This mechanism is still under development particularly for low-temperature N2O removal.

8. N2O activation mechanisms

N2O activation mechanisms follow different pathways, three of which are most common, namely dual-pathway mechanism, radical-initiated oxidative decomposition and electron-assisted surface decomposition. Representative reactions belonging to each category are indicated in this section.

8.1. Dual-pathway mechanism

It corresponds to the classical surface-mediated activation of H2O on catalytic active sites (*), described as follows:
N2O + * → N2 + *O

2*O → O2 + 2*

N2O + *O → N2 + O2 + *
When NH3 is present (NH3-assisted SCR), surface O2 can be removed as follows:
2NH3 + 3*O → N2 + 3H2O + 3*
Interaction of NH3 with N2O is given as
3N2O + 2NH3 → 4N2 + 3H2O
The Fe–MOR catalyst was employed for the systematic investigation of N2O reduction reaction mechanism by NH3, through which the roles of different oxygen species were examined317 (Fig. 16a).

image file: d5ma00668f-f16.tif
Fig. 16 Different mechanisms of N2O activation.

8.2. Radical-mediated oxidation of N2O

In this case, N2O activation occurs via the following main reaction routes (Fig. 16b):
N2O → NO + N

NO + O → NO2

NO + OH → HNO2

NO2 + OH → HNO3

8.3. Electron-mediated reduction of N2O

The solid solution system is fairly active toward the decomposition reaction of N2O in the temperature range of 300–450 °C. Winter proposed the mechanism of N2O decomposition reaction involving the following steps318 (Fig. 16c):
2N2O → N2 + 2O2
 
N2O(g) → N2O(ads) (1)

N2O(ads) + e → N2O(ads)

N2O(ads) → N2(ads) + O(ads)
 
O(ads) + O(ads) → O2(g) + 2e (2)

O + N2O → O2(g) + N2(g) + 2e
In governing the reaction rate, N2O adsorption [step 1] and oxygen (Oads) desorption as O2 [step 2] are the key steps. Using the following rate equations, the kinetic parameters of N2O decomposition have been derived:319
 
image file: d5ma00668f-t1.tif(3)

Here, α = P° N2O (i.e. initial pressure of N2O), x = Pt N2O (which represents the pressure of N2O at time ‘t’), and x/2 = PtO2 (i.e. pressure of oxygen released at any time ‘t’).

This rate equation corresponds to the non-inhibitory effect caused by O2. The rate eqn (3) corresponds to strong inhibition by O2 and represents weak inhibition of the decomposition kinetics by O2.320

9. Conclusions and future perspectives

In this review, we summarized various categories of N2O decomposition catalysts. Starting with the well-known category of TMI-exchanged zeolites, we then presented other different classes including hydrotalcites, spinel oxides, mixed metal oxides, mesoporous silica-supported metal oxides and pure metal oxides, and finally, the more recent categories of catalysts such as metal-doped nanotubes, nanosheets and nanocages, graphene and graphitic oxides/nitrides, MOFs and MOF-derived catalysts for adsorptive reduction of N2O. Nitrous oxide (N2O) is a potent greenhouse gas and a major contributor to stratospheric ozone depletion. Its anthropogenic and natural emissions through biological processes such as nitrification, denitrification and nitrifier denitrification along with abiotic pathways and industrial activities demand urgent and effective mitigation strategies. Catalytic decomposition of N2O into benign products (N2 and O2) stands out as a promising solution owing to its direct applicability in emission abatement technologies and potential integration into existing industrial setups.

This review comprehensively categorizes N2O decomposition catalysts based on their mechanisms, compositions and material structures, providing a unified understanding of their functionality and efficiency. Mechanistically, the catalysts are classified into those following direct catalytic decomposition (DCD), selective catalytic reduction (SCR) and other redox or surface-catalysed processes. Composition-wise, both precious metal-based catalysts (e.g. Rh, Pt, and Pd) and non-precious alternatives (e.g. Fe, Co, and Cu-based materials) have been investigated with emphasis on cost, stability and activity trade-offs. The type of support materials – ranging from metal oxides, spinels to mesoporous and layered structures – plays a vital role in modulating dispersion, redox behaviour and oxygen vacancy dynamics.

Importantly, this review offers a progress-based classification, moving from classical oxide-based catalysts to novel nanostructured materials and framework materials such as MOFs and MOF-derived catalysts, which have shown significant promise due to their high surface areas, tunable porosity and ability to accommodate multiple active sites. Emerging materials such as antenna-reactor systems and quantum dots illustrate the frontier of N2O decomposition research, hinting at the exciting potential of photonic and nanoscale control over activation energies.

Mechanistic pathways such as surface-catalysed decomposition, Mars–van Krevelen mechanisms, and single-atom catalysis reveal the intricate nature of N2O activation, where lattice oxygen, metal oxidation states and coordination environments dictate the reaction rate and selectivity. Novel mechanisms including photocatalytic and electron/radical-mediated pathways further expand the toolkit available for N2O conversion, particularly under mild conditions or light-driven environments.

Looking ahead, several key future perspectives emerge. First, the rational design of catalysts with high activity at low temperatures coupled with selectivity and durability under industrial gas compositions remains an urgent priority. The integration of computational modelling and machine learning can accelerate the discovery of optimal catalyst compositions and structures. Second, a shift towards earth-abundant, non-toxic elements is essential to scale these technologies sustainably. Single-atom catalysts and dual-function catalysts that combine decomposition with selective oxidation or reduction reactions open new frontiers in multifunctional reactor design.

Moreover, in situ and operando characterization techniques should be leveraged to elucidate real-time reaction mechanisms, bridging the gap between theoretical models and actual catalyst performance. Finally, interdisciplinary collaboration between catalysis, materials science, environmental engineering and process design is essential to translate laboratory successes into commercially viable solutions.

In summary, catalytic N2O decomposition is a vibrant and evolving field. Continued research on catalyst development, mechanistic understanding and scalable application will be pivotal in mitigating N2O emissions and achieving global climate and sustainability goals.

Conflicts of interest

Authors declare no competing financials interest.

Abbreviations

WHOWorld Health Organization
EPAEnvironmental Protection Agency
IPPCIntergovernmental Panel on Climate Change
AFEDArab Forum for Environment and Development
APANAsian Pacific Adaptation Network
CCCCCCaribbean Community Climate Change Centre
CANClimate Action Network
CCLCitizen Climate Lobby
CJAClimate Justice Alliance
EDFEnvironmental Defense Fund
HEAHealth and Environmental Alliance
NRDCNatural Resources Defense Council
GWPGlobal warming potential
AOBAmmonia-oxidizing bacteria
AOAAmmonia-oxidizing archaea
DCDDirect catalytic decomposition
SCRSelective catalytic reduction
ZSMZeolite Socony Mobil
SSZAluminosilicate zeolite
MOFMetal–organic framework
EaActivation energy
CFBCCirculating fluidized bed combustion
SNCRSelective non-catalytic reduction
EGEthylene glycol
DPDeposition precipitation
IMPImpregnation method
DP–ClDeposition precipitation by chloride precursor
PTAPhosphotungstic acid
SACsSingle-atom catalysts
DFTDensity functional theory
GCGas chromatography
XRDX-ray diffraction
XPSX-ray photoelectron spectroscopy
FTIRFourier transform infrared spectroscopy
TEMTransmission electron microscopy
HRTEMHigh-resolution transmission electron microscopy
TPRTemperature-programmed reduction
H2–TPRTemperature-programmed reduction with H2
XPSX-ray photoelectron spectroscopy
EXAFSExtended X-ray absorption fine structure
EDSEnergy-dispersive spectroscopy
STEMScanning transmission electron spectroscopy
EELSElectron energy loss spectroscopy
O2–TPRTemperature-programmed reduction with O2
nHAPNatural hydroxyapatites
SSASpecific surface area
HADDFHigh-angle annular dark field
O2-TPDTemperature-programmed desorption of chemisorbed O2
XRFX-ray fluorescence
TPSRTemperature-programmed surface reaction
SEMScanning electron microscopy
FESEMField emission scanning electron microscopy
SAEDSelected area electron diffraction
TPCRTransfer-PCR
QMSQuadrupole mass spectrometer
EPRElectron paramagnetic resonance
TGAThermogravimetric analysis
ICP–OESInductively coupled plasma-optical emission spectrometry
BETBrunauer, Emmett and Teller theory
MSMass spectrometry
GC–TCDGas chromatograph equipped with a thermal conductivity detector
DTADifferential thermal analysis
XRPDX-ray powder diffraction
QDsQuantum dots
DRSDiffuse reflectance spectroscopy
DRIFTSDiffuse reflectance infrared Fourier transform spectroscopy
CO-DRIFTSDiffuse reflectance infrared Fourier transform spectroscopy of CO adsorption
MCTMercury cadmium telluride
LDHsLayered double hydroxides
MPMirror plane
PILCPillared clay
VOCsVolatile organic compounds
LSLow saturation
SPSequential precipitation
TMIsTransition metal ions
IECIon exchange capacity
WEIWet ion exchange
XANESX-ray absorption near-edge structure
CHAChabazite
AASAtomic absorption spectroscopy
ICP–AESInductive coupled plasma atomic emission spectroscopy
FIDFlame ionization detector
ESRElectron spin resonance
NMRNuclear magnetic resonance
SBASanta Barbara Amorphous
NTsNanotubes
NSsNanosheets
NCsNanocages
CNTsCarbon nanotubes
TNTsTitania nanotubes
PLPhotoluminescence
CUSCoordinatively unsaturated sites
BDC1,4-Benzenedicarboxylate
BTC1,3,5-Benzenetricarboxylate
MCMMobil Composition of Matter
KITKorea Advanced Institute of Science and Technology
MCFMesoporous Cellular Foam
Fe-FERIron Ferrierite
TG–DTA–MSThermogravimetric differential thermal analysis mass spectrometry

Data availability

The data and necessary protocols of this study have been included as part of the SI.

