Next Article in Journal
Photo-Charging a Zinc-Air Battery Using a Nb2O5-CdS Photoelectrode
Next Article in Special Issue
PdPty/V2O5-TiO2: Highly Active Catalysts with Good Moisture- and Sulfur Dioxide-Resistant Performance in Toluene Oxidation
Previous Article in Journal
Improving Photocatalytic Stille Coupling Reaction by CuPd Alloy-Doped Ordered Mesoporous TiO2
Previous Article in Special Issue
Synthesis of Durian-like TiO2@CdS Core-Shell Structure and Study on H2 Generation Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Single-Atom Catalysts: Preparation and Applications in Environmental Catalysis

Beijing Key Laboratory for Green Catalysis and Separation, Key Laboratory of Beijing on Regional Air Pollution Control, Key Laboratory of Advanced Functional Materials, Education Ministry of China, Laboratory of Catalysis Chemistry and Nanoscience, Department of Chemical Engineering and Technology, Faculty of Environment and Life, Beijing University of Technology, Beijing 100124, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1239; https://doi.org/10.3390/catal12101239
Submission received: 14 September 2022 / Revised: 29 September 2022 / Accepted: 10 October 2022 / Published: 14 October 2022
(This article belongs to the Special Issue Exclusive Papers in Environmentally Friendly Catalysis in China)

Abstract

:
Due to the expensive price and the low reserve of noble metals in nature, much attention has been paid to single-atom catalysts (SACs)—especially single-atom noble metal catalysts—owing to their maximum atomic utilization and dispersion. The emergence of SACs greatly decreases the amount of precious metals, improves the catalytic activity, and makes the catalytic process progressively economic and sustainable. However, the most remarkable challenge is the active sites and their stability against migration and aggregation under practical conditions. This review article summarizes the preparation strategies of SACs and their catalytic applications for the oxidation of methane, carbon monoxide, and volatile organic compounds (VOCs) and the reduction of nitrogen oxides. Furthermore, the perspectives and challenges of SACs in future research and practical applications are proposed. It is envisioned that the results summarized in this review will stimulate the interest of more researchers in developing SACs that are effective in catalyzing the reactions related to the environmental pollution control.

1. Introduction

Heterogeneous catalysis is an important research hotspot in the global field. Due to its economic and energy-saving features, it is widely used in the transformation of various energy resources (e.g., petroleum, coal, natural gas, and solar energy) in nature [1,2,3,4], the synthesis of numerous industrial chemicals [5,6], the purification of vehicle exhaust, etc. [7]. It is estimated that catalysts are used in the production processes of about 80% of artificial chemicals at some stages, and about 35% of the world’s gross domestic product (GDP) is accounted for by catalytic processes [8]. In the field of catalysis, exploring the efficient, stable, and economical catalyst formulations has promoted the rapid development of nanomaterials synthesis methods. Traditional heterogeneous catalysts are metal nanoparticles (NPs) dispersed on supports [9]. Among them, noble metals are the most efficient catalysts. However, the reserve of noble metals in nature is only about 3%. Therefore, it is of practical economic value to maximize the utilization of noble metal atoms [10]. In 2011, Zhang et al. defined the concept of SACs: an isolated metal atom anchored by a support as a monatomic active site [11]. Since then, SACs have become a research hotspot in the past decade due to their 100% atomic utilization and maximum dispersion, which greatly improves the efficiency of different catalytic reactions. The alterations in surface and quantum size effects as well as metal−support interactions play important roles in boosting catalytic performance [12].
Currently, the research of SACs is mainly focused on synthesizing stable single atoms on suitable supports. The strategies are divided into “bottom-up” and “top-down” methods according to the difference in precursors. The former is to disperse metal precursors on the support on the atomic scale, and the latter uses metal NPs or bulk metal as precursors to prepare the SACs [13]. Based on the two strategies, many preparation methods have been developed, such as atomic layer deposition (ALD), wet-chemistry, pyrolysis, electrochemical or photochemical methods, and other methods. Different types of supports (e.g., metal oxides, porous metal−organic framework (MOF) materials, two-dimensional layered materials, and so on) are used to form strong chemical bonds between metal atoms and support to improve the stability of SACs in various reactions. For instance, the Pt1/FeOxcatalyst possessed a unique electronic structure of noble metal atoms and iron oxide [11]. The high specific surface area of Ru1/N−C [14] was beneficial for the occurrence of catalytic reactions, and the layered structures [15] were designed to embed single atoms into the support. Single-atom alloy catalysts were developed to further improve the catalytic performance, such as PdAu [16], PtPd [17], and CoNi-NC catalysts [18]. The as-prepared SACs are widely used in carbon monoxide oxidation, methane combustion, hydrogenation, water–gas shift reaction (WGS), hydrogen evolution reaction (HER), carbon dioxide reduction, nitrogen oxide reduction, etc.
In this review, we introduce the advanced strategies of SACs preparation and their applications in environmental catalysis, including methane combustion, carbon monoxide oxidation, volatile organic compound (VOCs) oxidation, and nitrogen oxide reduction (Figure 1). The most remarkable challenge is the active sites and their stability against migration and aggregation under practical conditions. The active component, calcination temperature, treatment atmosphere, support type, and defect site also play important roles in the synthesis process of SACs and directly influence the catalytic efficiency. Furthermore, the problems and perspectives in the scientific research and applications of SACs are presented in detail. The reasonable design of SACs with unique structures, a low cost, and high efficiencies is of great significance in future industrial applications.

2. Synthesis of SACs

2.1. The Atomic Layer Deposition (ALD) Method

Earlier than the proposal of the SACs concept, the ALD technology has already been an optimal choice for coating various metals and metal oxides uniformly on the inner surface of porous nanostructures, such as ZnO, Al2O3, W, Cu, Pt, and Ru [19,20,21,22,23,24]. The ALD method usually includes the following two elementary steps: (i) exposure to a precursor vapor; and (ii) purging the residual precursors and by-products. In the former step, all of the active sites in the support interact with the precursors, and then extra reactants and by-products are swept with an inert gas. Next, another cycle starts, which can realize the required film thickness or design of specific materials by controlling the ALD cycle times and the deposition temperature [23,24]. This is also the unique advantage of the ALD method. In recent years, the selected metal deposition of graphene materials via the ALD route was rapidly developed [25,26,27,28,29,30,31,32,33]; the metal atoms could be precisely deposited into the defect of graphene, such as the crystalline phase interface [25]. CeO2 [34], SiO2 [35], and other metal oxides have also been developed as supports of the SACs. Figure 2a shows that the atomically dispersed Fe1(OH)x catalyst was fabricated on Pt NPs. Atom-level control was achieved by using the self-limiting and selective deposition. However, the ALD technology is generally only suitable for laboratory investigations and not for commercial purposes due to its high cost and low catalyst output, although a number of small-scale trials have been implemented [36,37].

2.2. The Wet-Chemistry Method

The wet-chemistry method includes the impregnation, co-precipitation, deposition-precipitation, and so on. Using the precursors containing metal atoms can effectually prevent monatomic aggregation, which is thanks to the metal-support interaction.

2.2.1. The Impregnation Method

The impregnation method is the most promising method for the industrial production of SACs due to its simple operation and cheap raw reagents, but the uniform dispersion of single atoms is still a challenge. It was found that the dispersity and loading content of the metal depends on the ability of a support to adsorb precursors containing the metal atoms during the impregnation process. It is worth mentioning that a support rich in oxygen vacancies is more favorable for stabilizing single atoms, such as TiO2 [38,39], an amorphous MoOx phase with abundant defects [40], MnOx with an Ru−O−MnOx interface [41], CeO2 nanorods rich in Ce3+ [42,43,44], and SnO2 [45]. For example, specific surface functional groups (e.g., −OH) facilitate the electrostatic adsorption of the metal precursors, hence favoring the anchoring of the individual Pt atoms (Figure 2b). Furthermore, numerous investigations have found that supports with high surface areas are also beneficial for stabilizing the single atoms [15,46,47,48].

2.2.2. The Co-Precipitation Method

Many parameters in the process of preparing SACs by the co-precipitation method can influence the dispersion of the metal atoms, such as the order and speed of adding precursors, the temperature and pH value of the solution, the ageing time, and so on. A variety of SACs were successfully prepared using the co-precipitation strategy, and their catalytic performance was usually better than their NPs [49,50,51,52,53]. Except for the support with abundant oxygen vacancies, the obvious electron transfer between the metal atoms and support facilitates the regeneration of oxygen vacancies, which plays an important role in catalytic reactions. For example, the Pt sites occupy the higher electronic state and show a strong affinity for the orbitals of neighboring Co atoms, thus promoting the electron transfer between Pt and Co atoms and ulteriorly increasing the number of oxygen vacancies on the catalyst surface [54]. The interaction between Pt atoms and the support can be adjusted by doping the reducible Fe2O3 to the irreducible Al2O3 by means of the co-precipitation approach [55]. However, the shortcoming of such a method is that the partial atoms would be restricted to the inside of the support, and the atomic utilization does not reach 100%.

2.2.3. The Deposition–Precipitation (DP) Method

The operation sequence of the deposition–precipitation method is as follows: The target metal solution is added to the support suspension, and a mixed uniform suspension is formed after the reaction, thus depositing metal hydroxide or carbonate on the support. Generally speaking, this method easily obtains small-sized NPs or clusters; the SACs with a high dispersion are also synthesized via precisely controlling the reaction temperature and pH value of the precursor solutions. Although the loading content of the metal atoms obtained by this strategy is relatively low, the as-obtained samples exhibit excellent properties; therefore, this method is also widely used [56,57,58,59,60]. The active sites efficaciously inhibit carbon deposition [58,59], and the strong metal–support interactions make metal single atoms more strongly bound to the lattice oxygen of the support [60]. For the successful preparation of the SACs by the DP method, it is hence very important to accurately regulate the anchoring sites on the support surface and the metal–support interaction.

2.3. The Pyrolysis Approach

The pyrolysis approach usually employs the coordination-assisted strategy to achieve the preparation of SACs, especially for preparing the carbon-based SACs. Metal−organic frames (MOFs), as ideal precursors, possess a coordination environment with the well-defined atomically dispersed metal sites and an adjustable structure. For example, as shown in Figure 2c, the existence of −NH2 prevents Ru in the precursors from aggregating to Ru clusters in the nitrogen-doped porous carbon-supported ruthenium catalyst (Ru SAs/N−C), which takes advantage of the strong coordination of Ru3+ with −NH2 on the metal−organic skeleton [14]. Various single noble metal (Ru, Au, Pd, Ir, or Pt) atoms have been successfully immobilized at anchoring sites of the MOF-like hollow nanotubes [61], which further demonstrates that the SACs have a maximum atomic efficiency, unique hollow structure, and excellent mass transfer efficiency [62,63,64]. Such a structurally tunable synthesis strategy shows great potential in improving the catalytic performance of the SACs.