Acknowledgements

This work was financially supported and was provided access to the international journals and financial resources by the Higher Education Commission (HEC) of Pakistan (no. 377/IPFP-II) (Batch-I/) SRGP/NAHE/HEC/2020/. Dr. Ejaz Hussain thanks Govt. of Pakistan, WHO and EPA for arranging consistant seminars, conferences, debates and attention for Globle Warming.

References

  1. F. Charlson, S. Ali, T. Benmarhnia, M. Pearl, A. Massazza, J. Augustinavicius and J. G. Scott, Int. J. Environ. Res. Public Health, 2021, 18, 4486 CrossRef PubMed.
  2. S. Sinhababu, Y. Lakliang and N. P. Mankad, Dalton Trans., 2022, 51, 6129–6147 RSC.
  3. K. Hayashi and N. Itsubo, Int. J. Life Cycle Assess., 2023, 1–13 Search PubMed.
  4. N. J. Bassous, A. C. Rodriguez, C. I. L. Leal, H. Y. Jung, C. K. Lee, S. Joo, S. Kim, C. Yun, M. G. Hahm and M. H. Ahn, Adv. Sens. Res., 2023, 2300094 Search PubMed.
  5. R. Feng and X. Fang, Environ. Sci. Technol., 2022, 56, 5299–5301 CrossRef CAS PubMed.
  6. S. He, Y. Han and X. Qin, J. Environ. Sci., 2025, 157, 615–632 CrossRef CAS PubMed.
  7. B. K. Sovacool, S. Griffiths, J. Kim and M. Bazilian, Renewable Sustainable Energy Rev., 2021, 141, 110759 CrossRef CAS.
  8. X. Zhu, M. Burger, T. A. Doane and W. R. Horwath, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 6328–6333 CrossRef CAS PubMed.
  9. N. Wrage, G. L. Velthof, M. L. Van Beusichem and O. Oenema, Soil Biol. Biochem., 2001, 33, 1723–1732 CrossRef CAS.
  10. T. C. Robinson, D. E. Latta, L. Notini, K. E. Schilling and M. M. Scherer, Environ. Sci.: Processes Impacts, 2021, 23, 1531–1541 RSC.
  11. U. Deka, I. Lezcano-Gonzalez, B. M. Weckhuysen and A. M. Beale, ACS Catal., 2013, 3, 413–427 CrossRef CAS.
  12. Y. Zhang, G. Cao and X. Yang, Energy Fuels, 2021, 35, 6443–6464 CrossRef CAS.
  13. Y. Liu, F. Gao, H. Yi, C. Yang, R. Zhang, Y. Zhou and X. Tang, Environ. Sci. Pollut. Res., 2021, 28, 2549–2571 CrossRef CAS PubMed.
  14. H. Chen, L. Zeng, D. Wang, Y. Zhou and X. Yang, Water Res., 2020, 184, 116168 CrossRef CAS PubMed.
  15. M. Galle, D. Agar and O. Watzenberger, Chem. Eng. Sci., 2001, 56, 1587–1595 CrossRef CAS.
  16. A. E. Hughes, N. Haque, S. A. Northey and S. Giddey, Resources, 2021, 10, 93 CrossRef.
  17. Q. Zhang, S. Gao and J. Yu, Chem. Rev., 2022, 123, 6039–6106 CrossRef PubMed.
  18. N. Richards, J. H. Carter, E. Nowicka, L. A. Parker, S. Pattisson, Q. He, N. F. Dummer, S. Golunski and G. J. Hutchings, Appl. Catal., B, 2020, 264, 118501 CrossRef.
  19. S. Hinokuma, T. Iwasa, Y. Kon, T. Taketsugu and K. Sato, Catal. Commun., 2021, 149, 106208 CrossRef CAS.
  20. S. Hinokuma, T. Iwasa, Y. Kon, T. Taketsugu and K. Sato, Sci. Rep., 2020, 10, 21605 CrossRef CAS PubMed.
  21. M. Miao, M. Zhang, H. Kong, T. Zhou, X. Yang and H. Yang, Energies, 2021, 14, 6148 CrossRef CAS.
  22. Y. Yin, R. Xiao, L. Qiu, C. Wang and H. Chang, Fuel, 2024, 366, 131303 CrossRef CAS.
  23. Y. Li, S. Yang, H. Peng, W. Liu, Y. Mi, Z. Wang, C. Tang, D. Wu and T. An, J. Catal., 2021, 395, 195–209 CrossRef CAS.
  24. R. Li, Y. Li and Z. Liu, Fuel, 2024, 355, 129405 CrossRef CAS.
  25. N. Richards, J. H. Carter, L. A. Parker, S. Pattisson, D. G. Hewes, D. J. Morgan, T. E. Davies, N. F. Dummer, S. Golunski and G. J. Hutchings, ACS Catal., 2020, 10, 5430–5442 CrossRef CAS.
  26. A. Szymaszek-Wawryca, P. Summa, D. Duraczyńska, U. Díaz and M. Motak, Materials, 2022, 15, 7884 CrossRef CAS PubMed.
  27. F. Lin, T. Andana, Y. Wu, J. Szanyi, Y. Wang and F. Gao, J. Catal., 2021, 401, 70–80 CrossRef CAS.
  28. F. Saleem, M. Z. Abid, K. Rafiq, A. Rauf, K. Ahmad, S. Iqbal, R. Jin and E. Hussain, Int. J. Hydrogen Energy, 2024, 52, 305–319 CrossRef CAS.
  29. N. Wajid, K. Rafiq, M. Z. Abid, A. Ilyas, T. Najam, A. Rauf and E. Hussain, Mater. Chem. Phys., 2023, 306, 128062 CrossRef CAS.
  30. M. Sabir, K. Rafiq, M. Z. Abid, U. Quyyum, S. S. A. Shah, M. Faizan, A. Rauf, S. Iqbal and E. Hussain, Fuel, 2023, 353, 129196 CrossRef CAS.
  31. T. Gaidei, A. Kokorin, N. Pillet, M. Srukova, E. Khaustova, G. Shmurak and N. Yaroshenko, Russ. J. Phys. Chem. A, 2007, 81, 895–900 CrossRef CAS.
  32. L. Han, S. Cai, M. Gao, J.-Y. Hasegawa, P. Wang, J. Zhang, L. Shi and D. Zhang, Chem. Rev., 2019, 119, 10916–10976 CrossRef CAS PubMed.
  33. D. Shun, D. H. Bae, J. Y. Paek and Y. S. Park, Korean J. Chem. Eng., 2004, 21, 890–894 CrossRef CAS.
  34. A. Ilyas, K. Rafiq, M. Z. Abid, A. Rauf and E. Hussain, RSC Adv., 2023, 13, 2379–2391 RSC.
  35. M. Z. Abid, A. Ilyas, K. Rafiq, A. Rauf, M. A. Nadeem, A. Waseem and E. Hussain, Environ. Sci.: Water Res. Technol., 2023, 9, 2238–2252 RSC.
  36. M. Z. Abid, K. Rafiq, A. Rauf, S. S. A. Shah, R. Jin and E. Hussain, Nanoscale Adv., 2023, 5, 3247–3259 RSC.
  37. J. Liu, L. Wang, W. Song, M. Zhao, J. Liu, H. Wang, Z. Zhao, C. Xu and Z. Duan, ACS Sustainable Chem. Eng., 2018, 7, 2811–2820 CrossRef.
  38. L.-L. Zhang, X.-M. Chen and C.-G. Liu, Inorg. Chem., 2019, 58, 5221–5229 CrossRef CAS PubMed.
  39. V. G. Komvokis, M. Marti, A. Delimitis, I. A. Vasalos and K. S. Triantafyllidis, Appl. Catal., B, 2011, 103, 62–71 CrossRef CAS.
  40. P. S. S. Reddya, N. S. Babua, N. Lingaiaha, P. S. Prasada and I. Raob, Proc. Eur. Cong. Chem. Eng., 2007, 1–12 Search PubMed.
  41. S. N. Basahel, M. Mokhtar, T. T. Ali and K. Narasimharao, Catal. Today, 2020, 348, 166–176 CrossRef CAS.
  42. N. Zhang, C. He, Y. Jing, Y. Qian, T. Toyao and K.-I. Shimizu, Surf. Interfaces, 2024, 46, 104120 CrossRef CAS.
  43. M. K. Singh and G. Rajaraman, Inorg. Chem., 2019, 58, 3175–3188 CrossRef CAS PubMed.
  44. M. G. Galloni, S. Campisi, A. Gervasini, S. Morandi and M. Manzoli, Appl. Catal., A, 2023, 655, 119101 CrossRef CAS.
  45. X. Hu, L. Wu, S. Ju, C. Dong, Y. Yang and W. Qin, Environ. Eng. Sci., 2014, 31, 308–316 CrossRef CAS.
  46. X. Li, J. Chen, C. Lu, G. Luo and H. Yao, Fuel, 2021, 299, 120910 CrossRef CAS.
  47. M. D. Esrafili, H. Janebi and P. Mousavian, Appl. Surf. Sci., 2021, 569, 151001 CrossRef CAS.
  48. H. Yu, X. Qi, X. Du, Y. Pan, X. Feng, W. Shan and Y. Xiong, Mol. Catal., 2023, 537, 112960 CAS.
  49. E. M. Goliaei and N. Seriani, Appl. Surf. Sci., 2022, 579, 152215 CrossRef.
  50. K. Feng, R. Lin, T. Chen, Y. Hu, Y. Feng, D. Chen, M. H. Tahir and T. Cao, J. Environ. Chem. Eng., 2024, 111891 CrossRef CAS.
  51. H. Liu, S. Yang, G. Wang, H. Liu, Y. Peng, C. Sun, J. Li and J. Chen, Environ. Sci. Technol., 2022, 56, 16325–16335 CrossRef CAS PubMed.
  52. H. Liu, J. Chen, Y. Wang, S. Xiong, Z. Su, Y. Wang, W. Yang, X. Chu, W. Yang and Y. Peng, Chem. Eng. J., 2021, 414, 128643 CrossRef CAS.
  53. X. Hu, Y. Wang, R. Wu and Y. Zhao, Mol. Catal., 2021, 509, 111656 CAS.
  54. W. Gao, D. Wen, J. Ho and Y. Qu, Mater. Today Chem., 2019, 12, 266–281 CrossRef CAS.
  55. F. Zasada, J. Gryboś, E. Budiyanto, J. Janas and Z. Sojka, J. Catal., 2019, 371, 224–235 CrossRef CAS.
  56. Q. Zhao, Z. Yan, C. Chen and J. Chen, Chem. Rev., 2017, 117, 10121–10211 CrossRef CAS PubMed.
  57. G. Chen, C. Guyon, Z. Zhang, B. Da Silva, P. Da Costa, S. Ognier, D. Bonn and M. Tatoulian, Microfluid. Nanofluid., 2014, 16, 141–148 CrossRef CAS.
  58. Q. Shen, L. Li, J. Li, H. Tian and Z. Hao, J. Hazard. Mater., 2009, 163, 1332–1337 CrossRef CAS PubMed.
  59. M.-J. Kim, S.-J. Lee, I.-S. Ryu, M.-W. Jeon, S.-H. Moon, H.-S. Roh and S. G. Jeon, Mol. Catal., 2017, 442, 202–207 CAS.
  60. Y. Li, A. Sundermann, O. Gerlach, K.-B. Low, C. C. Zhang, X. Zheng, H. Zhu and S. Axnanda, Catal. Today, 2019, 355, 608–619 CrossRef.
  61. J. Pérez-Ramírez, F. Kapteijn, G. Mul and J. A. Moulijn, J. Catal., 2002, 208, 211–223 CrossRef.
  62. X. Wei, Y. Wang, X. Li, R. Wu and Y. Zhao, Mol. Catal., 2020, 491, 111005 CAS.
  63. L. Rocha-Meneses, A. Inayat, M. Ayoub, S. Ullah, S. R. Naqvi, S. Farrukh, A. Mustafa, A. Z. Abdullah and A. H. Bhat, Environ. Prog. Sustainable Energy, 2023, 42, e13982 CrossRef CAS.
  64. S. Wojcik, P. Indyka, Z. Sojka and A. Kotarba, Catal. Today, 2020, 348, 111–117 CrossRef CAS.
  65. G. Grzybek, P. Stelmachowski, S. Gudyka, J. Duch, K. Ćmil, A. Kotarba and Z. Sojka, Appl. Catal., B, 2015, 168–169, 509–514 CrossRef CAS.
  66. M. Inger, J. Rajewski, M. Ruszak and M. Wilk, Chem. Papers, 2019, 73, 1979–1986 CrossRef CAS.
  67. B. M. Abu-Zied, S. A. Soliman and S. E. Abdellah, J. Ind. Eng. Chem., 2015, 21, 814–821 CrossRef CAS.
  68. S. Li, J. Zhao, Z. Song, H. Wang, T. Zhang, J. Liu and Q. Jiang, Fuel, 2024, 362, 130745 CrossRef CAS.
  69. A. Inayat, L. Rocha-Meneses, M. Ayoub, S. Ullah, A. Z. Abdullah, S. R. Naqvi and A. H. Bhat, Environ. Sci. Pollut. Res., 2023, 1–12 Search PubMed.
  70. T. Franken and R. Palkovits, Appl. Catal., B, 2015, 176–177, 298–305 CrossRef CAS.
  71. A. Klyushina, K. Pacultová, K. Karásková, K. Jirátová, M. Ritz, D. Fridrichová, A. Volodarskaja and L. Obalová, J. Mol. Catal. A: Chem., 2016, 425, 237–247 CrossRef CAS.
  72. H. Yu, X. Wang, X. Wu and Y. Chen, Chem. Eng. J., 2018, 334, 800–806 CrossRef CAS.
  73. M. Tursun, X. Wang, F. Zhang and H. Yu, Catal. Commun., 2015, 65, 1–5 CrossRef CAS.
  74. S. Wójcik, G. Grzybek, P. Stelmachowski, Z. Sojka and A. Kotarba, Catalysts, 2020, 10, 41 CrossRef.
  75. M. Inger, B. Moszowski, M. Ruszak, J. Rajewski and M. Wilk, Catalysts, 2020, 10, 987 CrossRef CAS.
  76. P. Hermawan, J. Kim, M. H. Jeong, S. A. Choi, S.-M. Hwang and S. K. Jeong, Chemosphere, 2024, 367, 143642 CrossRef PubMed.
  77. A. Klegova, K. Pacultová, T. Kiška, P. Peikertová, A. Rokicińska, P. Kuśtrowski and L. Obalová, Mol. Catal., 2022, 533, 112754 CAS.
  78. B. Abu-Zied, S. Soliman and S. Abdellah, Curr. Catal., 2020, 9, 152–162 CrossRef CAS.
  79. B. Li, X. Duan, T. Zhao, B. Niu, G. Li, Z. Zhao, Z. Yang, D. Liu, F. Zhang and J. Cheng, Environ. Sci. Technol., 2024, 58(4), 2153–2161 CrossRef CAS PubMed.
  80. P. Lu, L. Ye, X. Yan, X. Chen, P. Fang, D. Chen, D. Chen and C. Cen, J. Hazard. Mater., 2021, 414, 125468 CrossRef CAS PubMed.
  81. S. Li, Y. Li, W. Shen, Y. Bai and L. Kong, J. Cleaner Prod., 2022, 380, 134961 CrossRef CAS.
  82. Y. Matsumura and J. B. Moffat, J. Catal., 1994, 148, 323–333 CrossRef CAS.
  83. S. Sugiyama, K. Abe, H. Hayashi and J. B. Moffat, Appl. Catal., A, 1999, 183, 135–142 CrossRef CAS.
  84. Y. Cui, H. Liu, Y. Lin and Z. Ma, J. Taiwan Inst. Chem. Eng., 2016, 67, 254–262 CrossRef CAS.
  85. C. Huang, Z. Ma, P. Xie, Y. Yue, W. Hua and Z. Gao, J. Mol. Catal. A: Chem., 2015, 400, 90–94 CrossRef CAS.
  86. M. G. Galloni, S. Campisi, A. Gervasini, S. Morandi and M. Manzoli, Appl. Catal., A, 2023, 655, 119101 CrossRef CAS.
  87. X. Tan, H. Chen, L. Shi, Q. Lu, S. Qi, C. Yi and B. Yang, Catal. Lett., 2023, 153, 3724–3733 CrossRef CAS.
  88. Y. Lin, T. Meng and Z. Ma, J. Ind. Eng. Chem., 2015, 28, 138–146 CrossRef CAS.
  89. Y. Wang, X. Zhou, X. Wei, X. Li, R. Wu, X. Hu and Y. Zhao, Mol. Catal., 2021, 501, 111370 CAS.
  90. M. G. Galloni, Development of metal functionalized hydroxyapatite catalysts for air-quality protection, PhD Thesis, Università degli Studi di Milano, 2022 Search PubMed.