2.4. The Electrochemical/Photochemical Method

The electrochemical/photochemical method refers to adsorbing a mononuclear metal precursor on the surface of a support by means of the electrochemical or photochemical process. Using ultraviolet light irradiation [65,66,67] or electrochemical potential deposition [68,69,70] could prepare high loadings of SACs [65,71]. The atoms are inserted into the lattice of the support via transferring the photoelectrons in a photochemical liquid-phase reaction. For instance, high-energy deep-ultraviolet photons induced photochemical cleavage and activated the metal and oxygen atoms to promote the generation of the M−O−M bond [68]. Constant potentials and special treatments are used to activate the catalyst for the optimum performance and steady state, such as in situ anchoring Ni atoms to the carbon support via carbonizing the as-synthesized Ni-MOF at 700 °C in the nitrogen atmosphere and then activating it by the electrochemical approach (Figure 2d). However, the shortcoming is that the atoms will be aggregated if the rates of nucleus formation and growth cannot be reasonably adjusted [72]. Thus, the photochemical reduction-freezing strategy was developed to successfully stabilize noble metal atoms [73]. At present, this method has been extended to other oxide supports, except for carbon materials and TiO2.

2.5. Other Synthesis Methods

Other synthesis methods for the efficient preparation of SACs include mass-selected soft-landing [74,75,76], the microwave-assisted method [77,78,79], ball-milling [80,81], laser ablation [82], and so on. A variety of SACs have been fabricated via these advanced synthesis routes. As shown in Figure 2e, the solid amine-functionalized graphene oxide (AGO) containing Co2+ was irradiated at 1000 W for 2 s; then, the Co-NG-MW catalyst was obtained, and the irradiation created a high-energy environment which was favorable for the simultaneous reduction and N-doping of GO, thus producing the defects or vacancies to anchor the metal atoms. Table 1 summarizes the supports, active components, loadings, and catalytic applications of some typical SACs in recent years. However, the commercial application of SACs has been hindered, so it is necessary to further develop SACs with distinct structures, good stability, and high loadings.
Figure 2. Schematic illustrations of preparing (a) FeOx-Pt/SiO2 by the ALD method. Reprinted/adapted with permission from Ref. [35]. Copyright 2019, copyright Springer Nature Limited, (b) Pt-SA/α-MoOx by the impregnation method. Reprinted/adapted with permission from Ref. [40]. Copyright 2020, copyright Elsevier Ltd, (c) Ru SAs/N−C by the pyrolysis approach. Reprinted/adapted with permission from Ref. [14]. Copyright 2017, copyright American Chemical Society, (d) Ni–C catalysts by the electrochemical method. Reprinted/adapted with permission from Ref. [72]. Copyright 2020, copyright Willy-VCH GmbH, and (e) Co-NG-MW by the microwave-assisted method. Reprinted/adapted with permission from Ref. [78]. Copyright 2018, copyright Willy-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 2. Schematic illustrations of preparing (a) FeOx-Pt/SiO2 by the ALD method. Reprinted/adapted with permission from Ref. [35]. Copyright 2019, copyright Springer Nature Limited, (b) Pt-SA/α-MoOx by the impregnation method. Reprinted/adapted with permission from Ref. [40]. Copyright 2020, copyright Elsevier Ltd, (c) Ru SAs/N−C by the pyrolysis approach. Reprinted/adapted with permission from Ref. [14]. Copyright 2017, copyright American Chemical Society, (d) Ni–C catalysts by the electrochemical method. Reprinted/adapted with permission from Ref. [72]. Copyright 2020, copyright Willy-VCH GmbH, and (e) Co-NG-MW by the microwave-assisted method. Reprinted/adapted with permission from Ref. [78]. Copyright 2018, copyright Willy-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 12 01239 g002

3. Catalytic Applications of SACs

A great number of studies on SACs have rapidly emerged in recent decades. In general, the detailed surface structure of a single atom can be obtained by a scanning tunneling microscope (STM) and aberration-corrected scanning transmission electron microscopy (STEM). X-ray absorption near edge spectra (XANES) and extended X-ray absorption fine structure spectra (EXAFS) can provide the valence states of the metal atoms and the properties and oxidation states of the neighbor species, which are the most effective ways to determine metal coordination numbers and judge whether SACs have been successfully prepared. Nuclear magnetic resonance spectroscopy (NMR) is used to provide coordinate information of metal species and ligands during in situ reactions. Density functional theory (DFT) calculations can reveal the geometrical configuration and strong interaction of the SACs according to the bond lengths, bond energies, electron transfer, binding energies, orbital hybridization, and density of the states. SACs exhibit excellent catalytic activity and high stability in heterogeneous catalytic reactions owing to their specific geometries, unique electronic structures, and strong interactions. The properties and development potential of SACs in the aforementioned reactions are listed in Table 2 and described below in detail.

3.1. CO Oxidation

CO oxidation is the most widely studied field of SACs. As early as 1999, researchers studied CO oxidation over the Pt/MgO catalyst and put forward the correlation between monodisperse Pt cluster size and catalytic activity [74]. The concept of SACs was first proposed by Zhang and colleagues when they concluded that partially vacant orbitals of high-valence Pt atoms contributed to reducing the adsorption energy of CO and the activation energy barrier of CO oxidation, thereby increasing the catalytic activity [11]. At present, the most commonly studied catalysts in SACs take Pd and Au as the active sites; FeOx [11], CeO2 [34,57,97], TiO2 [98], Co3O4 [59], and Al2O3 [88] are used as the supports for CO oxidation. Generally speaking, the catalytic performance of CO oxidation over SACs is better than that of its corresponding NPs, such as Pt1/FeOx [11], Pt/m-Al2O3 [88], and Au1/Co3O4 [89]. The activity of CO oxidation over the Pt/m-Al2O3 SAC was still maintained after 60 circles of testing in a month, while the temperature went up dramatically after 10 testing circles of the NPs catalyst [88]. The excellent activity and high stability of CO oxidation over the Au1/Co3O4 SAC depend upon the interaction between Au1 atoms and neighboring Co or O atoms (or oxygen vacancies), not only the individual Au1 atoms [89]. Nevertheless, SACs do not always exhibit a better performance than NPs catalysts—for example, the catalytic activity of Pt SAC/SiO2 is lower than that of Pt NPs/SiO2 [99]. The Pt SACs prepared by electrostatic adsorption and high-temperature atomic capture were not active for CO oxidation at low temperatures, but the CO oxidation activities of both catalysts were improved after the catalysts were treated in a CO atmosphere at 275 °C, which was mainly attributed to the reduction in lattice oxygen in the CeO2 support [100]. Another study reported that the active sites were the Pt−O−Pt sites, rather than the Pt−O−Ce interface. The Pt−O−Pt sites were more likely to activate oxygen molecules, and its CO oxidation activity was 100–1000 times higher than that of Pt1/CeO2 [101]. At the same time, the oxidation activity of CO was related to the migration of Pt sites to the position of clusters containing a small number of Pt atoms [102]. The reduced Au sites (Au0), which are the pivotal active sites in CO oxidation, demonstrate a higher CO adsorption capacity compared with the completely oxidized Au species [103]. In conclusion, the advantages of SACs in CO oxidation depend upon the interaction between metal atoms and support, but they cannot be simply generalized.
The activities of CO oxidation can also be further improved by doping other components to modified SACs. First, 1 wt% Ce was introduced to modify the TiO2 support particles, which could stabilize Pt single atoms through the strong electronic interaction. The as-formed CeOx−TiO2 mixed oxide interface could stabilize Pt single atoms through the strong electronic interaction. Compared with the Pt/TiO2 catalyst, the mass activity of the compact Pt SAC formed on CeOx/TiO2 mixed oxide increased by about 15 times [98]. Fe atoms, displaying a similar oxygen affinity to Pt, were inserted in nitrogen-doped graphene to prevent CO poisoning [90]. On the atomic scale, Luo et al. visualized the evolution of the AuCu alloy NPs supported on CeO2 during the CO oxidation process and found that CO exposure led to the formation of the metal-CO bonds, while O2 exposure resulted in the formation of the Cu2O–AuCu interface, thereby promoting CO–O interaction [97].

3.2. Methane Combustion

Methane is abundant in natural reserve, and it is also the main component of natural gas and biogas. It is used as low-value fuel at all times, and its emission leads to increased methane concentration in the atmosphere, not only causing a waste of resources but also aggravating the greenhouse effect and bringing great environmental pressure. Therefore, it is desirable to study the complete oxidation of methane and convert methane to high-value chemicals for air pollution control and the utilization of low-carbon alkanes, respectively. In view of the highly symmetric tetrahedral stability and strong C–H bond energy of CH4, the activation and directional transformation of CH4 molecules have been hot topics in the field of catalysis.
On the one hand, the present literature concentrates on the high-temperature tolerance [56] and water resistance [91,92] of SACs on CH4 combustion. The activity and stability were improved several times when Pt NPs were converted into Pt atoms in situ [56]. By combining the experiments with the DFT calculations, the volatilization of Pt at less than 600 °C is found to be thermally neutral or slightly endothermic, indicating that temperature plays a critical role in the dispersion of Pt NPs, that is, monoatomic dispersion can only be achieved at a temperature higher than 600 °C. Doping Fe2O3 into Al2O3 is beneficial to the dispersion of Pt atoms, because the reducible Fe2O3 displays a stronger interaction with Pt atoms than the irreducible Al2O3 displays with Pt atoms. In fact, a large amount of water vapor in the exhaust gas of methane combustion in actual situations will influence the performance of the Pd catalyst, even inducing deactivation. As shown in Figure 3a, two-dimensional Pt rafts were formed on the surface of cerium oxide (2Pt/1Pt@CeO2), which showed a higher reactivity in CH4 oxidation because of the weak interaction between the Pt atoms and support. However, deactivation after introducing water gas is inevitable (Figure 3b). The deactivation was mainly attributed to the generation and migration of the Pd–OH species strongly adsorbed on the PdO surface, and the surface active sites were poisoned due to the binding of water vapor to PdO, which was substantiated by the fact that the surface coverage of CH4 on PdO (101) in the presence of water vapor is only 3% of the CH4 saturation coverage [91]. In view of such a phenomenon, 1Pd/2Pt@CeO2 derived from the atomic trapping route exhibited higher CH4 combustion reactivity and better water resistance than 1Pd/CeO2, 2Pt@CeO2 and (1Pd + 2Pt)/CeO2. Hence, the addition of Pt atoms is favorable for forming the PtPd bimetallic atoms and can stabilize the Pd species in a metallic state, thus improving the water vapor resistance. The W single atoms-modified PdO (PdW1/Al2O3) catalyst [92] exhibited better water resistance than the traditional catalysts. When 3–10 vol% water vapor was first added to the reaction system and then cut off, the catalytic activity of PdW1/Al2O3 first decreased and then recovered (Figure 3c). Interestingly, as the testing time got longer, the CH4 conversion even exceeded the initial conversion, which was completely different from the activity of the traditional Pd catalyst (Figure 3d), which is due to the fact that water molecules participate in methane combustion in the form of the peroxy hydroxyl group, and W atoms change the electronic state of the PdO particles, thus regulating the adsorption energy and mobility of oxygen or water molecules. In general, the development of stable and efficient catalysts for CH4 catalytic combustion still needs further exploration—from the perspective of either catalyst preparation strategies or reaction mechanisms.

3.3. Volatile Organic Compounds (VOCs) Oxidation

Given that noble metal catalysts are of high cost, SACs show great potential in many oxidation reactions, such as the oxidation of alcohol [46,54,83], formaldehyde [85,86,87], and benzene/toluene [47,80,84]. These studies were mainly focused on Pt, Pd, Ag, and other SACs. The extensive applications of metal SACs in VOCs oxidation will be introduced in detail.
First of all, a proper reducible support is critical to successfully preparing SACs. Except for the excellent pollutant removal efficiency, maintaining stability for a long time is also important to the catalytic performance of SACs. As is well known, the metal–support interaction helps to maintain the thermodynamic stable state of SACs. For the supports with more oxygen vacancies, the oxygen vacancies strongly combine with the metal atoms and are difficult to replenish in time, resulting in a drop in catalytic activity. For the supports with a low number of oxygen vacancies, the binding force to the noble metal atoms is too weak to stabilize single atoms, so supports such as CeO2 [83], MnOx [84,85], and Co3O4 [54] were rapidly developed for the thermocatalytic oxidation of VOCs. For instance, Pt atoms are attached to the Co3O4 (111) crystal plane and occupy part of the Co2+ atomic sites. The electrons are transferred from Pt to Co sites, thus increasing the proportion of oxygen vacancies on the Pt1-Co3O4 surface. The regeneration of these oxygen vacancies reduces the adsorption energy of methanol and significantly promotes the dissociation of the C−H bond in methanol oxidation [54]. Furthermore, the majority SACs showed a better performance than their NPs counterparts in many VOCs oxidation reactions. For example, the reaction rate of 0.25Pt1/meso-Fe2O3 for benzene oxidation was seven times higher than that of the NPs sample, which was undoubtedly related to the good utilization of Pt atoms [47]. The 0.03 wt% Pd/Al2O3 SAC was the most active cinnamyl alcohol oxidation catalyst reported so far; the activity of each Pd atom increased by 30 times if the 4.7 wt% Pd cluster catalyst was prepared to the 0.03 wt% Pd SAC catalyst [46]. The above-mentioned cases indicate that the metal–support interaction in SACs highlights the advantages in the oxidation reaction. Interestingly, Jiang et al. further discovered that the Pt−O−Ce interface in Pt1/CeO2 SAC could spontaneously deform to balance the Fermi energy and charge density between Pt and the support, and the changes in the valence state and electronic structure increased the adsorption of oxygen and methanol, which was the essential reason for the excellent catalytic activity and thermal stability of the Pt1/CeO2 catalyst in the oxidation of oxygenated hydrocarbons [83].
Secondly, modifying the catalyst to form hydroxyl radicals can promote the oxidative activity for pollutant removal. Over the Pt/MnO2 SAC, a 100% conversion of toluene (0.42 ppm) was obtained at room temperature, and the oxygen vacancies and strong oxidizing hydroxyl radicals were considered to be the main reasons for its outstanding stability and excellent activity [84]. Moreover, introducing alkali metal ions (e.g., Li+, Na+, and K+) to Pt/TiO2 SAC promoted the highly efficient removal of formaldehyde. The formed PtO(OH)x-alkali metal species, as a promoter for the dispersion of Pt atoms, remarkably promoted the activity of formaldehyde oxidation by activating the simple reaction between hydroxyl and formate species on the surface [86].
In addition, the loading of metal or metal oxides, the addition of water vapor, and the special structure of the catalyst also seriously affect the performance of the catalyst. Pt/Mn−TiO2, with excellent activity and an acceptable cost, was screened as the best catalyst for formaldehyde oxidation. The loading of Pt enhanced the reactivity of surface lattice oxygen [87]. However, it is worth noting that only 0.5 wt% Pt loading can completely oxidize formaldehyde at low temperatures and achieve long-term stability under simulated conditions. The addition of water vapor can suppress the side reaction and favor the deep oxidation of formaldehyde. The MnOx nanorods-supported monoatomic Ag chain catalyst possessed a good activation ability for molecular and lattice oxygen at low temperatures [85], in which the Ag atoms were initially uniformly dispersed and then thermally migrated to the end of the support to generate a single-atom Ag chain. In general, SACs have shown promising application prospects in many fields. A number of targeted strategies have been developed to improve the efficiency of SACs, especially those aiming at multicomponent VOCs, which are still under development.

3.4. NOx Reduction

Nitrogen oxides (NOx, x = 1 or 2), derived from the exhaust emissions of diesel vehicles, are one of the important precursors of the atmospheric pollutants PM2.5 and the ozone. Their large amount of emissions lead to environmental pollution (e.g., haze, acid rain, and photochemical smog). The most effective NOx removal technology is using NH3 for the selective catalytic reduction of NOx (NH3-SCR) to generate environmentally friendly N2 and H2O. The key issue is the successful preparation of high-performance SCR catalysts with a high N2 selectivity.
The Pt group catalysts have a great potential for NOx emission control [94,105]. The 0.06 wt% Pt single-atom catalyst (Pt-SAC) could efficiently eliminate NO in a wide temperature range and showed a high N2 selectivity. Compared with the NPs catalyst, Pt-SAC possessed more oxygen vacancies for NO adsorption and the breaking ability of the N−O bond [105]. Pd SACs exhibited a high selectivity for converting NO into nitrate and an excellent resistance to nitrate poisoning [94]. However, when the Pd loading increased to more than 0.055 wt%, Pd atom clusters or particles will form and significantly decrease the activity. This indicates that the catalytic performance of SACs is governed by the content of noble metals.
Transition-metal SACs have also received much attention [95,106] in NOx reduction. Acidic metal oxides (e.g., Nb2O5, WO3, and MoO3)-modified CeO2 catalysts show excellent NH3-SCR performance at medium and high temperatures, especially the Nb2O5−CeO2 catalyst. Ulteriorly, Xie et al. prepared a Cu atom-promoted Nb2O5/CeO2 catalyst (NbCuCe), and the formed strong interaction of Nb−O−Cu and Cu−O−Ce effectively promoted the activation of reactant molecules [95]. After aging under actual reaction conditions, the NbCuCe catalyst showed better activity than the commercial Cu-CHA catalyst. Such a catalyst shows good application prospects in diesel vehicle exhaust low-temperature denitrification.
The Mo1/Fe2O3 catalyst has been used to study the active sites and reaction mechanisms of NOx reduction [96,106]. The formation of acidic-redox double active sites was caused by the Mo atoms and the surrounding Fe atoms (Figure 4a,b). The SCR performance was linearly related to the number of active catalytic sites (ACS) (Figure 4c−g). The redox ability of the binuclear W1−Fe1 site in Fe1/WO3 was much weaker than that in W1/Fe2O3, which mainly led to an SCR performance of Fe1/WO3 that was inferior to that of W1/Fe2O3 (Figure 4g). Thus, modulating the acidity or redox properties of W1/Fe2O3 and Fe1/WO3 SACs can improve the SCR activity [106]. In addition, commercial catalysts are often inactivated due to the occupation of the active sites by the formed NH4HSO4 in the presence of SO2. The acidic (Mo) sites and basic (Ti) sites in the Mo1/TiO2 catalyst adsorbed NH4+ and HSO42− in NH4HSO4, respectively, thus decomposing NH4HSO4 at 225 °C and releasing SO2 at low temperatures without causing the poisoning of the catalyst [96]. The exploration of the active sites for NOx reduction and the reaction mechanisms under complex industrial conditions will help to develop high-performance catalysts.

4. Conclusive Remarks and Perspectives

Recent studies have shown that SACs have been rapidly developed in heterogeneous catalysis. The preparation methods of SACs and their applications in different reactions were briefly reviewed in this paper. The importance of support selection, catalytic performance, active sites, and reaction mechanisms was discussed. Due to their unique electronic structures, unsaturated coordination states, quantum size effects, and metal–support interactions, SACs have the following advantages over their nanocluster or nanoparticle counterparts:
(i)
Efficiency: SACs have achieved the maximum atomic utilization efficiency and have made great progress in CO oxidation, methane oxidation, VOCs oxidation, and NOx reduction. In general, SACs are usually more active than their NPs catalysts, mainly due to the reducibility of the support (i.e., the number of oxygen vacancies, which is critical for anchoring single atoms) and the presence of OH groups on the support surface.
(ii)
Economy: The high atomic utilization reduces the amount of precious metal used in the catalyst and effectively reduces the cost.
(iii)
Uniqueness: The single active sites of SACs are helpful for further studying the active sites and catalytic mechanisms of various reactions.
However, there are the following challenges:
(i)
At present, there are many strategies to synthesize SACs, such as the ALD, wet chemical, and pyrolysis methods. However, the loading of monatomic metals is limited, so it is necessary to explore a common strategy to achieve the scale-up production of SACs with high metal loadings, a satisfactory catalyst output, and good activity.
(ii)
The agglomeration of single atoms is still a common phenomenon, and the influence of inorganic molecules in the reaction atmosphere under actual industrial conditions should also be considered, such as the presence of H2O, CO2, SO2, and HCl. Selecting a suitable metal to dope into a support or form an alloy structure can effectively modify the electronic structure of SACs and enhance the single atoms–support interaction, which is a good pathway to maintain the stability of SACs.
(iii)
SACs are currently limited to traditional pollutants (e.g., VOCs) in the atmosphere and need to be further expanded to other fields, such as the synthesis of high-value chemicals, the dehydrogenation of VOCs, the resource utilization of CO2, the purification of water, and so on.
(iv)
Optimizing the catalyst structures to improve catalytic performance: Since the atomic structure has a direct effect on reactivity, it is highly desirable to use in situ characterization techniques to study the catalyst structures so as to design the catalysts rationally. The catalyst support with a high surface area has a complex nanostructure and abundant surface active sites, which is beneficial for improving the catalytic performance.
(v)
In addition to the development of in situ STM, STEM, and XAFS as well as the DFT theoretical calculations, more techniques need to be developed to recognize single atoms and further identify controversial active sites and mechanisms. The understanding of catalyst structures has a guiding role for the design of highly efficient catalysts. In addition, the current DFT calculations can only provide theoretical models of simple pollutant molecules, and more computational methods need to be explored to evaluate the reaction mechanisms of complex molecules.
In conclusion, SACs show great potential in the environmental remediation field. Although certain advancements have been achieved in fabrication strategies, performance evaluation, characterization techniques, and mechanism research, there are still some challenges, which are mentioned above. The reasonable design of SACs with a clear structure, a high efficiency, economy, and simple and controllable preparation methods is of great significance in future potential industrial applications. It is envisioned that this review article can draw more researchers’ attention to achieving the goal of controlling single atoms.