  91. M. Galloni, S. Campisi and A. Gervasini, Iron functionalised hydroxyapatites as efficient eco-friendly catalysts for air-quality protection, 2020 Search PubMed.
  92. X. Liu, Y. Wang, R. Wu and Y. Zhao, Catal. Surveys Asia, 2021, 25, 168–179 CrossRef CAS.
  93. K.-M. Lee, G. Kwon, S. Hwang, J. A. Boscoboinik and T. Kim, Catal. Sci. Technol., 2021, 11, 7850–7865 RSC.
  94. D. Jia, K. Hanna, G. Mailhot and M. Brigante, Molecules, 2021, 26, 5748 CrossRef CAS PubMed.
  95. U. Quyyum, K. Rafiq, M. Z. Abid, F. Ahmad, A. Rauf and E. Hussain, Environ. Sci.: Water Res. Technol., 2023, 9, 1147–1160 RSC.
  96. M. Jalil, K. Rafiq, M. Z. Abid, A. Rauf, S. Wang, S. Iqbal and E. Hussain, Nanoscale Adv., 2023, 5, 3233–3246 RSC.
  97. K. U. Sahar, K. Rafiq, M. Z. Abid, A. Rauf, U. U. Rehman, M. A. Nadeem, R. Jin and E. Hussain, Colloids Surf., A, 2023, 674, 131942 CrossRef CAS.
  98. A. Aslam, M. Z. Abid, K. Rafiq, A. Rauf and E. Hussain, Sci. Rep., 2023, 13, 6306 CrossRef CAS PubMed.
  99. H. Beyer, J. Emmerich, K. Chatziapostolou and K. Koehler, Appl. Catal., A, 2011, 391, 411–416 CrossRef CAS.
  100. L. Han, M. Gao, J.-Y. Hasegawa, S. Li, Y. Shen, H. Li, L. Shi and D. Zhang, Environ. Sci. Technol., 2019, 53, 6462–6473 CrossRef CAS PubMed.
  101. C. Ratanatawanate, A. Chyao and K. J. Balkus Jr, J. Am. Chem. Soc., 2011, 133, 3492–3497 CrossRef CAS PubMed.
  102. T. Gaidei, A. Kokorin, N. Pillet, V. Sadov, M. Strukova, S. Filatov, E. Khaustova and N. Yaroshenko, Russ. J. Appl. Chem., 2010, 83, 1130–1138 CrossRef CAS.
  103. S. Imamura, J.-i Tadani, Y. Saito, Y. Okamoto, H. Jindai and C. Kaito, Appl. Catal., A, 2000, 201, 121–127 CrossRef CAS.
  104. Y. Li, A. Sundermann, O. Gerlach, K.-B. Low, C. C. Zhang, X. Zheng, H. Zhu and S. Axnanda, Catal. Today, 2020, 355, 608–619 CrossRef CAS.
  105. T. C. Peck, G. K. Reddy, M. Jones and C. A. Roberts, J. Phys. Chem. C, 2017, 121, 8435–8443 CrossRef CAS.
  106. M. Konsolakis, ACS Catal., 2015, 5, 6397–6421 CrossRef CAS.
  107. Y. Ni, Z. Han, Y. Chai, G. Wu and L. Li, EES Catal., 2023, 1, 459–494 RSC.
  108. H. Xu, Y. Ma, J. Chen, W.-X. Zhang and J. Yang, Chem. Soc. Rev., 2022, 51, 2710–2758 RSC.
  109. M. Zabilskiy, P. Djinović, E. Tchernychova and A. Pintar, Appl. Catal., B, 2016, 197, 146–158 CrossRef CAS.
  110. H. Zhu, Y. Li and X. Zheng, Appl. Catal., A, 2019, 571, 89–95 CrossRef CAS.
  111. X. Fan, S. Kang, J. Li and T. Zhu, RSC Adv., 2018, 8, 26998–27007 RSC.
  112. H. Zeng and X. Pang, Appl. Catal., B, 1997, 13, 113–122 CrossRef CAS.
  113. M. Jabłońska, M. A. Arán, A. M. Beale, K. Góra-Marek, G. Delahay, C. Petitto, K. Pacultová and R. Palkovits, RSC Adv., 2019, 9, 3979–3986 RSC.
  114. S. Carabineiro, E. Papista, G. Marnellos, P. Tavares, F. Maldonado-Hódar and M. Konsolakis, Mol. Catal., 2017, 436, 78–89 CAS.
  115. M. Lykaki, E. Papista, N. Kaklidis, S. A. Carabineiro and M. Konsolakis, Catalysts, 2019, 9, 233 CrossRef.
  116. K. Karásková, K. Pacultová, K. Jirátová, D. Fridrichová, M. Koštejn and L. Obalová, Catalysts, 2020, 10, 1134 CrossRef.
  117. X. Hu, Y. Wang, R. Wu, L. Zhao, X. Wei and Y. Zhao, Appl. Surf. Sci., 2020, 514, 145892 CrossRef CAS.
  118. T. Zhao, Y. Li, Q. Gao, Z. Liu and X. Xu, Catal. Commun., 2020, 137, 105948 CrossRef CAS.
  119. P. H. Ho, M. Jabłońska, M. Nocuń, G. Fornasari, F. Ospitali, A. Vaccari, P. Benito and R. Palkovits, ChemCatChem, 2019, 11, 5580–5592 CrossRef CAS.
  120. E. M. Iwanek, L. F. Liotta, G. Pantaleo, K. Krawczyk, E. Gdyra, J. Petryk, J. W. Sobczak and Z. Kaszkur, Catalysts, 2021, 11, 325 CrossRef CAS.
  121. B. M. Abu-Zied, L. Obalová, K. Pacultová, A. Klegova and A. M. Asiri, J. Ind. Eng. Chem., 2021, 93, 279–289 CrossRef CAS.
  122. E. M. Kostyukhin, A. L. Kustov, N. V. Evdokimenko, A. I. Bazlov and L. M. Kustov, J. Am. Ceram. Soc., 2021, 104, 492–503 CrossRef CAS.
  123. D. Wang, Q. Yao, S. Hui and Y. Niu, Fuel, 2019, 251, 23–29 CrossRef CAS.
  124. W. Muhammad, L. Wu, A. El Kasmi, A. Muhammad and Z. Tian, J. Therm. Sci., 2023, 32, 531–541 CrossRef CAS.
  125. Y. Zeng, Z. Wu, L. Guo, Y. Wang, S. Zhang and Q. Zhong, Mol. Catal., 2020, 488, 110916 CAS.
  126. Y. Zeng, F. Lyu, Y. Wang, S. Zhang, Q. Zhong and Z. Zhong, Mol. Catal., 2022, 525, 112356 CAS.
  127. L. Sheng, Z. Ma, S. Chen, J. Lou, C. Li, S. Li, Z. Zhang, Y. Wang and H. Yang, Chin. J. Catal., 2019, 40, 1070–1077 CrossRef CAS.
  128. Z. Z. Khan, I. A. Khan, I. Khan, M. H. S. Wattoo and A. Badshah, Solid State Sci., 2019, 98, 106035 CrossRef CAS.
  129. D. Wang, Q. Yao, C. Mou, S. Hui and Y. Niu, Fuel, 2019, 254, 115719 CrossRef CAS.
  130. F. Zhao, D. Wang, X. Li, Y. Yin, C. Wang, L. Qiu, J. Yu and H. Chang, Ind. Eng. Chem. Res., 2022, 61, 13854–13862 CrossRef CAS.
  131. M.-x Xu, H.-X. Wang, H.-d Ouyang, L. Zhao and Q. Lu, J. Hazard. Mater., 2021, 401, 123334 CrossRef CAS PubMed.
  132. M. Zabilskiy, I. Arčon, P. Djinović, E. Tchernychova and A. Pintar, ChemCatChem, 2021, 13, 1814–1823 CrossRef CAS.
  133. A. A. Khan, A. Ahmad, H. M. Al-Swaidan, S. Haider, M. S. Akhtar and S. U. Khan, Mol. Catal., 2022, 527, 112392 CAS.
  134. C.-G. Liu, Y.-J. Chu, L.-L. Zhang, C. Sun and J.-Y. Shi, Environ. Sci. Technol., 2019, 53, 12893–12903 CrossRef CAS PubMed.
  135. M. Wang, W. Li and S. Wang, J. Phys. Chem. C, 2023, 127, 20344–20353 CrossRef CAS.
  136. S. Xie, D. Kim, K. Ye, L. Tetard and F. Liu, J. Rare Earths, 2023, 41, 941–951 CrossRef CAS.