Author Contributions

Conceptualization, H.D. and J.D.; methodology, Y.L. and L.J., writing—original draft preparation, X.Y., W.P. and Z.H.; writing—review and editing, X.Y. and H.D.; funding acquisition, H.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Committee of China—Liaoning Provincial People’s Government Joint Fund (No. U1908204), the National Natural Science Foundation of China (Nos. 21876006, 21976009, and 21961160743), the Foundation on the Creative Research Team Construction Promotion Project of Beijing Municipal Institutions (No. IDHT20190503), the Natural Science Foundation of Beijing Municipal Commission of Education (No. KM201710005004), and the Development Program for the Youth Outstanding—Notch Talent of Beijing Municipal Commission of Education (No. CIT&TCD201904019).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhu, C.Z.; Du, D.; Eychmuller, A.; Lin, Y.H. Engineering ordered and nonordered porous noble metal nanostructures: Synthesis, assembly, and their applications in electrochemistry. Chem. Rev. 2015, 115, 8896–8943. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, X.Q.; Iocozzia, J.; Wang, Y.; Cui, X.; Chen, Y.H.; Zhao, S.Q.; Li, Z.; Lin, Z.Q. Noble metal−metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy Environ. Sci. 2017, 10, 402–434. [Google Scholar] [CrossRef]
  3. Cai, P.W.; Li, Y.; Wang, G.X.; Wen, Z.H. Alkaline-acid Zn-H2O fuel cell for the simultaneous generation of hydrogen and electricity. Angew. Chem. Int. Ed. 2018, 57, 3910–3915. [Google Scholar] [CrossRef] [PubMed]
  4. Li, Z.Y.; Liu, H.T.; Liu, Y.; He, P.; Li, J.H. A room-temperature ionic liquid-templated proton-conducting gelatinous electrolyte. J. Phys. Chem. B 2004, 108, 17512–17518. [Google Scholar] [CrossRef]
  5. Furstner, A. Gold and platinum catalysis—A convenient tool for generating molecular complexity. Chem. Soc. Rev. 2009, 38, 3208–3221. [Google Scholar] [CrossRef]
  6. Zheng, N.F.; Zhang, T. Preface: Single-atom catalysts as a new generation of heterogeneous catalysts. Nat. Sci. Rev. 2018, 5, 625. [Google Scholar] [CrossRef]
  7. Zhang, S.R.; Nguyen, L.; Liang, J.X.; Shan, J.J.; Liu, J.Y.; Frenkel, A.I.; Patlolla, A.; Huang, W.X.; Li, J.; Tao, F. Catalysis on singly dispersed bimetallic sites. Nat. Commun. 2015, 6, 7938. [Google Scholar] [CrossRef] [Green Version]
  8. Youngs, T.G.A.; Manyar, H.; Bowron, D.T.; Gladden, L.F.; Hardacre, C. Probing chemistry and kinetics of reactions in heterogeneous catalysts. Chem. Sci. 2013, 4, 3484. [Google Scholar] [CrossRef] [Green Version]
  9. Bell, A.T. The impact of nanoscience on heterogeneous. Catal. Sci. 2003, 299, 1688. [Google Scholar]
  10. Hunt, S.T.; Román-Leshkov, Y. Principles and methods for the rational design of core–shell nanoparticle catalysts with ultralow noble metal loadings. Acc. Chem. Res. 2018, 51, 1054. [Google Scholar] [CrossRef]
  11. Qiao, B.T.; Wang, A.Q.; Yang, X.F.; Allard, L.F.; Jiang, Z.; Cui, Y.T.; Liu, J.Y.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 2011, 3, 634. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, J.Y. Catalysis by supported single metal atoms. ACS Catal. 2017, 7, 34–59. [Google Scholar] [CrossRef]
  13. Ji, S.F.; Chen, Y.J.; Wang, X.L.; Zhang, Z.D.; Wang, D.S.; Li, Y.D. Chemical synthesis of single atomic site catalysts. Chem. Rev. 2020, 120, 11900–11955. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, X.; Chen, W.X.; Zhang, L.; Yao, T.; Liu, W.; Lin, Y.; Ju, H.X.; Dong, J.C.; Zheng, L.R.; Yan, W.S.; et al. Uncoordinated amine groups of metal−organic frameworks to anchor single Ru sites as chemoselective catalysts toward the hydrogenation of quinoline. J. Am. Chem. Soc. 2017, 139, 9419–9422. [Google Scholar] [CrossRef] [PubMed]
  15. Li, P.S.; Wang, M.Y.; Duan, X.X.; Zheng, L.R.; Cheng, X.P.; Zhang, Y.F.; Kuang, Y.; Li, Y.P.; Ma, Q.; Feng, Z.X.; et al. Boosting oxygen evolution of single-atomic ruthenium through electronic coupling with cobalt iron layered double hydroxides. Nat. Commun. 2019, 10, 1711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Liu, J.L.; Uhlman, M.B.; Montemore, M.M.; Trimpalis, A.; Giannakakis, G.; Shan, J.J.; Cao, S.F.; Hannagan, R.T.; Sykes, E.C.H. Integrated catalysis-surface science-theory approach to understand selectivity in the hydrogenation of 1-hexyne to 1-hexene on PdAu single-atom alloy catalysts. ACS Catal. 2019, 9, 8757–8765. [Google Scholar] [CrossRef]
  17. Zhang, L.; Liu, H.S.; Liu, S.H.; Banis, M.N.; Song, Z.X.; Li, J.J.; Yang, L.J.; Markiewicz, M.; Zhao, Y.; Li, R.Y.; et al. Pt/Pd single-atom alloys as highly active electrochemical catalysts and the origin of enhanced activity. ACS Catal. 2019, 9, 9350–9358. [Google Scholar] [CrossRef]
  18. He, Q.; Liu, D.B.; Lee, J.H.; Liu, Y.M.; Xie, Z.H.; Hwang, S.; Kattel, S.; Song, L.; Chen, J.C. Electrochemical conversion of CO2 to syngas with controllable CO/H2 ratios over Co and Ni single-atom catalysts. Angew. Chem. Int. Ed. 2020, 59, 3033–3037. [Google Scholar] [CrossRef]
  19. Kucheyev, S.O.; Biener, J.; Wang, Y.M.; Baumann, T.F.; Wu, K.J.; Buuren, T.V.; Hamza, A.V.; Satcher, J.H., Jr.; Elam, J.W.; Pellin, M.J. Atomic layer deposition of ZnO on ultralow-density nanoporous silica aerogel monoliths. Appl. Phys. Lett. 2005, 86, 083108. [Google Scholar] [CrossRef] [Green Version]
  20. Baumann, T.F.; Biener, J.; Wang, Y.M.M.; Kucheyev, S.O.; Nelson, E.J.; Satcher, J.H., Jr.; Elam, J.W.; Pellin, M.J.; Hamza, A.V. Atomic layer deposition of uniform metal coatings on highly porous aerogel substrates. Chem. Mater. 2006, 18, 6106–6108. [Google Scholar] [CrossRef]
  21. Biener, J.; Baumann, T.F.; Wang, Y.M.; Nelson, E.J.; Kucheyev, S.O.; Hamza, A.V.; Kemell, M.; Ritala, M.; Leskelä, M. Ruthenium/aerogel nanocomposites via atomic layer deposition. Nanotechnology. 2007, 18, 055303. [Google Scholar] [CrossRef]
  22. Kucheyev, S.O.; Biener, J.; Baumann, T.F.; Wang, Y.M.; Hamza, A.V.; Li, Z.; Lee, D.K.; Gordon, R.G. Mechanisms of atomic layer deposition on substrates with ultrahigh aspect ratios. Langmuir 2008, 24, 943–948. [Google Scholar] [CrossRef] [PubMed]
  23. King, J.S.; Wittstock, A.; Biener, J.; Kucheyev, S.O.; Wang, Y.M.; Baumann, T.F.; Giri, S.K.; Hamza, A.V.; Baeumer, M.; Bent, S.F. Ultralowloading Pt nanocatalysts prepared by atomic layer deposition on carbon aerogels. Nano Lett. 2008, 8, 2405–2409. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, C.; Wang, C.C.; Kei, C.C.; Hsueh, Y.C.; Perng, T.P. Atomic layer deposition of platinum nanoparticles on carbon nanotubes for application in proton-exchange membrane fuel cells. Small 2009, 5, 1535–1538. [Google Scholar] [CrossRef] [PubMed]
  25. Kim, K.; Lee, H.B.R.; Johnson, R.W.; Tanskanen, J.T.; Liu, N.; Kim, M.G.; Pang, C.; Ahn, C.; Bent, S.F.; Bao, Z. Selective metal deposition at graphene line defects by atomic layer deposition. Nat. Commun. 2014, 5, 4781. [Google Scholar] [CrossRef] [Green Version]
  26. Yan, H.; Zhao, X.X.; Guo, N.; Lyu, Z.Y.; Du, Y.H.; Xi, S.B.; Guo, R.; Chen, C.; Chen, Z.X.; Liu, W.; et al. Atomic engineering of high-density isolated Co atoms on graphene with proximal-atom controlled reaction selectivity. Nat. Commun. 2018, 9, 3197. [Google Scholar] [CrossRef] [Green Version]
  27. Yan, H.; Lin, Y.; Wu, H.; Zhang, W.H.; Sun, Z.H.; Cheng, H.; Liu, W.; Wang, C.L.; Li, J.J.; Huang, X.H.; et al. Bottom-up precise synthesis of stable platinum dimers on graphene. Nat. Commun. 2017, 8, 1070. [Google Scholar] [CrossRef] [Green Version]
  28. Yan, H.; Cheng, H.; Yi, H.; Lin, Y.; Yao, T.; Wang, C.L.; Li, J.J.; Wei, S.Q.; Lu, J.J. Single-atom Pd1/graphene catalyst achieved by atomic layer deposition: Remarkable performance in selective hydrogenation of 1,3-butadiene. J. Am. Chem. Soc. 2015, 137, 10484–10487. [Google Scholar] [CrossRef]
  29. Sun, S.H.; Zhang, G.X.; Gauquelin, N.; Chen, N.; Zhou, J.G.; Yang, S.L.; Chen, W.F.; Meng, X.B.; Geng, D.S.; Banis, M.N.; et al. Single-stom catalysis using Pt/graphene achieved through atomic layer deposition. Sci. Rep. 2013, 3, 1775. [Google Scholar] [CrossRef] [Green Version]
  30. Cao, Y.J.; Chen, S.; Luo, Q.Q.; Yan, H.; Lin, Y.; Liu, W.; Cao, L.L.; Lu, J.L.; Yang, J.L.; Yao, T.; et al. Atomic-level insight into optimizing the hydrogen evolution pathway over a Co1-N4 single-site photocatalyst. Angew. Chem. Int. Ed. 2017, 56, 12191–12196. [Google Scholar] [CrossRef]
  31. Zhang, L.; Si, R.T.; Liu, H.S.; Chen, N.; Wang, Q.; Adair, K.; Wang, Z.Q.; Chen, J.T.; Song, Z.X.; Li, J.J.; et al. Atomic layer deposited Pt−Ru dual-metal dimers and identifying their active sites for hydrogen evolution reaction. Nat. Commun. 2019, 10, 4936. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, L.W.; Long, R.; Zhang, Y.M.; Duan, D.L.; Xiong, Y.J.; Zhang, Y.J.; Bi, Y.P. Direct observation of dynamic bond evolution in single atom Pt/C3N4 catalysts. Angew. Chem. Int. Ed. 2020, 59, 6224–6229. [Google Scholar] [CrossRef] [PubMed]