  137. M. Jabłońska and R. Palkovits, Catal. Sci. Technol., 2016, 6, 49–72 RSC.
  138. D. F. Swearer, H. Robatjazi, J. M. P. Martirez, M. Zhang, L. Zhou, E. A. Carter, P. Nordlander and N. J. Halas, ACS Nano, 2019, 13, 8076–8086 CrossRef CAS PubMed.
  139. D. Tichit and B. Coq, CATTECH, 2003, 7, 206–217 CrossRef CAS.
  140. Y. Zhang, X. Wang, Y. Zhu, B. Hou, X. Yang, X. Liu, J. Wang, J. Li and T. Zhang, J. Phys. Chem. C, 2014, 118, 1999–2010 CrossRef CAS.
  141. X. Xing, N. Li, J. Cheng, Y. Sun, G. Wang, Z. Zhang, H. Xu, C. He and Z. Hao, Ind. Eng. Chem. Res., 2019, 58, 9362–9371 CrossRef CAS.
  142. F. Gholami, M. Tomas, Z. Gholami and M. Vakili, Sci. Total Environ., 2020, 714, 136712 CrossRef CAS PubMed.
  143. A. De Stefanis, M. Dondi, G. Perez and A. Tomlinson, Chemosphere, 2000, 41, 1161–1165 CrossRef CAS PubMed.
  144. E. E. Kiss, T. J. Vulić and A. F. Reitzmann, React. Kinet. Catal. Lett., 2005, 86, 233–240 CrossRef CAS.
  145. M. C. Karaeyvaz and S. Balci, Microporous Mesoporous Mater., 2021, 323, 111193 CrossRef.
  146. S. Kannan and C. Swamy, Catal. Today, 1999, 53, 725–737 CrossRef CAS.
  147. M. Jabłońska, M. A. Arán, A. M. Beale, G. Delahay, C. Petitto, M. Nocuń and R. Palkovits, Appl. Catal., B, 2019, 243, 66–75 CrossRef.
  148. Y. Zhang, Y. Guo, N. Li and Y. Feng, Pol. J. Environ. Stud., 2019, 28, 4477–4484 CrossRef CAS PubMed.
  149. P. H. Ho, M. Jabłońska, R. Palkovits, E. Rodríguez-Castellón, F. Ospitali, G. Fornasari, A. Vaccari and P. Benito, Chem. Eng. J., 2020, 379, 122259 CrossRef CAS.
  150. S. Xiong, J. Chen, N. Huang, S. Yang, Y. Peng and J. Li, Environ. Sci. Technol., 2019, 53, 10379–10386 CrossRef CAS PubMed.
  151. M. Armandi, T. Andana, S. Bensaid, M. Piumetti, B. Bonelli and R. Pirone, Catal. Today, 2020, 345, 59–70 CrossRef CAS.
  152. A. Inayat, M. Ayoub, A. Z. Abdullah, S. Ullah and S. R. Naqvi, Environ. Prog. Sustainable Energy, 2019, 38, 13129 CrossRef.
  153. D. Lopes, F. Zotin and L. A. Palacio, Appl. Catal., B, 2018, 237, 327–338 CrossRef CAS.
  154. J. Akil, C. Ciotonea, S. Siffert, S. Royer, L. Pirault-Roy, R. Cousin and C. Poupin, Catal. Today, 2022, 384, 97–105 CrossRef.
  155. C. d O. P. Teixeira, S. da Silva Montani, J. C. S. Soares, L. A. Palacio and F. M. Z. Zotin, Mol. Catal., 2024, 561, 114157 CAS.
  156. Z. Zhuang, B. Guan, J. Chen, C. Zheng, J. Zhou, T. Su, Y. Chen, C. Zhu, X. Hu and S. Zhao, Chem. Eng. J., 2024, 150374 CrossRef CAS.
  157. Z. Tišler, A. Klegová, E. Svobodová, J. Šafář, K. Strejcová, J. Kohout, S. Šlang, K. Pacultová, D. Rodríguez-Padrón and R. Bulánek, Catalysts, 2020, 10, 1398 CrossRef.
  158. G. He, B. Zhang, H. He, X. Chen and Y. Shan, Sci. Total Environ., 2019, 673, 266–271 CrossRef CAS PubMed.
  159. L. J. Lobree, I.-C. Hwang, J. A. Reimer and A. T. Bell, J. Catal., 1999, 186, 242–253 CrossRef CAS.
  160. A. Ates, A. Reitzmann, C. Hardacre and H. Yalcin, Appl. Catal., A, 2011, 407, 67–75 CrossRef CAS.
  161. M. Rutkowska, A. Jankowska, E. Różycka-Dudek, W. Dubiel, A. Kowalczyk, Z. Piwowarska, S. Llopis, U. Díaz and L. Chmielarz, Catalysts, 2020, 10, 1139 CrossRef CAS.
  162. J. Zhang, X. Tang, H. Yi, Q. Yu, Y. Zhang, J. Wei and Y. Yuan, Appl. Catal., A, 2022, 630, 118467 CrossRef CAS.
  163. Y. Li and J. N. Armor, Appl. Catal., B, 1992, 1, L21–L29 CrossRef CAS.
  164. S. A. Yashnik, O. P. Taran, T. A. Surovtsova, A. B. Ayusheev and V. N. Parmon, J. Environ. Chem. Eng., 2022, 10, 107950 CrossRef CAS.
  165. N. F. Biturini, A. P. N. Santos and M. S. Batista, React. Kinet., Mech. Catal., 2019, 126, 341–352 CrossRef CAS.
  166. P. J. Smeets, J. S. Woertink, B. F. Sels, E. I. Solomon and R. A. Schoonheydt, Inorg. Chem., 2010, 49, 3573–3583 CrossRef CAS PubMed.
  167. T. Žumbar, I. Arčon, P. Djinović, G. Aquilanti, G. Žerjav, A. Pintar, A. Ristić, G. Dražić, J. Volavšek and G. Mali, ACS Appl. Mater. Interfaces, 2023, 15, 28747–28762 CrossRef PubMed.
  168. B. Liu, D. Yao, F. Wu, L. Wei, X. Li and X. Wang, Ind. Eng. Chem. Res., 2019, 58, 20516–20527 CrossRef CAS.
  169. S. Yasumura, H. Ide, T. Ueda, Y. Jing, C. Liu, K. Kon, T. Toyao, Z. Maeno and K.-I. Shimizu, JACS Au, 2021, 1, 201–211 CrossRef CAS PubMed.
  170. X. Zhang, Q. Shen, C. He, C. Ma, J. Cheng, L. Li and Z. Hao, ACS Catal., 2012, 2, 512–520 CrossRef CAS.
  171. M. Jin, P. Wang, L. Zhang, C. Ao and L. Lei, Therm. Sci. Eng. Prog., 2023, 45, 102088 CrossRef CAS.
  172. M. Jabłońska, M. E. Potter and A. M. Beale, ChemCatChem, 2024, 16, e202301214 CrossRef.
  173. C. Ruan, X. Wang, C. Wang, L. Zheng, L. Li, J. Lin, X. Liu, F. Li and X. Wang, Nat. Commun., 2022, 13, 718 CrossRef CAS PubMed.
  174. W.-S. Ju, M. Matsuoka and M. Anpo, Catal. Lett., 2001, 71, 91–93 CrossRef CAS.
  175. S. Dzwigaj, D. Reja, S. Koné-Guira, A. Miche, G. Costentin and C. Thomas, Appl. Catal., A, 2023, 657, 119119 CrossRef CAS.
  176. M. Shukla, B. V. Chauhan, S. Verma and A. Dhar, Solids, 2022, 3, 665–683 CrossRef CAS.
  177. B. Bromley, C. Pischetola, L. Nikoshvili, F. Cárdenas-Lizana and L. Kiwi-Minsker, Molecules, 2020, 25, 3867 CrossRef CAS PubMed.
  178. B. Kang, M. Li, Z. Di, X. Guo, Y. Wei, J. Jia and R. Zhang, Catal. Today, 2022, 402, 17–26 CrossRef CAS.
  179. B. Kang, R. Zhang, M. Guo, X. Guo, Z. Di, Y. Wei and J. Jia, Energy Fuels, 2023, 37(23), 18019–18029 CrossRef CAS.
  180. D. Pietrogiacomi, M. C. Campa, L. Ardemani and M. Occhiuzzi, Catal. Today, 2019, 336, 131–138 CrossRef CAS.
  181. G. A. Zenkovets, R. A. Shutilov, V. I. Sobolev and V. Y. Gavrilov, Catal. Commun., 2020, 144, 106072 CrossRef CAS.