  33. Cheng, N.C.; Stambula, S.; Wang, D.; Banis, M.N.; Liu, J.; Riese, A.; Xiao, B.W.; Li, R.Y.; Sham, T.K.; Liu, L.M.; et al. Platinum single-atom and cluster catalysis of the hydrogen evolution reaction. Nat. Commun. 2016, 7, 13638. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, C.L.; Gu, X.K.; Yan, H.; Lin, Y.; Li, J.J.; Liu, D.D.; Li, W.X.; Lu, J.L. Water-mediated Mars−van Krevelen mechanism for CO oxidation on ceria-supported single-atom Pt1 catalyst. ACS Catal. 2017, 7, 887–891. [Google Scholar] [CrossRef]
  35. Cao, L.N.; Liu, W.; Luo, Q.Q.; Yin, R.T.; Wang, B.; Weissenrieder, J.; Soldemo, M.; Yan, H.; Lin, Y.; Sun, Z.H.; et al. Atomically dispersed iron hydroxide anchored on Pt for preferential oxidation of CO in H2. Nature 2019, 565, 631–635. [Google Scholar] [CrossRef] [PubMed]
  36. O’Neill, B.J.; Jackson, D.H.K.; Lee, J.; Canlas, C.; Stair, P.C.; Marshall, C.L.; Elam, J.M.; Kuech, T.F.; Dumesic, J.A.; Huber, G.W. Catalyst design with atomic layer deposition. ACS Catal. 2015, 5, 1804–1825. [Google Scholar] [CrossRef] [Green Version]
  37. Fonseca, J.; Lu, J.L. Single-atom catalysts designed and prepared by the atomic layer deposition technique. ACS Catal. 2021, 11, 7018–7059. [Google Scholar] [CrossRef]
  38. Liu, S.Z.; Wang, Y.J.; Wang, S.B.; You, M.M.; Hong, S.; Wu, T.S.; Soo, Y.L.; Zhao, Z.Q.; Jiang, G.Y.; Qiu, J.S.; et al. Photocatalytic fixation of nitrogen to ammonia by single Ru atom decorated TiO2 nanosheets. ACS Sustain. Chem. Eng. 2019, 7, 6813–6820. [Google Scholar] [CrossRef]
  39. Wan, J.W.; Chen, W.X.; Jia, C.Y.; Zheng, L.R.; Dong, J.C.; Zheng, X.S.; Wang, Y.; Yan, W.S.; Chen, C.; Peng, Q.; et al. Defect effects on TiO2 nanosheets: Stabilizing single atomic site Au and promoting catalytic properties. Adv. Mater. 2018, 30, 1705369. [Google Scholar] [CrossRef]
  40. Xu, J.; Zhang, C.X.; Liu, H.X.; Sun, J.Q.; Xie, R.C.; Qiu, Y.; Lu, F.; Liu, Y.F.; Zhuo, L.C.; Liu, X.J.; et al. Amorphous MoOx-stabilized single platinum atoms with ultrahigh mass activity for acidic hydrogen evolution. Nano Energy 2020, 70, 104529. [Google Scholar] [CrossRef]
  41. Wang, H.; Xu, D.Y.; Guan, E.J.; Wang, L.; Zhang, J.; Wang, C.T.; Wang, S.; Xu, H.; Meng, X.J.; Yang, B.; et al. Atomically dispersed Ru on manganese oxide catalyst boosts oxidative cyanation. ACS Catal. 2020, 10, 6299–6308. [Google Scholar] [CrossRef]
  42. Chen, J.Y.; Wanyan, Y.J.; Zeng, J.X.; Fang, H.H.; Li, Z.J.; Dong, Y.D.; Qin, R.X.; Wu, C.Z.; Liu, D.Y.; Wang, M.Z.; et al. Surface engineering protocol to obtain an atomically dispersed Pt/CeO2 catalyst with high activity and stability for CO oxidation. ACS Sustain. Chem. Eng. 2018, 6, 14054–14062. [Google Scholar] [CrossRef]
  43. Guo, Y.; Mei, S.; Yuan, K.; Wang, D.J.; Liu, H.C.; Yan, C.H.; Zhang, Y.W. Low-temperature CO2 methanation over CeO2-supported Ru single atoms, nanoclusters, and nanoparticles competitively tuned by strong metal−support interactions and H-spillover effect. ACS Catal. 2018, 8, 6203–6215. [Google Scholar] [CrossRef]
  44. Feng, Y.; Dai, L.Y.; Wang, Z.W.; Peng, Y.; Duan, E.H.; Liu, Y.X.; Jing, L.; Wang, X.; Rastegarpanah, A.; Dai, H.X.; et al. Photothermal synergistic effect of Pt1/CuO-CeO2 single-atom catalysts significantly improving toluene removal. Environ. Sci. Technol. 2022, 56, 8722–8732. [Google Scholar] [CrossRef] [PubMed]
  45. Kim, J.; Roh, C.W.; Sahoo, S.K.; Yang, S.; Bae, J.; Han, J.W.; Lee, H. Highly durable platinum single-atom alloy catalyst for electrochemical reactions. Adv. Energy Mater. 2018, 8, 1701476. [Google Scholar] [CrossRef]
  46. Hackett, S.F.J.; Brydson, R.M.; Gass, M.H.; Harvey, I.; Newman, A.D.; Wilson, K.; Lee, A.F. High-activity, single-site mesoporous Pd/Al2O3 catalysts for selective aerobic oxidation of allylic alcohols. Angew. Chem. Int. Ed. 2007, 46, 8593–8596. [Google Scholar] [CrossRef] [PubMed]
  47. Yang, K.; Liu, Y.X.; Deng, J.G.; Zhao, X.T.; Yang, J.; Han, Z.; Hou, Z.Q.; Dai, H.X. Three-dimensionally ordered mesoporous iron oxide-supported single-atom platinum: Highly active catalysts for benzene combustion. Appl. Catal. B 2019, 244, 650–659. [Google Scholar] [CrossRef]
  48. Kwak, J.H.; Kovarik, L.; Szanyi, J. CO2 reduction on supported Ru/Al2O3 catalysts: Cluster size dependence of product selectivity. ACS Catal. 2013, 3, 2449–2455. [Google Scholar] [CrossRef]
  49. Yu, W.Z.; Wang, W.W.; Li, S.Q.; Fu, X.P.; Wang, X.; Wu, K.; Si, R.; Ma, C.; Jia, C.J.; Yan, C.H. Construction of active site in a sintered copper-ceria nanorod catalyst. J. Am. Chem. Soc. 2019, 141, 17548–17557. [Google Scholar] [CrossRef]
  50. Zhang, D.; Luo, J.J.; Wang, J.J.; Xiao, X.; Liu, Y.F.; Qi, W.; Su, D.S.; Chu, W. Ru/FeOx catalyst performance design: Highly dispersed Ru species for selective carbon dioxide hydrogenation. Chin. J. Catal. 2018, 39, 157–166. [Google Scholar] [CrossRef]
  51. Millet, M.M.; Algara-Siller, G.; Wrabetz, S.; Mazheika, A.; Girgsdies, F.; Teschner, D.; Seitz, F.; Tarasov, A.; Leychenko, S.V.; Schloegl, R.; et al. Ni single atom catalysts for CO2 activation. J. Am. Chem. Soc. 2019, 141, 2451–2461. [Google Scholar] [CrossRef] [PubMed]
  52. Lin, J.; Wang, A.Q.; Qiao, B.T.; Liu, X.Y.; Yang, X.F.; Wang, X.D.; Liang, J.X.; Li, J.; Liu, J.Y.; Zhang, T. Remarkable performance of Ir1/FeOx single atom catalyst in water gas shift reaction. J. Am. Chem. Soc. 2013, 135, 15314–15317. [Google Scholar] [CrossRef] [PubMed]
  53. Wei, H.S.; Liu, X.Y.; Wang, A.Q.; Zhang, L.L.; Qiao, B.T.; Yang, Y.F.; Huang, Y.Q.; Miao, S.; Liu, J.Y.; Zhang, T. FeOx-supported platinum single-atom and pseudo-single-atom catalysts for chemoselective hydrogenation of functionalized nitroarenes. Nat. Commun. 2014, 5, 5634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Jiang, Z.Y.; Feng, X.B.; Deng, J.L.; He, C.; Douthwaite, M.; Yu, Y.K.; Liu, J.; Hao, Z.P.; Zhao, Z. Atomic-scale insights into the low-temperature oxidation of methanol over a single-atom Pt1-Co3O4 catalyst. Adv. Funct. Mater. 2019, 29, 1902041. [Google Scholar] [CrossRef]
  55. Lang, R.; Xi, W.; Liu, J.C.; Cui, Y.T.; Li, T.B.; Lee, A.F.; Chen, F.; Chen, Y.; Li, L.; Li, L.; et al. Non defect-stabilized thermally stable single-atom catalyst. Nat. Commun. 2019, 10, 234. [Google Scholar] [CrossRef] [Green Version]
  56. Qiao, B.T.; Liu, J.X.; Wang, Y.G.; Lin, Q.Q.; Liu, X.Y.; Wang, A.Q.; Li, J.; Zhang, T.; Liu, J.Y. Highly efficient catalysis of preferential oxidation of CO in H2-rich stream by gold single-atom catalysts. ACS Catal. 2015, 5, 6249–6254. [Google Scholar] [CrossRef]
  57. Wang, H.; Shen, J.H.; Huang, J.F.; Xu, T.J.; Zhu, J.R.; Zhu, Y.H.; Li, C.Z. Atomically dispersed Au catalysts supported on CeO2 foam: Controllable synthesis and CO oxidation reaction mechanism. Nanoscale 2017, 9, 16817–16825. [Google Scholar] [CrossRef]
  58. Guo, X.G.; Fang, G.Z.; Li, G.; Ma, H.; Fan, H.J.; Yu, L.; Ma, C.; Wu, X.; Deng, D.H.; Wei, M.M.; et al. Direct, nonoxidative conversion of methane to ethylene, aromatics, and hydrogen. Science 2014, 344, 616–619. [Google Scholar] [CrossRef]
  59. Qiao, B.T.; Lin, J.; Wang, A.Q.; Chen, Y.; Zhang, T.; Liu, J.Y. Highly active Au1/Co3O4 single-atom catalyst for CO oxidation at room temperature. Chin. J. Catal. 2015, 36, 1505–1511. [Google Scholar] [CrossRef]
  60. Gu, X.K.; Qiao, B.T.; Huang, C.Q.; Ding, W.C.; Sun, K.J.; Zhan, E.S.; Zhang, T.; Liu, J.Y.; Li, W.X. Supported single Pt1/Au1 atoms for methanol steam reforming. ACS Catal. 2014, 4, 3886–3890. [Google Scholar] [CrossRef]
  61. He, T.; Chen, S.M.; Ni, B.; Gong, Y.; Wu, Z.; Song, L.; Gu, L.; Hu, W.P.; Wang, X. Zirconium–porphyrin-based metal–organic framework hollow nanotubes for immobilization of noble-metal single atoms. Angew. Chem. Int. Ed. 2018, 57, 3493. [Google Scholar] [CrossRef] [PubMed]
  62. Vile, G.; Albani, D.; Nachtegaal, M.; Chen, Z.P.; Dontsova, D.; Antonietti, M.; Lopez, N.; Perez-Ramírez, J. A stable single-site palladium catalyst for hydrogenations. Angew. Chem. Int. Ed. 2015, 54, 11265–11269. [Google Scholar] [CrossRef] [PubMed]
  63. Fei, H.L.; Dong, J.C.; Arellano-Jimenez, M.J.; Ye, G.L.; Kim, N.D.; Samuel, E.L.; Peng, Z.W.; Zhu, Z.; Qin, F.; Bao, J.M.; et al. Atomic cobalt on nitrogen-doped graphene for hydrogen generation. Nat. Commun. 2015, 6, 8668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Xi, J.B.; Sung, H.Y.; Wang, D.; Zhang, Z.Y.; Duan, X.M.; Xiao, J.W.; Xiao, F.; Liu, L.M.; Wang, S. Confined-interface-directed synthesis of palladium single-atom catalysts on graphene/amorphous carbon. Appl. Catal. B 2018, 225, 291. [Google Scholar] [CrossRef] [Green Version]