  182. C. Gao, J. Li, J. Zhang and X. Sun, Catalysts, 2020, 10, 646 CrossRef CAS.
  183. J. B. Lim, S. H. Cha and S. B. Hong, Appl. Catal., B, 2019, 243, 750–759 CrossRef CAS.
  184. K. M. Farhan, A. N. Thabassum, T. M. Ismail and P. Sajith, Catal. Sci. Technol., 2022, 12, 1466–1475 RSC.
  185. B. Zhang, G. He, Y. Shan and H. He, Catal. Today, 2019, 327, 177–181 CrossRef CAS.
  186. Y. You, S. Chen, J. Li, J. Zeng, H. Chang, L. Ma and J. Li, J. Hazard. Mater., 2020, 383, 121117 CrossRef CAS PubMed.
  187. L. M. Kustov, S. F. Dunaev and A. L. Kustov, Molecules, 2022, 27, 398 CrossRef CAS PubMed.
  188. S. Li, J. Wang, R. Shang, J. Zhao, Q. Xu, H. Wang and J. Liu, Mol. Catal., 2024, 552, 113706 CAS.
  189. X. Liang, H. Tang, F. Yang, G. Tu, F. Zhang, Q. Xiao, Y. Zhong and W. Zhu, Microporous Mesoporous Mater., 2019, 290, 109655 CrossRef CAS.
  190. G. Zhao, E. Benhelal, A. Adesina, E. Kennedy and M. Stockenhuber, J. Phys. Chem. C, 2019, 123, 27436–27447 CrossRef CAS.
  191. J. Zeng, S. Chen, Z. Fan, C. Wang, H. Chang and J. Li, Ind. Eng. Chem. Res., 2020, 59, 19500–19509 CrossRef CAS.
  192. J. Cheng, D. Zheng, G. Yu, R. Xu, C. Dai, N. Liu, N. Wang and B. Chen, ACS Catal., 2022, 13, 934–947 CrossRef.
  193. T. Zhang, Y. Qiu, G. Liu, J. Chen, Y. Peng, B. Liu and J. Li, J. Catal., 2020, 392, 322–335 CrossRef CAS.
  194. J. Han, A. Wang, G. Isapour, H. Harelind, M. Skoglundh, D. Creaser and L. Olsson, Ind. Eng. Chem. Res., 2021, 60, 17826–17839 CrossRef CAS.
  195. S. Li, C. Zhang, J. Li, Y. Li, H. Wang, C. Li and Y. Song, Res. Chem. Intermed., 2019, 45, 3601–3616 CrossRef CAS.
  196. G. Sádovská, M. Bernauer, B. Bernauer, E. Tabor, A. Vondrová and Z. Sobalík, Catal. Commun., 2018, 112, 58–62 CrossRef.
  197. M. L. Bols, B. E. Snyder, H. M. Rhoda, P. Cnudde, G. Fayad, R. A. Schoonheydt, V. Van Speybroeck, E. I. Solomon and B. F. Sels, Nat. Catal., 2021, 4, 332–340 CrossRef CAS.
  198. T. Wu, Y. Shen, L. Feng, Z. Tang and D. Zhang, J. Chem. Eng. Data, 2019, 64, 3473–3482 CrossRef CAS.
  199. A. J. Shih, J. M. González, I. Khurana, L. P. Ramírez, A. Peña, L. A. Kumar and A. L. Villa, ACS Catal., 2021, 11, 10362–10376 CrossRef CAS.
  200. S. J. Lee, I. S. Ryu, S. G. Jeon and S. H. Moon, Environ. Prog. Sustainable Energy, 2019, 38, 451–456 CrossRef CAS.
  201. M. C. Campa, D. Pietrogiacomi, C. Catracchia, S. Morpurgo, J. Olszowka, K. Mlekodaj, M. Lemishka, J. Dedecek, A. Kornas and E. Tabor, Appl. Catal., B, 2024, 342, 123360 CrossRef CAS.
  202. E. Tabor, K. Mlekodaj, G. Sádovská, M. Bernauer, P. Klein, P. Sazama, J. Dědeček and Z. Sobalík, Microporous Mesoporous Mater., 2019, 281, 15–22 CrossRef CAS.
  203. E. Tabor, G. Sádovská, M. Bernauer, P. Sazama, J. Nováková, V. Fila, T. Kmječ, J. Kohout, K. Závěta and Z. Sobalík, Appl. Catal., B, 2019, 240, 358–366 CrossRef CAS.
  204. N. Pal and A. Bhaumik, RSC Adv., 2015, 5, 24363–24391 RSC.
  205. P. Verma, Y. Kuwahara, K. Mori, R. Raja and H. Yamashita, Nanoscale, 2020, 12, 11333–11363 RSC.
  206. Y. Qin, Z. Qu, C. Dong, Y. Wang and N. Huang, J. Environ. Sci., 2019, 76, 208–216 CrossRef CAS PubMed.
  207. M. Marafi and E. Furimsky, Energy Fuels, 2017, 31, 5711–5750 CrossRef CAS.
  208. X. Wei, X.-F. Yang, A.-Q. Wang, L. Li, X.-Y. Liu, T. Zhang, C.-Y. Mou and J. Li, J. Phys. Chem. C, 2012, 116, 6222–6232 CrossRef CAS.
  209. L. Yu, H. Hu, H. B. Wu and X. W. Lou, Adv. Mater., 2017, 29, 1604563 CrossRef PubMed.
  210. K. Pacultová, A. Klegova, T. Kiška, D. Fridrichová, A. Martaus, A. Rokicińska, P. Kuśtrowski and L. Obalová, Mater. Res. Bull., 2020, 129, 110892 CrossRef.
  211. M. C. Campa, A. M. Doyle, G. Fierro and D. Pietrogiacomi, Catal. Today, 2022, 384, 76–87 CrossRef.
  212. B. Bozorgi, J. Karimi-Sabet and P. Khadiv-Parsi, Environ. Technol. Innovation, 2022, 26, 102344 CrossRef CAS.
  213. T. Umegaki, H. Katori, K. Otake, R. Yamamoto and Y. Kojima, J. Sol-Gel Sci. Technol., 2019, 92, 715–722 CrossRef CAS.
  214. S. Wang, B. Yan, J. Chai, T. Li, H. Yu, T. Li, P. Cao, F. Yang, X. Yuan and H. Yin, Eur. J. Inorg. Chem., 2021, 2201–2210 CrossRef.
  215. O. Muccioli, E. Meloni, S. Renda, M. Martino, F. Brandani, P. Pullumbi and V. Palma, Processes, 2023, 11, 1511 CrossRef CAS.
  216. Z. Li, B. Li, C. Yu, H. Wang and Q. Li, Adv. Sci., 2023, 10, 2206605 CrossRef CAS PubMed.
  217. T. Boningari, D. K. Pappas and P. G. Smirniotis, J. Catal., 2018, 365, 320–333 CrossRef CAS.
  218. D. K. Pappas, T. Boningari, P. Boolchand and P. G. Smirniotis, J. Catal., 2016, 334, 1–13 CrossRef CAS.
  219. P. Nematollahi and M. D. Esrafili, RSC Adv., 2016, 6, 59091–59099 RSC.
  220. J.-H. Lee, Y.-J. An, H. Choi, J. Lee, S.-Y. Lee and S.-J. Park, Appl. Catal., O, 2024, 206937 CAS.
  221. M. D. Esrafili and N. Saeidi, Appl. Surf. Sci., 2017, 403, 43–50 CrossRef CAS.
  222. M. D. Esrafili and N. Saeidi, Appl. Surf. Sci., 2018, 444, 584–589 CrossRef CAS.
  223. M. D. Esrafili, Phys. Lett. A, 2017, 381, 2085–2091 CrossRef CAS.
  224. B. Gao, J.-X. Zhao, Q.-H. Cai, X.-G. Wang and X.-Z. Wang, J. Phys. Chem. A, 2011, 115, 9969–9976 CrossRef CAS PubMed.
  225. S. K. Tiwari, S. Sahoo, N. Wang and A. Huczko, J. Sci.: Adv. Mater. Devices, 2020, 5, 10–29 Search PubMed.
  226. J. Li, M. Zhang, Q. Li and J. Yang, Appl. Surf. Sci., 2017, 391, 184–193 CrossRef CAS.
  227. P. Liu, X. Gu, K. Kang, H. Zhang, J. Cheng and H. Su, ACS Appl. Mater. Interfaces, 2017, 9, 10759–10767 CrossRef CAS PubMed.
  228. T. Boningari, D. K. Pappas and P. G. Smirniotis, J. Catal., 2018, 365, 320–333 CrossRef CAS.