  65. Yang, M.; Allard, L.F.; Flytzani-Stephanopoulos, M. Atomically dispersed Au−(OH)x species bound on titania catalyze the low temperature water-gas shift reaction. J. Am. Chem. Soc. 2013, 135, 3768–3771. [Google Scholar] [CrossRef]
  66. Ge, X.X.; Zhou, P.; Zhang, Q.H.; Xia, Z.H.; Chen, S.L.; Gao, P.; Zhang, Z.; Gu, L.; Guo, S.J. Palladium single atoms on TiO2 as a photocatalytic sensing platform for analyzing organophosphorus pesticide chlorpyrifos. Angew. Chem. Int. Ed. 2019, 131, 232–236. [Google Scholar]
  67. Liu, P.X.; Zhao, Y.; Qin, R.X.; Mo, S.G.; Chen, G.X.; Gu, L.; Chevrier, D.M.; Zhang, P.; Guo, Q.; Zang, D.D.; et al. Photochemical route for synthesizing atomically dispersed palladium catalysts. Science 2016, 352, 797–800. [Google Scholar] [CrossRef] [Green Version]
  68. Kim, Y.H.; Heo, J.S.; Kim, T.H.; Park, S.; Yoon, M.H.; Kim, J.; Oh, M.S.; Yi, G.R.; Noh, Y.Y.; Park, S.K. Flexible metal-oxide devices made by room-temperature photochemical activation of sol-gel films. Nature 2012, 489, 128–132. [Google Scholar] [CrossRef]
  69. Zhang, Z.R.; Feng, C.; Liu, C.X.; Zuo, M.; Qin, L.; Yan, X.P.; Xing, Y.L.; Li, H.L.; Si, R.; Zhou, S.M.; et al. Electrochemical deposition as a universal route for fabricating single-atom catalysts. Nat. Commun. 2020, 11, 1215. [Google Scholar] [CrossRef] [Green Version]
  70. Fan, L.; Liu, P.F.; Yan, X.C.; Gu, L.; Yang, Z.Z.; Yang, H.G.; Qiu, S.L.; Yao, X.D. Atomically isolated nickel species anchored on graphitized carbon for efficient hydrogen evolution electrocatalysis. Nat. Commun. 2016, 7, 10667. [Google Scholar] [CrossRef] [Green Version]
  71. Liu, P.X.; Zhao, Y.; Qin, R.X.; Gu, L.; Zhang, P.; Fu, G.; Zheng, N.F. A vicinal effect for promoting catalysis of Pd1/TiO2: Supports of atomically dispersed catalysts play more roles than simply serving as ligands. Sci. Bull. 2018, 63, 675–682. [Google Scholar] [CrossRef] [Green Version]
  72. Liu, T.T.; Gao, W.B.; Wang, Q.Q.; Dou, M.; Zhang, Z.P.; Wang, F. Selective loading of atomic Pt on a RuCeOx support enables stable hydrogen evolution at high current densities. Angew. Chem. Int. Ed. 2020, 59, 20423–20427. [Google Scholar] [CrossRef] [PubMed]
  73. Wei, H.H.; Huang, K.; Wang, D.; Zhang, R.Y.; Ge, B.H.; Ma, J.Y.; Wen, B.; Zhang, S.; Li, Q.Y.; Lei, M.; et al. Iced photochemical reduction to synthesize atomically dispersed metals by suppressing nanocrystal growth. Nat. Commun. 2017, 8, 1490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Heiz, U.; Sanchez, A.; Abbet, S.; Schneider, W.D. Catalytic oxidation of carbon monoxide on monodispersed platinum clusters:  Each atom counts. J. Am. Chem. Soc. 1999, 121, 3214–3217. [Google Scholar] [CrossRef]
  75. Vajda, S.; White, M.G. Catalysis applications of size-selected cluster deposition. ACS Catal. 2015, 5, 7152–7176. [Google Scholar] [CrossRef]
  76. Li, X.N.; Yuan, Z.; He, S.G. CO oxidation promoted by gold atoms supported on titanium oxide cluster anions. J. Am. Chem. Soc. 2014, 136, 3617–3623. [Google Scholar] [CrossRef]
  77. Bilecka, I.; Niederberger, M. Microwave chemistry for inorganic nanomaterials synthesis. Nanoscale 2010, 2, 1358–1374. [Google Scholar] [CrossRef]
  78. Fei, H.L.; Dong, J.C.; Wan, C.Z.; Zhao, Z.P.; Xu, X.; Lin, Z.Y.; Wang, Y.L.; Liu, H.T.; Zang, K.T.; Luo, J.; et al. Microwave-assisted rapid synthesis of graphene-supported single atomic metals. Adv. Mater. 2018, 30, 1802146. [Google Scholar] [CrossRef]
  79. Yao, Y.G.; Huang, Z.N.; Xie, P.F.; Wu, L.P.; Ma, L.; Li, T.Y.; Pang, Z.Q.; Jiao, M.L.; Liang, Z.Q.; Gao, J.L.; et al. High temperature shockwave stabilized single atoms. Nat. Nanotechnol. 2019, 14, 851–857. [Google Scholar] [CrossRef]
  80. Cui, X.J.; Xiao, J.P.; Wu, Y.H.; Du, P.P.; Si, R.; Yang, H.X.; Tian, H.F.; Li, J.Q.; Zhang, W.H.; Deng, D.H.; et al. A graphene composite material with single cobalt active sites: A highly efficient counter electrode for dye-sensitized solar cells. Angew. Chem. Int. Ed. 2016, 55, 6708–6712. [Google Scholar] [CrossRef]
  81. Deng, D.H.; Chen, X.Q.; Yu, L.; Wu, X.; Liu, Q.F.; Liu, Y.; Yang, H.X.; Tian, H.F.; Hu, Y.F.; Du, P.P.; et al. A single iron site confined in a graphene matrix for the catalytic oxidation of benzene at room temperature. Sci. Adv. 2015, 1, e1500462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Zhao, Y.X.; Li, Z.Y.; Yuan, Z.; Li, X.N.; He, S.G. Thermal methane conversion to formaldehyde promoted by single platinum atoms in PtAl2O4 cluster anions. Angew. Chem. Int. Ed. 2014, 53, 9482–9486. [Google Scholar] [CrossRef] [PubMed]
  83. Jiang, Z.Y.; Jing, M.Z.; Feng, X.B.; Xiong, J.C.; He, C.; Douthwaitee, M.; Zheng, L.R.; Song, W.Y.; Liu, J.; Qu, Z.G. Stabilizing platinum atoms on CeO2 oxygen vacancies by metal-support interaction induced interface distortion: Mechanism and application. Appl. Catal. B 2020, 278, 119304. [Google Scholar] [CrossRef]
  84. Zhang, H.Y.; Sui, S.H.; Zheng, X.M.; Cao, R.R.; Zhang, P.Y. One-pot synthesis of atomically dispersed Pt on MnO2 for efficient catalytic decomposition of toluene at low temperatures. Appl. Catal. B 2019, 257, 117878. [Google Scholar] [CrossRef]
  85. Huang, Z.W.; Gu, X.; Cao, Q.Q.; Hu, P.P.; Hao, J.M.; Li, J.H.; Tang, X.F. Catalytically active single-atom sites fabricated from silver particles. Angew. Chem. Int. Ed. 2012, 51, 4198–4203. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, C.B.; Liu, F.D.; Zhai, Y.P.; Ariga, H.; Yi, N.; Liu, Y.C.; Asakura, K.; Flytzani-Stephanopoulos, M.; He, H. Alkali-metal-promoted Pt/TiO2 opens a more efficient pathway to formaldehyde oxidation at ambient temperatures. Angew. Chem. Int. Ed. 2012, 51, 9628–9632. [Google Scholar] [CrossRef]
  87. Chen, J.; Jiang, M.Z.; Xu, W.J.; Chen, J.; Hong, Z.X.; Jia, H.P. Incorporating Mn cation as anchor to atomically disperse Pt on TiO2 for low temperature removal of formaldehyde. Appl. Catal. B 2019, 259, 118013. [Google Scholar] [CrossRef]
  88. Zhang, Z.L.; Zhu, Y.H.; Asakura, H.; Zhang, B.; Zhang, J.G.; Zhou, M.X.; Han, Y.; Tanaka, T.; Wang, A.Q.; Zhang, T.; et al. Thermally stable single atom Pt/m-Al2O3 for selective hydrogenation and CO oxidation. Nat. Commun. 2017, 8, 16100. [Google Scholar] [CrossRef] [Green Version]
  89. Lou, Y.; Cai, Y.F.; Hu, W.D.; Wang, L.; Dai, Q.G.; Zhan, W.C.; Guo, Y.L.; Hu, P.; Cao, X.M.; Liu, J.Y.; et al. Identification of active area as active center for CO oxidation over single Au atom catalyst. ACS Catal. 2020, 10, 6094–6101. [Google Scholar] [CrossRef]
  90. Kropp, T.; Mavrikakis, M. Transition metal atoms embedded in graphene: How nitrogen doping increases CO oxidation sctivity. ACS Catal. 2019, 9, 6864–6868. [Google Scholar] [CrossRef]
  91. Xiong, H.F.; Kunwar, D.; Jiang, D.; García-Vargas, C.E.; Li, H.Y.; Du, C.C.; Canning, G.; Pereira-Hernandez, X.I.; Wan, Q.; Lin, S.; et al. Engineering catalyst supports to stabilize PdOx two-dimensional rafts for water-tolerant methane oxidation. Nat. Catal. 2021, 4, 830–839. [Google Scholar] [CrossRef]
  92. Hou, Z.Q.; Dai, L.Y.; Deng, J.G.; Zhao, G.F.; Jing, L.; Wang, Y.S.; Yu, X.H.; Gao, R.Y.; Tian, X.R.; Dai, H.X.; et al. Electronically engineering water resistance in methane combustion with an atomically dispersed tungsten on PdO catalyst. Angew. Chem. Int. Ed. 2022, 61, e202201655. [Google Scholar] [CrossRef] [PubMed]
  93. Luo, L.H.; Luo, J.; Li, H.L.; Ren, F.N.; Zhang, Y.F.; Liu, A.D.; Li, W.X.; Zeng, J. Water enables mild oxidation of methane to methanol on gold single-atom catalysts. Nat. Commun. 2021, 12, 1218. [Google Scholar] [CrossRef] [PubMed]
  94. Fujiwara, K.; Pratsini, S.E. Single Pd atoms on TiO2 dominate photocatalytic NOx removal. Appl. Catal. B 2018, 226, 127–134. [Google Scholar] [CrossRef]
  95. Xie, S.H.; Tan, W.; Li, Y.J.; Ma, L.; Ehrlich, S.N.; Deng, J.G.; Xu, P.; Gao, F.; Dong, L.; Liu, F.D. Copper single atom-triggered niobia−ceria catalyst for efficient low-temperature reduction of nitrogen oxides. ACS Catal. 2022, 12, 2441–2453. [Google Scholar] [CrossRef]
  96. Qu, W.Y.; Liu, X.N.; Chen, J.X.; Dong, Y.Y.; Tang, X.F.; Chen, Y.X. Single-atom catalysts reveal the dinuclear characteristic of active sites in NO selective reduction with NH3. Nat. Commun. 2020, 11, 1532. [Google Scholar] [CrossRef] [Green Version]
  97. Luo, L.L.; Chen, S.Y.; Xu, Q.; He, Y.; Dong, Z.J.; Zhang, L.F.; Zhu, J.F.; Du, Y.G.; Yang, B.; Wang, C.M. Dynamic atom clusters on AuCu nanoparticle surface during CO oxidation. J. Am. Chem. Soc. 2020, 142, 4022–4027. [Google Scholar] [CrossRef]
  98. Yoo, M.; Yu, Y.S.; Ha, H.; Lee, S.; Choi, J.S.; Oh, S.; Kang, E.; Choi, H.; An, H.; Lee, K.S.; et al. A tailored oxide interface creates dense Pt single-atom catalysts with high catalytic activity. Energy Environ. Sci. 2020, 13, 1231–1239. [Google Scholar] [CrossRef]
  99. Ding, K.L.; Gulec, A.; Johnson, A.M.; Schweitzer, N.M.; Stucky, G.D.; Marks, L.D.; Stair, P.C. Identification of active sites in CO oxidation and water-gas shift over supported Pt catalysts. Science 2015, 350, 189. [Google Scholar] [CrossRef] [Green Version]