  229. M. D. Esrafili and N. Saeidi, Appl. Surf. Sci., 2017, 403, 43–50 CrossRef CAS.
  230. M. D. Esrafili and S. Heydari, Mol. Phys., 2020, 1–8,  DOI:10.1080/00268976.2020.1759830,.
  231. M. D. Esrafili, Phys. Lett. A, 2017, 381, 2085–2091 CrossRef CAS.
  232. M. D. Esrafili and S. Heidari, Chem. Phys. Lett., 2019, 725, 52–58 CrossRef CAS.
  233. J. Sun, L. Wang, L. Zhang, Y. Zhao, Y. Chi, H. Wang, C. Li, J. Liu and J. Liu, ACS Appl. Energy Mater., 2021, 4, 8496–8505 CrossRef CAS.
  234. X. Liu and L. Sheng, Mater. Today Commun., 2021, 28, 102585 CrossRef CAS.
  235. G. Fan, Q. Wang, H. Xu, X. Wang, X. Tu and X. Chu, Appl. Surf. Sci., 2021, 544, 148776 CrossRef CAS.
  236. Y. Xiong, Y. Zhao, W. Shan, X. Feng, J. Cui, Z. Lou, G. Shao, M. Dong and H. Yu, Chemosphere, 2022, 303, 135257 CrossRef CAS PubMed.
  237. Y. Xiong, Y. Zhao, X. Qi, J. Qi, Y. Cui, H. Yu and Y. Cao, Environ. Sci. Technol., 2021, 55, 13335–13344 CAS.
  238. E. Meloni, M. Martino, S. Renda, O. Muccioli, P. Pullumbi, F. Brandani and V. Palma, Catalysts, 2022, 12, 1405 CrossRef CAS.
  239. X. Guan, H. Asakura, R. Han, S. Xu, H.-X. Liu, L. Chen, Z. Yao, J. H. C. Yan, T. Tanaka and Y. Guo, ACS Catal., 2023, 13, 13816–13827 CrossRef CAS PubMed.
  240. A. Hassanpour, M. Kamel, S. Ebrahimiasl, A. G. Ebadi, S. Arshadi and Z. Ghulinezhad Ahangari, J. Mol. Model., 2022, 28, 1–14 CrossRef PubMed.
  241. X. Hu, E. Zhang, W. Li, L. Wu, Y. Zhou, H. Zhang and C. Dong, Appl. Sci., 2022, 12, 5034 CrossRef CAS.
  242. X. Tan, S. Qi, G. Cheng, C. Yi and B. Yang, SSRN, 2023, preprint DOI:10.2139/ssrn.4470798.
  243. M. Derdare, A.-G. Boudjahem and M. Boulbazine, Struct. Chem., 2022, 33, 2043–2062 CrossRef CAS.
  244. M. D. Esrafili, S. Asadollahi and S. Heydari, J. Mol. Graphics Modell., 2019, 89, 41–49 CrossRef CAS PubMed.
  245. A. A. Khan, M. D. Esrafili, F. Ali, R. Ahmad and I. Ahmad, J. Mol. Graphics Modell., 2022, 114, 108186 CrossRef PubMed.
  246. P. Li, T. Zhang, H. Sun, Y. Gao, Y. Zhang, Y. Liu, C. Ge, H. Chen, X. Dai and X. Zhang, Nano Res., 2022, 15, 3001–3009 CrossRef CAS.
  247. J. Qi, X. Qi, Y. Pan, J. Cui, Y. Xiong, W. Shan and H. Yu, Appl. Surf. Sci., 2023, 611, 155657 CrossRef CAS.
  248. Y. Bai, Y. Hou, Y. Guo, N. Xiang, X. Han, H. Wang, Z. Wu and Z. Huang, J. Colloid Interface Sci., 2022, 616, 55–66 CrossRef CAS PubMed.
  249. Z. Liu, H. Wang, Y. Gao and J. Zhao, Molecules, 2023, 28, 4485 CrossRef CAS PubMed.
  250. M. D. Esrafili and B. Nejadebrahimi, Chem. Phys. Lett., 2019, 716, 11–16 CrossRef CAS.
  251. X. Dai, Y. Cheng, T. Liu and L. Mao, Energy Fuels, 2024, 38(11), 9262–9279 CrossRef CAS.
  252. T. Das, B. K. Sharma, A. K. Katiyar and J.-H. Ahn, J. Semicond., 2018, 39, 011007 CrossRef.
  253. A. Akça, O. Karaman and C. Karaman, ECS J. Solid State Sci. Technol., 2021, 10, 041003 CrossRef.
  254. X. Liu and L. Sheng, Adv. Theory Simul., 2023, 6, 2200680 CrossRef CAS.
  255. R. Gholizadeh, Y.-X. Yu and Y. Wang, Appl. Surf. Sci., 2017, 420, 944–953 CrossRef CAS.
  256. M. D. Esrafili, Chem. Phys. Lett., 2018, 708, 94–99 CrossRef CAS.
  257. M. D. Esrafili, Chem. Phys. Lett., 2018, 705, 44–49 CrossRef CAS.
  258. M. Vakili, R. Gholizadeh, A. Ghadi, E. Salmasi and M. Sinnokrot, J. Mol. Graphics Modell., 2020, 101, 107752 CrossRef CAS PubMed.
  259. Z. Liu, X.-r Cheng, Y.-m Yang, H.-z Jia, B.-q Bai and L. Zhao, Materials, 2019, 12, 2611 CrossRef CAS PubMed.
  260. I. Majeed, U. Manzoor, F. K. Kanodarwala, M. A. Nadeem, E. Hussain, H. Ali, A. Badshah, J. A. Stride and M. A. Nadeem, Catal. Sci. Technol., 2018, 8, 1183–1193 RSC.
  261. L. Hu, J. Yan, C. Wang, B. Chai and J. Li, Chin. J. Catal., 2019, 40, 458–469 CrossRef CAS.
  262. K. U. Sahar, K. Rafiq, M. Z. Abid, U. ur Rehman, A. Rauf and E. Hussain, React. Chem. Eng., 2023, 8, 2522–2536 RSC.
  263. I. Troppová, M. Šihor, M. Reli, M. Ritz, P. Praus and K. Kočí, Appl. Surf. Sci., 2018, 430, 335–347 CrossRef.
  264. D. Ma, X. Cao and Z. Cao, J. Phys. Chem. C, 2023, 127, 5800–5809 CrossRef CAS.
  265. M. D. Esrafili, Chem. Phys. Lett., 2018, 708, 94–99 CrossRef CAS.
  266. M. D. Esrafili and E. Vessally, Surf. Sci., 2018, 667, 105–111 CrossRef CAS.
  267. R. Gholizadeh, Y.-X. Yu and Y. Wang, Appl. Surf. Sci., 2017, 420, 944–953 CrossRef CAS.
  268. M. D. Esrafili, F. Mohammadian-Sabet and P. Nematollahi, RSC Adv., 2016, 6, 64832–64840 RSC.
  269. P. Praus, L. Svoboda, M. Ritz, I. Troppová, M. Šihor and K. Kočí, Mater. Chem. Phys., 2017, 193, 438–446 CrossRef CAS.
  270. K. Kočí, M. Reli, I. Troppová, M. Šihor, J. Kupková, P. Kustrowski and P. Praus, Appl. Surf. Sci., 2017, 396, 1685–1695 CrossRef.
  271. M. Reli, L. Svoboda, M. Šihor, I. Troppová, J. Pavlovský, P. Praus and K. Kočí, Environ. Sci. Pollut. Res., 2018, 25, 34839–34850 CrossRef CAS PubMed.
  272. K. Kočí, M. Reli, I. Troppová, M. Šihor, T. Bajcarová, M. Ritz, J. Pavlovský and P. Praus, Catalysts, 2019, 9, 735 CrossRef.
  273. M. Reli, I. Troppová, M. Šihor, J. Pavlovský, P. Praus and K. Kočí, Appl. Surf. Sci., 2019, 469, 181–191 CrossRef CAS.
  274. P. Praus, J. Lang, A. Martaus, L. Svoboda, V. Matějka, M. Kormunda, M. Šihor, M. Reli and K. Kočí, J. Inorg. Organomet. Polym. Mater., 2019, 29, 1219–1234 CrossRef CAS.
  275. V. Matějka, M. Šihor, M. Reli, A. Martaus, K. Kočí, M. Kormunda and P. Praus, Mater. Sci. Semicond. Process., 2019, 100, 113–122 CrossRef.
  276. P. Maitarad, A. Junkaew, V. Promarak, L. Shi and S. Namuangruk, Appl. Surf. Sci., 2020, 508, 145255 CrossRef CAS.