  100. Pereira-Hernández, X.I.; DeLaRiva, A.; Muravev, V.; Kunwar, D.; Xiong, H.F.; Sudduth, B.; Engelhard, M.; Kovarik, L.; Hensen, E.J.M.; Wang, Y.; et al. Tuning Pt-CeO2 interactions by high-temperature vapor-phase synthesis for improved reducibility of lattice oxygen. Nat. Commun. 2019, 10, 1358. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, H.; Liu, J.X.; Allard, L.F.; Lee, S.; Liu, J.L.; Li, H.; Wang, J.Q.; Wang, J.; Oh, S.H.; Li, W.; et al. Surpassing the single-atom catalytic activity limit through paired Pt−O−Pt ensemble built from isolated Pt1 atoms. Nat. Commun. 2019, 10, 3808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Maurer, F.; Jelic, J.; Wang, J.; Gänzler, A.; Dolcet, P.; Wöll, C.; Wang, Y.; Studt, F.; Casapu, M.; Grunwaldt, J.D. Tracking the formation, fate and consequence for catalytic activity of Pt single sites on CeO2. Nat. Catal. 2020, 3, 824–833. [Google Scholar] [CrossRef]
  103. Guo, L.W.; Du, P.P.; Fu, X.P.; Ma, C.; Zeng, J.; Si, R.; Huang, Y.Y.; Jia, C.J.; Zhang, Y.W.; Yan, C.H. Contributions of distinct gold species to catalytic reactivity for carbon monoxide oxidation. Nat. Commun. 2016, 7, 13481. [Google Scholar] [CrossRef] [PubMed]
  104. Tang, X.; Wang, L.; Yang, B.; Fei, C.; Yao, T.Y.; Liu, W.; Lou, Y.; Dai, Q.G.; Cai, Y.F.; Cao, X.M.; et al. Direct oxidation of methane to oxygenates on supported single Cu atom catalyst. Appl. Catal. B 2021, 285, 119827. [Google Scholar] [CrossRef]
  105. Lin, J.; Qiao, B.T.; Li, N.; Li, L.; Sun, X.C.; Liu, J.Y.; Wang, X.D.; Zhang, T. Little do more: A highly effective Pt1/FeOx single-atom catalyst for the reduction of NO by H2, Chem. Commun. 2015, 51, 7911–7914. [Google Scholar] [CrossRef]
  106. Chen, J.X.; Fang, X.; Ren, Z.H.; Qu, W.Y.; Hu, X.L.; Ma, Z.; Chen, L.W.; Liu, X.; Chen, Y.X.; Tang, X.F. Single Mo atoms paired with neighboring Ti atoms catalytically decompose ammonium bisulfate formed in low-temperature SCR. J. Mater. Chem. A 2022, 10, 6065. [Google Scholar] [CrossRef]
Figure 1. Supports, preparation strategies, and applications of SACs in environmental catalysis.
Figure 1. Supports, preparation strategies, and applications of SACs in environmental catalysis.
Catalysts 12 01239 g001
Figure 3. (a) Light-off curves of CH4 combustion under dry conditions. (b) Time-on-stream measurements for CH4 combustion at 500 °C under different steam concentrations over 1Pd/2Pt@CeO2. Reprinted/adapted with permission from Ref. [91]. Copyright 2021, copyright Springer Nature Limited. The effect of water vapor concentration on the catalytic activity and selectivity of (c) PdW1/Al2O3 and (d) Pd/Al2O3 for methane combustion. Reprinted/adapted with permission from Ref. [92]. Copyright 2022, copyright Willy-VCH Verlag GmbH. On the other hand, SACs are also used to convert CH4 into high-value chemicals, such as methanol [93,104], ethylene, and aromatic hydrocarbons [58]. The general strategy in industry is to first convert CH4 into syngas (CO and H2) and then into higher-value products, but this process is energy-intensive and tends to produce by-products, such as CO2. So, how should methane be directly oxidized to higher-value chemicals under mild conditions? How can we ensure the high methane conversion and high selectivity of higher-value products? These are worldwide problems that attract much attention in chemistry. From the perspective of active sites, Cu1/ZSM-5 SAC can realize high CH4 conversion and high methanol selectivity (99%) because the active sites preferentially activate CH4 over methanol [104]. The Au1/BP (black phosphorus) catalyst showed excellent activity and selectivity in the photocatalytic oxidation of CH4 to methanol, and water plays an important role [93]. In addition to oxidizing methane to another chemical, Guo et al. proposed a non-oxidative pathway by which to convert methane to ethylene and aromatic hydrocarbons over Fe/SiO2 SAC. At 1090 °C, the methane conversion, ethylene selectivity, and total hydrocarbon selectivity were 48.1, 48.4, and more than 99%, respectively. This catalyst also showed good stability, and no deactivation took place within 60 h of continuous reaction, since the absence of the adjacent Fe sites suppressed the catalytic C−C coupling and ensured that the methyl radical was to be formed without the deposition of solid carbon [58]. As mentioned above, studies on the active sites and catalytic performance for CH4 combustion and selective oxidation are still being explored, which are of great significance for the world to control air pollution problems and resource utilization.
Figure 3. (a) Light-off curves of CH4 combustion under dry conditions. (b) Time-on-stream measurements for CH4 combustion at 500 °C under different steam concentrations over 1Pd/2Pt@CeO2. Reprinted/adapted with permission from Ref. [91]. Copyright 2021, copyright Springer Nature Limited. The effect of water vapor concentration on the catalytic activity and selectivity of (c) PdW1/Al2O3 and (d) Pd/Al2O3 for methane combustion. Reprinted/adapted with permission from Ref. [92]. Copyright 2022, copyright Willy-VCH Verlag GmbH. On the other hand, SACs are also used to convert CH4 into high-value chemicals, such as methanol [93,104], ethylene, and aromatic hydrocarbons [58]. The general strategy in industry is to first convert CH4 into syngas (CO and H2) and then into higher-value products, but this process is energy-intensive and tends to produce by-products, such as CO2. So, how should methane be directly oxidized to higher-value chemicals under mild conditions? How can we ensure the high methane conversion and high selectivity of higher-value products? These are worldwide problems that attract much attention in chemistry. From the perspective of active sites, Cu1/ZSM-5 SAC can realize high CH4 conversion and high methanol selectivity (99%) because the active sites preferentially activate CH4 over methanol [104]. The Au1/BP (black phosphorus) catalyst showed excellent activity and selectivity in the photocatalytic oxidation of CH4 to methanol, and water plays an important role [93]. In addition to oxidizing methane to another chemical, Guo et al. proposed a non-oxidative pathway by which to convert methane to ethylene and aromatic hydrocarbons over Fe/SiO2 SAC. At 1090 °C, the methane conversion, ethylene selectivity, and total hydrocarbon selectivity were 48.1, 48.4, and more than 99%, respectively. This catalyst also showed good stability, and no deactivation took place within 60 h of continuous reaction, since the absence of the adjacent Fe sites suppressed the catalytic C−C coupling and ensured that the methyl radical was to be formed without the deposition of solid carbon [58]. As mentioned above, studies on the active sites and catalytic performance for CH4 combustion and selective oxidation are still being explored, which are of great significance for the world to control air pollution problems and resource utilization.
Catalysts 12 01239 g003
Figure 4. (a) AC-STEM image of Mo1/Fe2O3, (b) intensity surface plot and the corresponding structural model of the selected area, (ce) EDX mappings and (f) AC-STEM image of W1/Fe2O3, (g) SCR rates of W1/Fe2O3 and Fe1/WO3 with different ACS numbers at 270 °C [96].
Figure 4. (a) AC-STEM image of Mo1/Fe2O3, (b) intensity surface plot and the corresponding structural model of the selected area, (ce) EDX mappings and (f) AC-STEM image of W1/Fe2O3, (g) SCR rates of W1/Fe2O3 and Fe1/WO3 with different ACS numbers at 270 °C [96].
Catalysts 12 01239 g004
Table 1. The SACs are prepared by the ALD, wet-chemistry, pyrolysis, electrochemical/photochemical, and other synthesis methods.
Table 1. The SACs are prepared by the ALD, wet-chemistry, pyrolysis, electrochemical/photochemical, and other synthesis methods.
MethodCatalystLoaded MetalSupportCatalytic ReactionRef.
The ALD methodCo1/grapheneCo (2.50 wt%)GrapheneHydrogenation of nitroarenes[26]
Pd1/GraphenePd (0.25 wt%)GrapheneSelective hydrogenation of 1,3-butadiene[28]
Co1-N4/g-C3N4Co (1.00 wt%)g-C3N4Photocatalytic hydrogen production[30]
Pt/C3N4Pt (0.14–0.32 wt%)C3N4photocatalytic hydrogen production[32]
Pt1/CeO2Pt (0.22 wt%)CeO2CO oxidation[34]
Fe−Pt/SiO2Fe (0–0.75 wt%)SiO2Preferential oxidation of CO in hydrogen (PROX)[35]
The impregnation methodRu/TiO2Ru (1.00 wt%)TiO2Photocatalytic fixation of nitrogen to ammonia[38]
Au-SA/TiO2Au (0.25 wt%)TiO2CO oxidation[39]
Pt-SA/α-MoOxPt (0.28 wt%)α-MoOxHydrogen evolution reaction (HER)[40]
Ru/MnO2-rRu (0.10 wt%)MnO2-rOxidative cyanation of alcohols[41]
Pt/CeO2Pt (1.00 wt%)CeO2CO oxidation[42]
Ru/CeO2Ru (0.89 wt%)CeO2CO2 methanation[43]
Pt1/CuO−CeO2Pt (0.39 wt%)CuO−CeO2Photothermal catalytic oxidation of toluene[44]