  277. S. Su, J. Ma, Z. Liu, D. Holiharimanana and H. Sun, Catalysts, 2023, 13, 578 CrossRef CAS.
  278. M. D. Esrafili and S. Heydari, Mol. Phys., 2020, 118, e1759830 CrossRef.
  279. X. Hu, Y. Wang, R. Wu and Y. Zhao, Appl. Surf. Sci., 2021, 538, 148157 CrossRef CAS.
  280. D. Masih, Y. Ma and S. Rohani, Appl. Catal., B, 2017, 206, 556–588 CrossRef CAS.
  281. J. O. Ighalo, M. L. Smith, A. Al Mayyahi and P. B. Amama, Appl. Catal., B, 2024, 124352 CrossRef CAS.
  282. S. Zhao, D. Chen, N. Li, Q. Xu, H. Li and J. Lu, J. Alloys Compd., 2023, 960, 170830 CrossRef CAS.
  283. B. Huang, Y. F. Wu, Z. Zhang, R. Chen, G. Y. Ren, N. Zhou, N. Li and Y. Qian, J. Mater. Chem. A, 2025, 13, 16970–16980 RSC.
  284. Y.-X. Yu, J. Colloid Interface Sci., 2025, 137799 CrossRef CAS PubMed.
  285. B. Huang, G. y Ren, R. Chen and N. Li, Chem. Phys. Lett., 2025, 874–875, 142178 CrossRef CAS.
  286. X. Hu, Y. Wang, Z. Ling, H. Song, Y. Cai, Z. Li, D. Zu and C. Li, Appl. Surf. Sci., 2021, 556, 149817 CrossRef CAS.
  287. A. Al Mayyahi, S. Sarker, B. M. Everhart, B. Tonyali, U. Yucel and P. B. Amama, J. Phys. Chem. Solids, 2022, 170, 110875 CrossRef CAS.
  288. A. Hermawan, T. Hasegawa, Y. Asakura and S. Yin, Sep. Purif. Technol., 2021, 270, 118815 CrossRef CAS.
  289. A. A. Mayyahi, S. Sarker, B. M. Everhart, X. He and P. B. Amama, Mater. Today Commun., 2022, 32, 103835 CrossRef CAS.
  290. L. Kong, M. Wang and C.-M. L. Wu, ACS Mater. Lett., 2024, 6, 1711–1721 CrossRef CAS.
  291. H. D. Mai, K. Rafiq and H. Yoo, Chem. – Eur. J., 2017, 23, 5631–5651 CrossRef CAS PubMed.
  292. S. M. Rogge, A. Bavykina, J. Hajek, H. Garcia, A. I. Olivos-Suarez, A. Sepúlveda-Escribano, A. Vimont, G. Clet, P. Bazin and F. Kapteijn, Chem. Soc. Rev., 2017, 46, 3134–3184 RSC.
  293. A. Bavykina, N. Kolobov, I. S. Khan, J. A. Bau, A. Ramirez and J. Gascon, Chem. Rev., 2020, 120, 8468–8535 CrossRef CAS PubMed.
  294. K. Rafiq, H. D. Mai, J. K. Kim, J. M. Woo, B. M. Moon, C. H. Park and H. Yoo, Sens. Actuators, B, 2017, 251, 472–480 CrossRef CAS.
  295. F. Ahmad, K. Rafiq, T. Najam, E. Hussain, M. Sohail, M. Z. Abid, A. Mahmood, M. S. Javed and S. S. A. Shah, Int. J. Hydrogen Energy, 2023, 48(90), 35075–35111 CrossRef CAS.
  296. S. Ketrat, T. Maihom, S. Wannakao, M. Probst, S. Nokbin and J. Limtrakul, Inorg. Chem., 2017, 56, 14005–14012 CrossRef CAS PubMed.
  297. P. Zhao, F. Qin, Z. Huang, C. Sun, W. Shen and H. Xu, Chem. Eng. J., 2018, 349, 72–81 CrossRef CAS.
  298. V. Paluka, T. Maihom, M. Probst and J. Limtrakul, Phys. Chem. Chem. Phys., 2020, 22, 13622–13628 RSC.
  299. S. Chen, Y. Zhou, J. Li, Z. Hu, F. Dong, Y. Hu, H. Wang, L. Wang, K. K. Ostrikov and Z. Wu, ACS Catal., 2020, 10, 10185–10196 CrossRef CAS.
  300. H. Louis, T. C. Egemonye, T. O. Unimuke, B. E. Inah, H. O. Edet, E. A. Eno, S. A. Adalikwu and A. S. Adeyinka, ACS Omega, 2022, 7, 34929–34943 CrossRef CAS PubMed.
  301. K. D. Vogiatzis, E. Haldoupis, D. J. Xiao, J. R. Long, J. I. Siepmann and L. Gagliardi, J. Phys. Chem. C, 2016, 120, 18707–18712 CrossRef CAS.
  302. S. Ketrat, T. Maihom, S. Wannakao, M. Probst, S. Nokbin and J. Limtrakul, Inorg. Chem., 2017, 56, 14005–14012 CrossRef CAS PubMed.
  303. V. Paluka, T. Maihom, M. Probst and J. Limtrakul, Phys. Chem. Chem. Phys., 2020, 22, 13622–13628 RSC.
  304. S. Chen, Y. Zhou, J. Li, Z. Hu, F. Dong, Y. Hu, H. Wang, L. Wang, K. Ostrikov and Z. Wu, ACS Catal., 2020, 10(17), 10185–10196 CrossRef CAS.
  305. D. Denysenko, J. Jelic, O. V. Magdysyuk, K. Reuter and D. Volkmer, Microporous Mesoporous Mater., 2015, 216, 146–150 CrossRef CAS.
  306. A. Dey, T. Albert, R. Y. Kong, S. N. MacMillan, P. Moënne-Loccoz, K. M. Lancaster and D. P. Goldberg, Inorg. Chem., 2022, 61, 14909–14917 CrossRef CAS PubMed.
  307. H. Tang, Y. He, P. Liu, J. Shao, F. Lin and Z. Wang, Energy Fuels, 2021, 35, 18664–18679 CrossRef CAS.
  308. Z. Yao, D. Qu, Y. Guo, Y. Yang and H. Huang, Adv. Mater. Sci. Eng., 2019, 2019, 1–9 Search PubMed.
  309. N. J. Halas, P. Nordlander, H. Robatjazi, D. F. Swearer, C. Zhang, H. Zhao and L. Zhou, US Pat., US10766024B2, 2020 Search PubMed.
  310. D. F. Swearer, H. Zhao, L. Zhou, C. Zhang, H. Robatjazi, J. M. P. Martirez, C. M. Krauter, S. Yazdi, M. J. McClain and E. Ringe, Proc. Natl. Acad. Sci. U. S. A., 2016, 113, 8916–8920 CrossRef CAS PubMed.
  311. K. Rafiq, I. Sadia, M. Z. Abid, M. Z. Waleed, A. Rauf and E. Hussain, ACS Biomater. Sci. Eng., 2024, 10, 7268–7313 CrossRef CAS PubMed.
  312. R. Nava, B. Pawelec, P. Castaño, M. C. Álvarez-Galván, C. V. Loricera and J. L. G. Fierro, Appl. Catal., B, 2009, 92, 154–167 CrossRef CAS.
  313. G. J. Kubas, Chem. Rev., 2007, 107, 4152–4205 CrossRef CAS PubMed.
  314. B. Kim, Z. Li, B. D. Kay, Z. Dohnálek and Y. K. Kim, J. Phys. Chem. C, 2014, 118, 9544–9550 CrossRef CAS.
  315. S. Royer, D. Duprez, F. Can, X. Courtois, C. Batiot-Dupeyrat, S. Laassiri and H. Alamdari, Chem. Rev., 2014, 114, 10292–10368 CrossRef CAS PubMed.
  316. I. Fechete, Y. Wang and J. C. Védrine, Catal. Today, 2012, 189, 2–27 CrossRef CAS.
  317. T. Selvaraj, P. Aghalayam and J. J. Varghese, Ind. Eng. Chem. Res., 2024, 63, 6591–6599 CrossRef CAS.
  318. E. Winter, J. Catal., 1969, 15, 144–152 CrossRef CAS.
  319. E. V. Kondratenko and J. Pérez-Ramírez, J. Phys. Chem. B, 2006, 110, 22586–22595 CrossRef CAS PubMed.
  320. S. Subramanian and C. Swamy, Catal. Lett., 1995, 35, 361–372 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.