Pt1/ATOPt (1.0, 4.0, 8.0 wt%)SnO2Formic acid oxidation reaction[45]
Pd/Al2O3-Al2O3Selective aerobic oxidation of allylic alcohols[46]
Pt1/meso-Fe2O3Pt (0.15 wt%)Fe2O3Benzene combustion[47]
Ru/CoFe-LDHsRu (0.45 wt%)CoFe-LDHsOxygen evolution reaction.[15]
Ru/Al2O3Ru (0.1–5.0 wt%)Al2O3CO2 reduction[48]
The co-precipitation methodCuCeCu (1.00 wt%)CeO2CO oxidation[49]
Ni/MgONi (0–18.7 wt%)MgOReverse water–gas shift (rWGS)[51]
Ir1/FeOxIr (0.01 wt%)FeOxWater–gas shift reaction (WGS)[52]
Pt1/FeOxIr (0.08 wt%)FeOxChemoselective hydrogenation[53]
Pt1-Co3O4Pt (0.02–0.30 wt%)Co3O4Methanol oxidation[54]
Pt/Fe2O3Pt (0.30 wt%)Fe2O3Methane oxidation[55]
The deposition–precipitation methodAu1/CeO2Au (0.05, 0.30 wt%)CeO2Preferential oxidation of CO (PROX)[56]
Au1/Co3O4Au (0.05 wt%)Co3O4CO oxidation[59]
Pt1/ZnO
Au1/ZnO
Pt (0.0125 wt%)
Au (0.0125 wt%)
ZnOMethanol steam reforming[60]
The pyrolysis approachRu SAs/N−CRu (0.30 wt%)N−CChemoselective hydrogenation[14]
HNTM-Ir/PtIr (1.05 wt%),
Pt (2.54 wt%)
MOF (HNTM)Photocatalytic H2 evolution[61]
Co-N-grapheneCo (0.57 at%)N-grapheneHydrogen generation[63]
RGO@AC/PdPd (0.43–0.87 wt%)RGO@AC (carbon)4-Nitrophenol (4-NP) reduction[64]
The electrochemical/photochemical methodAu–(OH)x/TiOxAu (~1.0 wt%)TiOxWGS[65]
Pd SA/TiO2Pd (0.53 wt%)TiO2HER[66]
Pd1/TiO2Pd (1.50 wt%)TiO2Hydrogenation of aldehydes[67]
Ir1/Co(OH)2Ir (1.2, 2.0 wt%)Co(OH)2HER[69]
Ni−CarbonNi (1.50 wt%)CarbonHER[70]
Pd1/TiO2Pd (1.20 wt%)TiO2Toluene and methane oxidation[71]
Pt/RuCeOxPt (0.49 wt%)RuCeOxHER[72]
Pt1/MCPt (2.60 wt%)CarbonHER[73]
Other synthesis methodsMicrowave-assistedCo-NG-MWCo (1.10 wt%)N-grapheneHER[78]
Shock-heatingPt-CA-CNFPt (0.24 wt%)CarbonCO oxidation, direct methane conversion[79]
Ball millingCoN4/GNCo (8.40%)GrapheneInterconversion of the redox couple I/I3−[80]
FeN4/GNFe (1.50 wt%)GrapheneBenzene oxidation[81]
Mass-selectedAux(TiO2)yOz-TiO2CO oxidation[76]
Laser ablationPtAlxOy-Pt/Al mixed metalThermal methane conversion to formaldehyde[82]
Table 2. Summary of the SACs for applications in environmental catalysis.
Table 2. Summary of the SACs for applications in environmental catalysis.
ReactionSACs CatalystCatalytic ActivityStabilityRef.
Photothermocatalytic oxidation of toluenePt1/CuO-CeO2Toluene conversion at 160−180 °C was increased by 48% after light illumination.No obvious change in activity within 25 h of on-stream toluene oxidation at 185 °C.[44]
Formic acid oxidationPt1/ATOThe Pt1/ATO catalyst showed activity and durability superior to those of the commercial Pt/C catalyst.The structure was retained after a test of 1800 cycles.[45]
Crotyl alcohol, cinnamyl alcohol, benzyl alcohol oxidationPd/meso-Al2O3TOF = 7080 h−1 at 60 °C in crotyl alcohol oxidation; the most active catalyst reported at present for cinnamyl alcohol oxidation (TOF = 4400 h−1); TOF = 4096 h−1 and more than 99% selectivity of benzaldehyde for benzyl alcohol oxidation.[46]
Benzene oxidationPt1/meso-Fe2O3The TOFPt (2.69 s−1) over Pt1/meso-Fe2O3 was much higher than that (1.16 s−1) over 0.25 Pt NP/meso-Fe2O3 at 160 °C.No significant change in activity after 40 h of on-stream reaction.[47]
Methanol oxidationPt1-Co3O4The reaction rate at 70 °C (0.013 mol/(gPt s)) over Pt1-Co3O4 was over four times higher than that (0.03 mol/(gPt s)) over Pt/Co3O4.After water-resistant tests, notable losses were observed; the Pt1-Co3O4 catalyst possessed a better CO2 (5 vol%) resistance.[54]
Toluene oxidationPd1/TiO2Complete toluene removal was achieved at 190 oC (Pd/TiO2 required a much higher temperature at 250 °C).[69]
Oxidation of benzene to phenolFeN4/GNThe conversion was 23.4% and the yield was 18.7% at room temperature.Benzene oxidation took place efficiently at 0 °C, with a phenol yield of 8.3% after 24 h of reaction.[80]
Methanol oxidationPt1-CeO2{100}The TOFPt was 0.68 s−1 at 50 °C (T50% = 33.7 °C, T90% = 51.9 °C).Negligible deactivation can be observed for an extended 800 min at 52 °C in a mixture airflow.[83]
Toluene oxidationPt/MnO2100% conversions of 0.42 ppm and 10 ppm toluene were achieved at 28 and 80 °C, respectively. The complete oxidation of toluene into CO2 took place at 220 °C.Kept stable at 220 °C within 10 h of reaction.[84]
HCHO oxidationAg-HMOThe TOF at 80 °C (0.035 s−1) over Ag-HMO was about seven times higher than that (ca. 0.005 s−1) over Ag/HMO.[85]
Alkali-Metal- Pt/TiO2A 100% conversion of HCHO was achieved at 40 °C over the 1% Na-Pt/TiO2 catalyst.[86]
Pt/Mn-TiO2The complete oxidation of 100 ppm HCHO was achieved at room temperature under the conditions of WHSV = 60,000 mL/(g h) and RH = 50%.During the long-term activity test (>75 h), Pt/Mn-TiO2 displayed good stability, even at a WHSV of 120,000 mL/(g h).[87]
CO oxidation0.22Pt1/CeO2100% CO conversion was obtained at 160 °C, and the temperature was remarkably decreased to 98 °C in the presence of water.The activity was virtually preserved at 92 °C after 96 h.[34]
0.25Au/TiO2The full conversion temperature was 120 °C, which was significantly lower than that over Au/Per-TiO2.Without an obvious decrease in catalytic activity after 600 min of reaction.[39]
CeO2-AA-PtThe temperature of the CO complete conversion was 60 °C.No activity losses in five cycle tests.[42]
CuCeA CO consumption rate of 6100 μmolCO/(gCu s) was achieved at 120 °C, which was much higher than that over the other reported copper catalysts.Possessed remarkable cyclic stability at 600 and 200 °C (5 mg of catalyst, GHSV = 240,000 mL/(g h)).[49]
Pt-HT-SAsKept the excellent stability in CO or methane conversion at 220 °C for 50 h.[78]
Pt/m-Al2O3TOF was 0.023 s−1 at 200 °C and steadily increased to 0.175 s−1 at 250 °C (GHSV = 4.8 × 107 mL/(g h)).Kept the stable activity after 60 cycles within a month.[88]
Au1/Co3O4TOF was 3.1 s−1 at −75 °C, which was one of the most active Au catalysts for CO oxidation (T90% = −100 °C).Kept stable after 10 cycle tests.[89]
CoOx/Pt/CeO2TOF was 0.88 s−1 at 25 °C, with a GHSV of 75,000 mL/(g h) (T50% = 6 oC, T90% = 20 °C).Kept stable during 36 h of on-stream reaction.[90]
CH4 oxidationPt/Fe2O3The specific activity (2.01 molCH4/(gPt h)) over Pt/Fe2O3 was 4 times higher than that that (0.47 molCH4/(gPt h)) over Pt/Fe2O3-NP and 20 times higher than that over Pt/Al2O3.Maintained a similar activity for over 16 h of on-stream reaction.[55]
Pd1/TiO2The temperature (380 °C) for the complete oxidation of methane was substantially decreased.[71]
1Pd/2Pt@CeO2At 300 °C, the reaction rate (3.27 mmolCH4/(molPd s)) was three times higher than that (0.91 mmolCH4/(molPd s)) of (1Pd + 2Pt)/CeO2.A loss of 3.8% in reactivity in the presence of 4 v/v% water vapor and a loss of 15.7% in reactivity in the presence of 10 v/v% water.[91]
PdW1/Al2O3The TOF at 260 °C over PdW1/Al2O3 was much higher than that over Pd/Al2O3.No deactivation during the 30 h reaction process.[92]
Cu1/ZSM-5The Cu1/ZSM-5 catalyst was highly active (C1 oxygenates productivity of 4800 μmol/gcat at 50 °C and 12,000 μmol/gcat at 70 °C within 30 min) and selective (C1 oxygenates selectivity of 99% at 50 °C).
C1 oxygenate productivity of 4800 and 12,000 μmol/gcat was achieved at 50 and 70 °C, respectively. A C1 oxygenate selectivity of 99% was achieved at 50 °C.
[82]
Au1/BPThe mass activity reached 113.5 μmol/gcat in the presence of water with 33 bar of mixed gas (CH4:O2 = 10:1) at 90 °C under light irradiation (1.2 W).After 10 in situ cycles (20 h), about 205.2 μmol of methanol was generated.[93]
NOx reductionPd/TiO2The presence of Pd led to a drastic increase (4–5 times) in NOx removal efficiency after 5 h of reaction.[94]
NbCuCe80% NOx conversion was achieved at 150 °C, and a high N2 selectivity was detected during the activity testing process.Under the conditions of 5 vol% H2O, 5 vol% CO2, and 200 ppm C3H6, NbCuCe still exhibited a better catalytic performance than Cu-CHA and NbCe.[95]
Mo1/Fe2O3TOF at 270 °C over Mo1/Fe2O3 was ca. 1.7 × 10−3 s−1, comparable to TOF at 277 °C (1.3 × 10−3 s−1) and TOF at 323 °C (2.4 × 10−3 s−1) over V2O5/TiO2.[96]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, X.; Deng, J.; Liu, Y.; Jing, L.; Hou, Z.; Pei, W.; Dai, H. Single-Atom Catalysts: Preparation and Applications in Environmental Catalysis. Catalysts 2022, 12, 1239. https://doi.org/10.3390/catal12101239

AMA Style

Yu X, Deng J, Liu Y, Jing L, Hou Z, Pei W, Dai H. Single-Atom Catalysts: Preparation and Applications in Environmental Catalysis. Catalysts. 2022; 12(10):1239. https://doi.org/10.3390/catal12101239

Chicago/Turabian Style

Yu, Xiaohui, Jiguang Deng, Yuxi Liu, Lin Jing, Zhiquan Hou, Wenbo Pei, and Hongxing Dai. 2022. "Single-Atom Catalysts: Preparation and Applications in Environmental Catalysis" Catalysts 12, no. 10: 1239. https://doi.org/10.3390/catal12101239

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop