Single–atom catalysts based on Fenton-like/peroxymonosulfate system for water purification: design and synthesis principle, performance regulation and catalytic mechanism

Liping Hao a, Chao Guo ab, Zhenyu Hu ab, Rui Guo *ab, Xuanwen Liu *ab, Chunming Liu a and Ye Tian c
aSchool of Materials Science and Engineering, Northeastern University, Shenyang 110819, China
bSchool of Resources and Materials, Northeastern University at Qinhuangdao, Qinhuangdao 066004, China
cThe First Hospital of Qinhuangdao 066099, China

Received 30th May 2022 , Accepted 9th August 2022

First published on 10th August 2022


Abstract

Novel single-atom catalysts (SACs) have become the frontier materials in the field of environmental remediation, especially wastewater purification because of their nearly 100% ultra-high atomic utilization and excellent properties. SACs can be used in Fenton-like catalytic reactions to activate various peroxides (such as hydrogen peroxide (H2O2), ozone (O3), and persulfate (PSs)) to release active radicals and non-radicals, acting on target pollutants, and realize their decomposition and mineralization. Among them, peroxymonosulfate (PMS) in PS systems has gradually become an important oxidant in Fenton-like processes due to its asymmetric molecular structure and characteristics of easy storage and transportation. Focusing on the numerous proposed strategies for the synthesis and performance regulation of Fenton-like SACs, it has been confirmed that the coordination of isolated metal atoms and the support/carrier enhances the structural robustness and chemical stability of these catalysts and optimizes their catalytic activity and kinetics. Moreover, the tunability of the coordination environment and electronic properties of SACs can improve their other catalytic properties, such as cycle stability and selectivity. Thus, to systematically explain the relationship between the active center, catalyst performance and the corresponding potential catalytic mechanism, herein, we focus on the representative scientific work on the preparation strategy, catalytic application and performance regulation of Fenton-like SACs. Specifically, we review the typical Fenton-like SAC reaction processes and catalytic mechanisms for the degradation of refractory organic compounds in advanced oxidation processes (AOPs). Finally, the future development and challenges of Fenton-like SACs are presented.


1. Introduction

The seminal study of creat strong oxidants using Fe(II) and H2O2 to degrade tartaric acid by Fenton H.J.H in 1894 opened the door of novel oxidation processes.1 Subsequently, the Fenton reaction has become one of the few organic chemical reactions named after a person. However, it was not until 1964 that Eisenhaner, a Canadian scholar, first applied the Fenton reagent in water treatment. The homogeneous catalyst involved in the reaction exhibited the advantages of high efficiency, low cost and wide applicability.2 However, the insignificant separation and recovery of iron species, the narrow optimal pH range (pH ∼ 3) and the production of ferric hydroxide sludge, which is regarded as secondary pollution, seriously limited its practical application value.3–5 Accordingly, Fenton-like technology using heterogeneous catalysts such as metal oxides and metal sulfides is a successful alternative strategy to overcome the main defects of the traditional Fenton reaction.6 However, due to the limited utilization of atoms that can contact and participate in the reaction on the surface of these catalysts, their catalytic efficiency is not satisfactory.7–9 Therefore, reducing the particle size of the catalyst plays an indispensable role in improving the ability of its surface atomic sites to participate in the reaction and coordinating the reaction activity and homogeneous/heterogeneous reaction kinetics.

In the last decade, isolated monatomic metal sites formed by downsizing nanoparticles to the atomic level and anchoring them on support materials have been defined as single-atom catalysts (SACs) (Fig. 1).10 The term SACs was first proposed in 2011, regarding single Pt atom anchored on FeOx as the active site for CO oxidation, its catalytic effect was 3-times higher than that of the corresponding Pt bulk.11 Since then, the oxidation or reduction process with single atoms as the reaction center has become a research hotspot of catalysis. Importantly, SACs have linked heterogeneous and homogeneous catalysts, successfully bridging the gap between them.12 Moreover, owing to the mononuclear property of their reaction sites, SACs show the highest atomic utilization efficiency and the strong covalent bonds between the isolated metal atoms and anchored substrate endow SACs with prominent structural robustness and chemical stability compared to their nanoparticle counterparts. To date, multiple types of SACs based on noble/non-noble metals (e.g., Pt, Ag, Au, Fe, Cu, Co, Mn and Cr) dispersed and fixed on various metal-free organic/semiconductor substrates (carbon supports, g-C3N4, TiO2, FeOx, Ti2AlN, Ti3AlC2, VO2, MoS2, etc.) have been reported.13–15 These SACs with a unique coordination environment and electronic characteristics have been applied in various catalytic reactions, such as the hydrogen evolution reaction (HER), and oxygen reduction reaction (ORR), CO oxidation, hydrogenation reactions, and CO2 reduction reaction (CRR).16–20 However, the bottleneck limiting the applicability and adaptability of SACs in practical is related to the accurate regulation of their atomic configuration and the large-scale batch synthesis of SACs. This is mainly because metal atoms with high surface free energy tend to spontaneously migrate and gather into clusters or even particles.21 Thus, the isolation and anchoring of single metal atoms in precursors is of vital importance in the design and synthesis of SACs.


image file: d2nr02989h-f1.tif
Fig. 1 Number of publications of SACs has increased sharply in the past decade. The illustration vividly illustrates the possible anchoring strategy of isolated metal atoms in SACs.

Advanced oxidation processes (AOPs) using SACs, especially Fenton-like catalytic systems are still in their infancy.22 Also, to date, most of the designed and synthesized Fenton-like SACs are still limited to carbon-based SACs and g-C3N4-based SACs, with relatively few studies on semiconductor-based SACs. Fenton-like catalysts can activate the oxide species (e.g., hydrogen peroxide (H2O2), ozone (O3), and persulfate (PSs)) adsorbed on their surface to release active free radicals or non-radicals to act on target pollutants and degrade them into small molecular compounds or CO2 and H2O.23–25 Specifically, Fenton-like SACs/PS oxidation systems are powerful catalytic systems for the decomposition and mineralization of toxic organic macromolecules because they generate reactive oxygen species (ROS), such as hydroxyl radical (HO˙) and sulfate (SO4˙).26–29 Benefitting from the rapid development of atomic-resolution characterization technologies, it was possible to observe isolated atoms directly. Moreover, the free radical capture experiment and theoretical density functional theory (DFT) calculation also provide a vital contribution to the verification and elucidation of their catalytic mechanism. Thus, Fenton-like SACs have broken through the catalytic activity and kinetic limitations of traditional catalysts and become the most potential candidate catalysts in the field of environmental remediation. However, the reports on the application of Fenton-like SACs in the field of environmental remediation are still rare.

In this report, we aim to elaborate the concept of SACs, the design and synthesis principles, catalytic performance regulation and mechanism of Fenton-like SACs. The main method for the preparation of Fenton-like SACs using a pure carbon substrate, heteroatom-doped carbon matrix and semiconductor support to disperse and anchor a variety of isolated metal atoms is mainly discussed. The designed and synthesized catalysts have a unique coordination environment, geometric configuration and electronic characteristics, which significantly affect the reaction activity, kinetics and cycle stability of the catalytic system. This is due to the following reasons: (i) each metal atom is endowed with unique electronic properties (oxidation state or spin state), and thus a change in the metal atoms coordinated on the surface of the substrate will result in significant differences in the catalytic effect and the decomposition path of the target substance; (ii) the synthesized Fenton-like SACs also perfectly inherit the characteristics of the carefully selected support materials, such as the optimized electron transfer characteristics of carbon-based SACs, the photocatalytic or electrocatalytic properties given by semiconductor-based SACs, and the porous structure retained by MOF-based, improving their specific surface area to enhance mass transfer; (iii) the coordination environment (including heteroatoms such as O, S, Se and N) of a single metal atom reaction center can also adjust the charge characteristics and reaction activity of metal sites, and may also increase the loading content of metal atoms. Finally, the catalytic mechanism of free radical and non-radical approaches applied to sewage treatment is explained from the perspective of experimental verification and theoretical calculation in Fenton-like SAC/PMS systems for the decomposition of target pollutants.

2. Concept

One of the original purposes of introducing SACs was to achieve a breakthrough improvement in atomic utilization after downsizing nanomaterials, that is, the surface/interface of nano-catalysts involved in the reaction are endowed with a larger effective surface area and higher density of reaction centers.30–32 Based on the definition by Gate and Thomas, with a decrease in the metal size, entities with a diameter of 2 to 8 nm are used to describe nanoparticles, whereas molecular aggregates between 2 to 30 atoms are considered clusters. According to the definition of SACs, the active metal in the form of a single atom is anchored and dispersed on the surface of a carrier through covalent or ionic interaction with adjacent atoms, and there is no interaction between each isolated metal atom.33–35 However, the coordination environment of isolated metal species may not be completely consistent. When they are completely consistent, single-atom catalysis is also known as unit point catalysis. It is worth noting that the active center of SACs does not mean that the existing state of a single metal species is zero valence. On the contrary, due to the coordination effects such as electron transfer between the isolated metal species and other species on the carrier surface, they often show a certain charge. The concept of “single-site” and “isolated metal” catalyst indicate that each reaction center is characterized by a metal species. In some catalytic systems, the synergy between the isolated metal atoms and adjacent anchored atoms is an indispensable way to regulate their catalytic activity. To date, the number of articles related to SACs has increased annually. Among them, a variety of transition metal atoms and noble metal atoms has been successfully developed in the research on Fenton-like SACs (Fig. 2). Also, they have been proven to be popular and potential catalysts in the field of environmental remediation, especially wastewater purification.
image file: d2nr02989h-f2.tif
Fig. 2 Publications on single-atom heterogeneous catalysts for all elements in the transition group in the periodic table since March 2017.

3. Single-atom catalyst design principle

In recent years, the methods for the synthesis of Fenton-like SACs have developed unprecedentedly, and a variety of physical and chemical preparation methods has been reported, such as direct pyrolysis, hydrothermal/solvothermal treatment, chemical vapor deposition and high-energy ball milling.36–39 These synthetic routes can be roughly divided into “bottom-up” and “top-down” approached using mononuclear metal complexes and bulk or nanoparticles metals as the initial precursors, respectively. The conditions in the “bottom-up” methods need to be accurately controlled, and the metal loading of the prepared catalyst is usually substantially lower than that of the starting precursor because it is difficult to disperse and stably form a single metal atom in the key post-treatment process of removing mononuclear metal complexes.40,41 The “top-down” route usually accelerates the thermal movement of metal atoms under sufficient energy (high temperature), reaches the fracture of metal–metal bonds existing in metal nanoparticles or bulk metals, and makes a strong interaction between the escaped isolated metal species and the anchor point for the preparation of stable and completely dispersed SACs.42–45

Based on the above discussion, the classification of the design and synthesis strategies for Fenton-like catalysts herein is based on the difference in their active center and support/carrier, which is preliminarily divided into three types, such as C-coordinated metal (M–C), heteroatom (e.g., N, P, O, and S)-introduced carbon substrate-fixated metal, and semiconductor-dispersed metal atom. The preparation process of Fenton-like catalysts with various active centers in the above-mentioned classification and the selection of precursor raw materials are explained in detail in the subsequent sections. These Fenton-like SACs synthesized by different design routes and methods show different geometric structures, coordination environments and electronic characteristics, which will result in significant differences in oxidation activity and kinetics in heterogeneous reactions, especially the Fenton-like systems highlighted in this review.

3.1 C-coordinated metal

At present, the “bottom-up” strategy is widely used in the preparation of Fenton-like SACs, where the single metal species existing in metal precursors are captured, reduced and anchored by coordinated in a defect-rich support. As is known, the isolated metal atoms are usually thermodynamically unstable and tend to move spontaneously to form clusters or even nanoparticles to improve their stability. Accordingly, the synthesis strategy of carbon material extraction is helpful to prevent the agglomeration of single metal elements caused by high surface energy. So far, a variety of carbon materials has been developed to promote the capture and fixation of noble or non-noble metal elements. For example, carbon materials rich in oxygen and nitrogen functional groups, such as carboxyl, anhydride, phenolic hydroxyl, carbonyl and other oxygen-containing functional groups, and nitrogen-containing functional groups, such as amide, imide, lactam, pyrrole and pyrimidine, have been widely developed and used in the preparation of Fenton-like SACs.46–49 In addition to the capture of metal atoms by oxygen functional groups, the unsaturated coordination edges of graphene (such as serrated edges, armchair row edges, and single carbon vacancies or double vacancies) also play a significant role in the dispersion and anchoring of metal atoms.50,51 The preparation method of carbon extraction and coordination of metal atoms is simple, easy to implement and has the advantage of economic cost. However, the only disadvantage to prevent its mass production is the unsatisfactory catalytic efficiency caused by the low metal atom loading.

3.2 Heteroatom-doped carbon material-fixed metal

The design and synthesis strategy of anchoring metal atoms and dispersing them in a heteroatom-doped carbon matrix by electrostatic interaction can realize a satisfactory loading content and stable chemical environment in atom-dispersed metal Fenton-like catalysts. This is mainly due to the greater number of unsaturated lone pair electrons and structural defects provided by the introduction of heteroatoms, which make the capture of metal precursors more efficient and a significant increase in metal atom loading possible.52 Among the variety of heteroatom dopants, the nitrogen atom has become the most widely studied and mature electronegative non-metallic atom for extracting and fixing metal atoms in a carbon matrix. Moreover, the simplest and most efficient methods for preparing Fenton-like SACs based on nitrogen-doped carbon substrates coordinated with isolated metal atoms include the one-step sintering method, impregnation and copolymerization synthesis strategy, and hydrothermal or solvothermal method.53–55 Importantly, various rich-nitrogen carbon precursor materials (e.g., organic small molecule compounds, metal–organic frameworks (MOFs), polymers, and biomass) have also been successfully developed and applied to prepare Fenton-like SACs for environmental remediation, especially the water purification system highlighted herein (Fig. 3).
image file: d2nr02989h-f3.tif
Fig. 3 Synthesis routes of various rich-nitrogen carbon precursor materials heteroatom-doped carbon materials-fixed metal atom catalysts: (a) organic small molecule compound, (b) metal-organic frameworks (MOFs), (c) biomass and (d) characterization methods.
3.2.1 Organic small molecule compounds. Metal precursors are mainly dispersed on carefully selected nitrogen-doped carbon precursors in the form of physical adsorption/encapsulation or chemical coordination, and monoatomic Fenton-like SACs corresponding to the precursor can be successfully prepared through the subsequent synthesis process. At present, a variety of nitrogen-rich carbon organic small molecules that can act as nitrogen-doped carbon precursors has been developed and utilized, such as urea, dicyandiamide, thiourea, cyanamide, melamine and 1,10-phenanthroline.56–61 Interestingly, the organic-conjugated graphite carbon nitride (g-C3N4) matrix with rich covalent and hydrogen bonds obtained via the one-step pyrolysis of the above-mentioned compounds can realize the coordination and fixation of isolated metal atoms, and this design and synthesis strategy has the advantages of universality and reproducibility. Particularly, g-C3N4 is often used as a green carrier and energy storage material in the traditional catalytic field because of its excellent chemical inertness, high specific surface area and a variety of nano multi-level structures.62 Therefore, this metal–nitrogen coordination SACs also inhibits the leakage of metal ions and reduces the potential risk to the environment, which is ascribed to its coordination field effect, geometric configuration and chemical stability. It also opens up a new way for Fenton-like SACs to develop in the direction of wide pH range and sustainable catalysis. Thus, the g-C3N4-confined metal Fenton-like SACs synthesized from nitrogen-rich small molecule organic compounds have become a potential candidate to replace the traditional transition metal oxides.

Recently, Wang and colleagues demonstrated a sandwich-structure catalyst strategy for the synthesis of iron atoms dispersed and fixed on a C3N4-rGO bilayer support, and this unique double-layered structure has the characteristics of preventing metal atom agglomeration and leakage on the premise of ensuring the activity of reaction sites (Fig. 4a).63 According to the DFT calculation of the optimized material structure, it can be proven that the novel structure of limiting isolated metal atoms via the bond interaction provided by the support layers on both sides has stronger stability than the traditional binary C3N4–Fe composite. Moreover, this metal atom-linked sandwich structure can make the excited electrons and holes migrate to the C3N4 side and rGO layer, respectively (Fig. 4b). In contrast, although the process for the synthesis of SAC based on graphite carbon nitride rich in pyrrole nitrogen (PN-g-C3N4) developed by Chen and co-workers was not as simple as that by Wang, the advantage of this preparation process was to realize the large-scale production of Fenton-like catalysts and the replicable synthesis of various metal atoms.64 Specifically, the electrostatic self-assembly MCAXT polymer acted as the support precursor and synthesized via the hydrogen bonding of melamine (M), cyanuric acid (CA) and xanthine (XT) in a homogeneous aqueous phase, which contained numerous imidazole groups and acted as the coordination-fixed unit of TM. Finally, the TM-loaded MCAXT polymer was calcined under an N2 atmosphere to successfully obtain SA-TM/PN-g-C3N4 (Fig. 4c). The successful preparation of the monatomic SA-Cr/PN-g-C3N4 catalyst was verified by the bright-spot image due to the high Z contrast in the HAADF-STEM characterization, as highlighted in the partially enlarged view in Fig. 4f. The pyrrolic N in the N 1s XPS spectra (Fig. 4l) indicates that the imidazole group was successfully doped in the tri-s-triazine ring of g-C3N4. It was confirmed that the introduced pyrrole group provided rich binding sites for the adsorption and fixation of TM. The atomic configuration and coordination environment of the reaction unit in monatomic Fenton-like catalysts are critical for the regulation of their catalytic efficiency. The X-ray absorption near-edge spectroscopy (XANES) study showed that the valence state of the Cr species in SA–Cr/PN-g-C3N4 was +3. The extended X-ray absorption fine structure (EXAFS) curve (Fig. 4n) revealed the existence of a Cr–N bond anchoring mode, and the coordination environment of the Cr species was similar to that of Cr(III) tetraphenylporphyrin chloride (CrTPPCl) due to their almost consistent κ2χ(κ) oscillations. The fitting curve (Fig. 4p) intuitively proves the coordination of Cr–N4 with an average distance of 1.91 Å.


image file: d2nr02989h-f4.tif
Fig. 4 Preparation of C3N4–Fe–rGO composites and DFT calculations: (a) schematic of the design and key synthesis steps of C3N4–Fe–rGO catalysts. (b) Chemical configuration of C3N4–Fe–rGO composite. (c) Projected density of states (PDOS) and local electronic structure of C3N4-Fe-rGO.63 Copyright 2020, Elsevier B.V. All rights reserved. Preparation and characterization of SA-TM/PN-g-C3N4, (d) synthetic route of SA-TM/PN-g-C3N4, (e) HAADF-STEM, and (f) corresponding enlarged picture. (g) HRTEM image, (h–j) EDX mapping, (k) XRD patterns, (l) high-resolution N 1s XPS, and (m) Cr 2p XPS of SA-Cr/gPN-C3N4. (n) Cr K-edge XANES. (o) Fourier-transform EXAFS and (p) corresponding EXAFS image.64 Copyright 2021, Wiley-VCH GmbH.

In addition to the above-mentioned pyrrolic N, the common forms of N species in Fenton-like SACs include oxidized N, graphitic N, and pyridinic N. The effects of different forms of N species on the chemical structure and catalytic reaction kinetics of synergistic catalysts are briefly described in this paragraph. Among them, the pyrrole group is expected to be a satisfactory anchoring unit for the extraction, fixation and coordination of transition metal elements. Pyrrolic N is usually considered as an adsorption site, which can optimize the adsorption capacity between N-doped carbons and organic pollutants.65 Moreover, pyridinic N has been proven to be the active center of the reaction in some works. Its existence adjusts the local charge distribution, accelerates the electron transfer, and helps to reduce the adsorption energy barrier of the adjacent carbon atoms.66 Graphitic N with an electron-deficient feature can obviously optimize the conductivity and selectivity of the catalyst, which can be modulated by electron-rich pyrrolic N/pyridinic N and graphitization structures.67 Because the electrons in pyrrolic and pyridinic N tend to be delocalized through the π-conjugate graphitization plane. Therefore, the controllable synthesis of N species in Fenton-like SACs is significant for regulating their performance and stability.

As discussed above, Fenton-like catalysts based on various metal atoms have been successfully synthesized, which are dispersed and anchored on nitrogen-doped carbon substrates via the generalized “extraction and coordination” method. Also, various dispersed noble (Ag, Pt, and Ir) and non-noble (Fe, Co, Cr, Mn, and Cu) metals have been anchored on N-doped C substrates to form metal-coordinated N center (M–N–C or M–Nx) materials. This M–N–C or M–Nx active center effectively prevents the leakage of metal atoms under extreme conditions, and experiments have proven that catalysts with this type of active site have higher catalytic activity, durability and atomic utilization. Therefore, the process for the preparation of Fenton-like SACs by the derivatization of small molecular organic compounds, which mainly focuses on the chemical structure and electronic configuration of reaction sites, is a simple, feasible and low-cost method and expected to realize their large-scale mass production.

3.2.2 MOFs. Metal organic frameworks (MOFs) are organic–inorganic hybrid materials with intramolecular pores formed by the self-assembly of organic ligands and metal ions or clusters through coordination bonds.68,69 Besides their large surface area, MOFs have excellent properties, including rich nano-cavities and open channels, which not only enhance the mass transfer in the catalytic process, but also play a role in the separation and anchoring of metal atoms, that is, micropore restriction strategy.70–72 Taking advantage of this intriguing feature, MOFs have recently been certified as superior candidate templates for heterogeneous Fenton-like SACs. At present, the MOF precursors for the production of Fenton-like catalysts, which have been applied to nitrogen-doped carbon precursors to disperse single metal atoms, mainly include Prussian blue analogue (PBA, AxM1[M2(CN)6]y·zH2O),73 zeolite imidazole frameworks (ZIF-8 and ZIF-67),74 and Fe-based frameworks (MIL-101 and MIL-88).75–78

Chen et al. used an Fe-doped Bi-MOF as a precursor to design and synthesize a FeBi-NC Fenton-like catalyst with double active centers, which had an ultra-high monatomic loading (FeBi-MOF, 2.61 wt% Fe and 8.01 wt% Bi).77 The highlight of its design was that the introduced Fe not only acted as the active site, but also expanded the distance between adjacent Bi atoms in FeBi-MOF like a “fence”, thus maximizing the dispersion of metals in the pyrolysis process (Fig. 5a). Polyvinylpyrrolidone (PVP) can flake the bulk MOF precursor, making it easier for metal atoms to evaporate and atomize to form isolated monatomic sites during carbonization. Tang et al. modified the metal nodes of the inherent MOF to optimize its catalytic activity because most of the metal nodes were saturated with organic ligands, which could prevent the metal sites from participating in catalytic reaction.78 In their approach, the thermal treatment under an inert atmosphere or vacuum condition caused the partially non-bridged terminal organic ligands (e.g., H2O and halogen/hydroxide anion ligands) to leave the metal center to construct coordination unsaturated metal sites (CUSs). This could optimize the adsorption capacity of the MOF and caused the oxidant to have better affinity and regioselectivity with the active site. Moreover, H2O2 as a Lewis base was more conducive to bind to CUSs as a Lewis acid site and accelerate its catalytic splitting to produce rich hydroxyl radical (˙OH). Specifically, CUS-MIL-100(Fe) was prepared via hydrothermal treatment (150 °C, 8 h) with iron salt and 1,3,5-benzoic acid (H3BTC) as raw materials and modified under high temperature (230 °C) and vacuum conditions. The reaction activation pathway for CUS-MIL-100 (Fe) to activate H2O2 and degrade SMT was further clarified, as illustrated in Fig. 5e. Its large specific surface area and channel (∼6.57 nm pore size) fully exposed the unsaturated metal active sites and favored the enrichment and diffusion of oxidants and organic pollutants in its internal materials. Therefore, the activity and stability of the Fenton-like catalytic reaction were significantly optimized. The coordinated and supported M–Nx complexes Fenton-like catalysts in the form of MOFs not only stabilize the metal species and reduce the leaching of metal ions, but also accelerate the metal ion redox cycle. Therefore, the pH range in which they can work effectively is optimized. Nevertheless, they have a fatal disadvantage, i.e., poor stability due to the self-oxidation and aggregation of their organic ligands, thus greatly limiting their service lifetime and practical application.


image file: d2nr02989h-f5.tif
Fig. 5 Synthesis and theoretical calculation of FeBi-NC: (a) schematic diagram of the preparation process, (b) theoretical study on the mechanism of FeBi-NC enzyme oxidase: optimized structure of various intermediates produced in FeBi-NC enzyme oxidase reaction and (c) free energy diagram of the reaction on M–N4 active site with TMB as the reducing agent.77 ©2021, Elsevier B.V. All rights reserved. (d) and (e) Morphological characterization of CUS-MIL-100(Fe) and schematic diagram of the reaction mechanism for the activation of H2O2.78 ©2018, the American Chemical Society.

In summary, the ordered pores/channels and stable chemical structure produced by electrostatic self-assembly between metal atoms and organic ligands are excellent in nitrogen-doped carbon precursors for the preparation of Fenton-like catalysts. Also, the synthesis process was simple, where they are usually produced by direct pyrolysis under the conditions of high temperature and inert gas protection. However, one of the unavoidable problems in the pyrolytic synthesis strategy of metal atoms based on MOFs is that metal elements were relatively easy to move and form clusters or even nanoparticles. Consequently, the solution to this problem its usually further acid treatment to obtain clean and perfectly anchored single metal atoms on the MOF-derived NC matrix.

3.2.3 Biomass derived. The above-mentioned organic small molecules and organic ligands in MOFs as nitrogen-doped carbon precursor substrates usually have certain toxicity and will bring potential risks to the environment. Therefore, it is urgent to develop alternative environment-friendly anchoring matrices. Biomass provides a promising solution for the amplification methods to synthesize gram-quantities of isolated metal atom center Fenton-like catalysts to enable their application in oxidation reactions and adaptability for practical use. For example, Zhu's group used the inherent abundant biological functional groups of chitosan and the unique hypersaline conditions of ZnCl2 and CoCl2 salts (mass ratio of CoCl2·6H2O to chitosan was 1[thin space (1/6-em)]:[thin space (1/6-em)]1) to successfully synthesize a single-center catalyst supported on N-doped two-dimensional nanobelts with an ultra-high specific surface area (Co-ISA/CNB, 2513 m2 g−1).79 Specifically, in the low-temperature pyrolysis stage, the monomer chitosan was thermally dehydrated and crosslinked on the continuous 2D active salt template formed by the melting of inorganic crystalline salt, and the in-plane carbon was reconstructed to form a large-area nanobelt layer. At high temperature, carbothermal reduction reduced the inorganic salts to metal crystals, thus promoting the graphitization of the carbon skeleton. At high temperature, the active salt was reduced into large metal crystals by carbothermal reduction, thereby accelerating the graphitization of the carbon-containing frameworks. ZnCl2 salt can be used as a dehydrating agent, foaming agent and pore forming agent in the process of polymerization and carbonization to form porous nanostructures with rich micropores and mesopores, significantly improving the specific surface area. CoCl2 salt was added to regulate the formation of a graphitized carbon skeleton and disperse the loaded metal species.

Lignin, the second most abundant cross-linked phenolic polymer, is a significant “solid waste” in the pulp and paper industry.80 However, because its abundant functional groups can coordinate with transition metal ions to form insoluble supramolecular metal-lignin complexes and promote the separation and anchoring of metal atoms, it has become the most promising candidate among natural biomass materials to replace organic ligands. Accordingly, Qi developed a one-pot pyrolysis strategy with lignin as an anchor rather than organic ligand to prepare a Co single-atom catalyst (SA Co–N/C catalyst).81 The Co content determined by ICP-MS was up to 2.45 wt% and it was proven that the Co content was positively correlated with the catalytic rate constant (R2 = 0.9675) (Fig. 6a). In addition, the study of catalytic performance showed that the adsorption of PMS (about 65%) mainly occurs on the N/C carrier, while the isolated atom Co (2–16 mM) was the active center for the activation of PMS, accelerating the consumption of PMS and effectively degrading pollutants (Fig. 6b). Also, the Co–N3(II) active center significantly improved the utilization of PMS, where due to the high Co loading (40 mM), the consumption of PMS gradually decreased.


image file: d2nr02989h-f6.tif
Fig. 6 Characterization of catalytic performance of SA Co–N/C materials: (a) linear correlation between Co content and rate constant in the SA Co–N/C Fenton-like SACs. (b) PMS concentration changes in different Fenton-like SACs/PMS systems. (c) Selectivity of different SACs. (d) Adsorption coefficients (kobs) of different pollutants in SA Co–N/C/PMS, NPs Co/C/PMS and Co2+/PMS systems. (e) Co leaching rate.81 ©2021, Elsevier B.V. All rights reserved.

Moreover, the performance of SA Co–N/C Fenton-like catalysts for the selective degradation of pollutants has also attracted attention, as shown in Fig. 6c. The degradation rates of contaminants (e.g., NPX, PCM, CIP and BPA) in SA Co–N/C/PMS systems were significantly higher than that of CP and MNZ containing electron-withdrawing groups such as carboxyl and nitro groups (Fig. 6d). Therefore, SA Co–N/C Fenton-like catalysts have high selectivity for refractory contaminants through PMS activation. Basically, the Co–N3 active sites dispersed and anchored in the carbon supports are suitable for operation in a wide pH range and minimize the Co leaching to obtain excellent stability and high catalytic capacity, which have potential application prospects (Fig. 6e).

Although it was easy to replace small organic molecules with natural biomass to coordinate and fix metal atoms, its catalytic activity is usually limited. Therefore, doping heteroatoms (such as N, S and P) in the carbon support can improve the electron density and spin density of adjacent carbon atoms, which is an efficient means to enhance its Fenton-like activity. Du et al. developed a simple and easy path to realize the co-doping of electronegative elements such as nitrogen and sulfur in biomass materials.82 Specifically, the strong interaction between biomass-derived bio-sorbent reed containing high quaternary ammonium salt groups and SO42− makes the adsorbed sulfate ions become potential sulfur dopants, which coordinate with nitrogen and anchored Co during pyrolysis. Reed bio-sorbent derivatives saturated with nitrate, sulfate or phosphate were successfully employed to prepare Fenton-like catalysts by N, S or P co-doping (Co–S(P)@NC). As can be seen in Fig. 7c, the content of introduced single atom Co in the Co–S@NC Fenton-like catalysts was positively correlated with the content of N. Co combined with the N atoms in reed derivatives with adsorbed saturated nitrate to form a functional rigid structure and induced the formation of Co–N sites during pyrolysis. It is worth noting that the highly linear correlation (R2 = 0.938) between pyridine N content (at%) and Co atoms content (at%) proved that biomass with a high N content corresponds to the presence of high pyridine N in the graphite layer, which is conducive to the anchoring of metal species during pyrolysis. The cyclic test in a tap water matrix also proved that the catalyst has the potential of long-term utilization in practical water treatment. In addition to the above-mentioned biomass-derived nitrogen-doped carbon carrier, other biomass such as Enteromorpha, Auricularia auricula, rubber powder and spirulina can be employed for the green and sustainable synthesis of Fenton-like SACs.83–85 However, the performance of biomass-derived catalysts is usually limited because the content of metal elements contained in biomass itself is not usually high.


image file: d2nr02989h-f7.tif
Fig. 7 Schematic of the preparation and performance characterization of monatomic cobalt derived from a biosorbent: (a) N species content (at%) in various Co@NC catalysts. (b) N 1s XPS of Co@NC and NC. (c) Relationship between pyridine N content and Co content in different Co@NC catalysts. (d) Design and synthetic route for Co–S@NC. (e) Stability test of Co–S@NC catalysts.82 ©2019, Elsevier B.V. All rights reserved.

3.3 M-semiconductor

3.3.1 Vacancy defect anchoring strategy. The defect limitation strategy for single-atom anchoring often appears on eco-friendly MAX ceramics. MAX ceramics (including Ti2AlN and Ti3AlC2) are novel and concerned machinable materials86 with the basic chemical formula of M(n+1)AXn, where M, A and X represent the transition metal element, main group element (IIIA and IVA) and C or N atom, respectively.87 Etching the A atoms of MAX ceramics with HF or HCl/LiF solution will generate vacancy defect cavities to anchor isolated transition metal atoms and form two-dimensional (2D) layered MXene structures to expose more active sites. For example, in the synthesis of Co@Ti2−xN and Cu@Ti2−xN Fenton-like catalysts with 2D MXene structures, molten transition metal salts were used to complete the etching of the MAX ceramics and the anchoring of monatomic metals in one step.88 This method is simple and applicable for the production of isolated metal atom catalysts. However, the migration of highly active metal single atoms on the carrier cannot be prevented through this reaction pathway, and thus it is often accompanied by the formation of metal clusters or particles, which need subsequent acid etching treatment.
3.3.2 Cation-substitution strategy. The cation exchange reaction is a mature post-synthesis method for the synthesis of semiconductor-semiconductor nano-heterostructures. It allows the cations in the preformed nanocrystals to be selectively replaced by new target cations, while the anion skeleton is retained, thus the fine-tuning of the size, shape, spatial orientation, composition and crystal structure of nano-materials can be realized.89 Typically, this method is used to regulate the energy band arrangement of photocatalytic semiconductor heterojunctions to strictly control the physical and optical properties of binary or ternary heterojunctions. Through the geometric lattice matching effect, the solute atoms enter the solvent lattice, that is, a substitutional solid solution is formed, and the crystal structure of the solvent components is retained. Inspired by the cation exchange reaction and substitutional solid solution, isolated metal atoms are directly fixed to the support through the element coordination effect of the support, thus effectively avoiding the agglomeration of transition metal single atoms in the preparation and catalytic engineering and sintering into clusters or particles. An atom-dispersed metal atom alloy (Au/VO2) in which the guest Au atom replaces the host V atom in the 2D VO2 nanobelt carrier was synthesized by Lu's research group via the solvent-assisted spontaneous stripping method.90 This lattice-doped Au atom enhanced the interaction between the metal and the carrier, which not only realized the fixation of the substitute Au atom on the carrier due to the atomic radius approximation (Au (0.134 nm) and V (0.122 nm)), but also improved the electron transfer and the charge redistribution due to the difference in electronegativity. This cation substitution strategy was scalable and easy to operate to activate Fenton-like oxidation to realize the efficient decomposition and mineralization of high molecular weight toxic substances.
3.3.3 Chemical vapor transport. Chemical vapor deposition (CVD) is a process in which gaseous or vapor substances react in the gas phase or at the gas–solid interface to form solid sediments.91 This process is divided into three important stages, i.e., reaction gas diffusion to the substrate surface, reaction gas adsorption on the substrate surface, and chemical reaction on the substrate surface to form solid deposits, and separation of gas-phase by-products from the substrate surface.92 The most common chemical vapor deposition reactions include thermal decomposition reaction, chemical synthesis reaction and chemical transport reaction.93–95 CVD can be employed to obtain Fenton-like SACs with a uniform geometry, especially large-area two-dimensional films. However, the CVD strategy requires the use of complex instruments, it is difficult to control the synthetic conditions and results in low productivity. Therefore, its practical application is greatly limited.

4. Performance regulation strategy

Generally, the strong coordination and anchoring interactions between isolated transition metals in Fenton-like SACs and adjacent atoms in the support/carrier are covalent bonds and ionic bonds. Also, with the improvement of characterization methods in recent years, research has focused on the excellent performance and catalytic mechanism of each active metal site in the oxidation process. In addition to developing and improving the existing synthetic processes for the preparation of highly atomically dispersed monatomic active center Fenton-like catalysts to maximize the atomic utilization, another indispensable strategy for the performance regulation of catalytic oxidation systems lies in the optimization of the inherent characteristics of these catalysts. Therefore, an increasing number of studies tend to deeply develop Fenton-like SACs, which can accurately regulate the reaction center and coordination environment of various transition metal species, thus endowing them with good oxidation activity, stability and selectivity in the process of catalytic oxidation and meeting the practical application requirements in industrial Fenton-like reactions (Table 1).
Table 1 Summary of single metal-based Fenton-like SACs for water treatment
Single atom Catalyst Support Active center Anchor method Loading [wt%, ICP] Performance Ref.
Noble metal Au Au/VO2 2D VO2(B) nanobelt Au Electronic metal–support interactions 3.7 RhB: TOF = 21.42 min−1 90
Phenol: TOF = 16.19 min−1
BPA: TOF = 80.89 min−1
Au@Mn/MoS2 MoS2 Mn Deposition-reduction and immobilization 0.5 Au enhanced the H2O2 productivity by about 2.5 times from bare MoS2 125
Ag Ag/mpg-C3N4 g-C3N4 Ag (visible light capture) Copolymerization BPA: TOC removal = 80% (60 min) Visible light photo-Fenton-like catalyst 148
Transition metal Cr SA–Cr/PN-g-C3N4 Pyrrolic N-rich g-C3N4 Cr(II)–N4 sites Molecular engineering approach Photocatalysis and single-atom catalysis 64
Mn Mn-ISAs@CN N-doped porous carbon MnN4 Coordination bond self-assembly BPA: k = 1.138 min−1 122
Pyrrolic N site (adsorption site) TOF > 5.69 min−1
Nonselective degradation: most organic pollutants
Fe Fe SA/NPCs N-doped porous carbon Fe–N4 Complex-sequestered strategy 2.17 RhB: k = 19.657 min−1 23
Pyrrolic N (adsorption site) TOF = 39.31 min−1
Nonselective (degradation most organic pollutants)
d-g-C3N4–Fe g-C3N4 Fe(III) and oxygen vacancies (OVs) Cation-π interaction Selective degradation: hydroxyl-containing pollutants ≫ hydrophobic pollutants 24
Fe/NC-30 3D N-doped carbon nanosheets Fe–N4 In situ CVD method 0.61 Phenol: k = 2.139 min−1 25
TOC removal = 83.56% (120 min)
Enteromorpha-derived N-doped carbon matrixes FeIV[double bond, length as m-dash]O and FeV[double bond, length as m-dash]O sites Coordination effect (between single metal atom and pyridine N) 0.84 Selective degradation: the aromatics with electron-donating groups such as hydroxyl and amido groups ≫ electron-withdrawing groups such as carboxyl and nitro groups 83
Fe–N–C
FeSA–N–C N-doped carbon matrixes FeIV=O sites Chemically Fe-doped zeolitic imidazolate frame-works 0.93 BPA: k = 0.240 min−1 53
Fe–N@C Biomass rich in N and Fe Graphitic N (PMS adsorption site) In situ pyrolysis 0.60 Paracetamol (PCM): k = 0.1213 min−1 137
Fe@MoS2 MoS2 Mo Fe Chemical vapor transport growth 0.66 Atrazine: k = 1.30 min−1 >3 times than that of pristine MoS2 134
FeCo@NC N-doped porous carbon FeCoN6 Cage-encapsulated-precursor pyrolysis method Fe: 0.98 BPA: k = 6.062 min−1 84
Pyrrolic N (adsorbs organic molecules) Co: 0.79 TOF > 60.62 min−1
Fe-POM/R-VO-TiO2 TiO2 Fe Oxygen vacancy rearrangement Synergetic effect between photo-catalysis and Fenton-like reactions 85
Co *SACo-NGs N-doped carbon Co–π structures (Co2+–N–Cπ) Metal cation–π orbital interaction BPA: k = 0.60 min−1 120
SA Co–N/C N-doped carbon Co–N3(II) Atom-size metal particles capture strategy 2.45 Naproxen (NPX): TOF = 4.82 min−1 81
SA–Co CNP 2D carbon nanoplate CoN4 One-step pyrolysis 0.1 Acetaminophen (APAP): k = 0.51 min−1 65
FeCo-NC-2 Porous CoN4 Pyrolysis of Fe–Co BPA: k = 1.252 min−1 62
N-doped graphene Pyrrolic N (adsorbs organic molecules) Prussian blue analogue with acid washing TOF = 12.51 min−1
Co–S@NC Sulfate saturated bio-sorbent Co One-step pyrolysis Dinotefuran (DIN): k = 0.048 min−1 82
TOC removal = 34.50% (90 min)
Co@Ti2–xN Ti2AlN MAX ceramic Single-atom Co Ti vacancy defect anchoring strategy Excellent performance in actual river water 88
Ru CNRu N-doped carbon Ru–N2 Coordinating to pyridine nitrogen Diclofenac (DCF): k = 0.628 (PMS, 0.63 mM) 138
Cu Cu-SA/NGO N-doped graphene CuN4 moieties Coordination and complexation 5.8 APAP: k = 0.052 (H2O2) 121


4.1 Transition metal species

To date, studies on Fenton-like SACs based on various transition metal atoms, such as the widely studied Fe-, Co-, Mn-, and Ni-based heterogeneous catalysts, have proven that different anchored active metal atoms may lead to different Fenton-like oxidation activity and catalytic mechanisms. Various transition metal species coordinate with the same heteroatoms to form Fenton-like SACs with the same coordination environment (such as electronic structure and geometric configuration), which can more clearly explain the activity and mechanism of different transition metal active sites through experimental, characterization and simulation evidence.96–98 For example, a series of SA-TM/PN-g-C3N4 catalysts with transition metal atoms (e.g., Cr, Mn, Fe, Co, and Cu) anchored on a PN-g-C3N4 support was prepared via the wet chemical method and top-down pyrolysis process in an inert atmosphere.64 By maintaining the same coordination environment and geometric structure of the catalyst, the relationship between different coordination metal atoms and the oxidation performance of the catalyst was discussed. The deep insight into the catalytic mechanism was closely related to the development of model construction and DFT theoretical calculation. The DFT calculation of the active center and relative adsorption energy showed that the catalytic activity of the SA-TM/PN-g-C3N4 catalysts with the same coordination configuration followed the order of Cr > Fe > Cu ≈ Co > Mn, which was due to the difference between the dispersed and fixed monatomic metals.

Moreover, the intrinsic electronic properties (oxidation state or spin state) of transition metal species can be regulated at the atomic level in the metal atom-nitrogen coordination active center.100 Importantly, the spatial electron cloud density of the d orbitals of transition group elements is closely related to their intrinsic properties. Generally, the eg orbital of transition metals participates in σ-bonding with oxygen-containing species by overlapping with the O2p orbitals, thus affecting the bonding strength and electron transfer rate between the metal element center and oxidants.101 The study of the spin states of transition metals can be further conducted to understand the mechanism of metal element-regulated PMS activation and crucial for the design and synthesis of N-coordinated metal (M–N) Fenton-like SACs for environmental applications. For example, Miao and co-workers developed M–N-CNT (M = Co, Fe, Mn, or Ni) Fenton-like SACs via pyrolysis and acid etching, focused on the relationship between the transition metal spin state and catalytic activity, and provided a deep understanding and clarification of the mechanism of metal element-regulated PMS activation (Fig. 8a).99 The Curie–Weiss law102 can be expressed as follows:

 
image file: d2nr02989h-t1.tif(1)
 
image file: d2nr02989h-t2.tif(2)
where C and μeff are the Curie constant and the effective magnetic moment, respectively. μeff can be calculated through the linear fitting of χ−1T. μeff can also be obtained from the spin states (S) of the 3d transition metals because its intrinsic origin is the spin splitting of d orbitals (eg and t2g) not fully filled with electrons.103
 
image file: d2nr02989h-t3.tif(3)
 
g = 2(4)
 
VHS + VLS = 1(5)


image file: d2nr02989h-f8.tif
Fig. 8 Schematic diagram of the process for the preparation of M–N-CNTs and characterization of metal spin properties: (a) key synthesis route of M–N-CNTs. (b) According to the Curie Weiss law, the temperature-dependent anti-polarizability curve related to the anchored metal species in M–N-CNT Fenton-like SACs. (c) Relationship among M-PMS*, M–N and PMS and effective magnetic moment (μeff) in M–N-CNTs. (d) Linear correlation curve between Eads and experimental results (p < 0.05). (e) Linear correlation curve between μeff and Eads (p < 0.05) and (f) model of possible oxidation reaction mechanism in M–N-CNTs.99 Copyright 2021, the American Chemical Society.

According to this formula, although the spin states of transition metals are complex and diverse, the S value is always positively correlated with the μeff value. Also, the experimental μeff and S value follow the order of Co–N-CNTs (μeff = 4.37μB, S = 2 or 3/2) > Ni–N-CNTs (μeff = 2.24μB, S = 1) > Fe–N-CNTs (μeff = 1.75μB, S = 1/2) > Mn–N-CNTs (μeff = 1.69μB, S = 1/2) (Fig. 8c). Based on the space electron configuration theory of the d orbitals in transition metals, they correspond to the interaction of oxygen-containing adsorbates on the metal central unit in the form of a large magnetic moment and high spin state and facilitate the rapid transfer of electrons with spin orientation. Therefore, the PMS adsorption, intermediate oxidation potential and electron transfer in the catalytic process are significantly improved. Meanwhile, the linear regression curves of the S value and adsorption energy (Eads) and Eads and μeff were established, respectively (Fig. 8e). According to the linear correlation coefficient, it was further clarified that the intrinsic factor of PMS activation is Eads, whereas the external factors controlling its activation are the magnetic moment and spin state of transition metal in the M-PMS system. Therefore, due to the spatial configuration and spin state of the d orbital of the single-metal atom M in M–N-CNTs, PMS (or O–O group) prefers to adsorb on the M–N site, generating M-PMS* intermediates with a higher oxidation potential and accelerating the electron transfer rate, resulting in the fast activation of PMS and decomposition of SMX via the non-radical process (Fig. 8f).

Interestingly, the synergy between the active metal atoms and the introduced second promoting metal atoms not only can increase the number of new Fenton-like reaction centers, but also improve the activity and stability of metal species, playing a greater role in continuous or intermittent cyclic reactions. For example, Xu's team selected Fe and Mn bimetallic species with low biological toxicity and rich geological reserves and electrostatic self-assembly complexes of Prussian blue analogues as precursors to prepare nitrogen-doped graphene-coated FeMn bimetallic catalysts (FeMn@N–C) through high-temperature pyrolysis to activate PMS and produce active free radicals in a Fenton-like process to achieve the mineralization of chlorothiazide (CTD).104 Taking FeMn@N–C as an example, this work focused on the synergistic effect of Fe and Mn in the Fenton-like oxidation process. Between them, the thermodynamically favorable electron transfer from Fe(II) to Mn(III) reduces the redox potential barrier between different valence states of the same metal species and accelerates the cyclic regeneration of high-valence and low-valence metal active centers. The DFT theoretical analysis showed that the redox cycle of multivalent FeMn bimetallic Fenton-like catalysts can accelerate the formation of free radicals and ensure the stability of iron under continuous PMS activation. As is known, the chemical rate constant of the reduction of the high valence state to the low valence state of transition metal species is very low, which has become the bottleneck in the transformation of Fenton-like catalysis to practical industrial application. Therefore, the strategy of the synergistic effect of multiple transition metal species can realize the rapid cycling and regeneration between redox valence states, realizing a breakthrough in the decisive step of Fenton oxidation to a certain extent.

In conclusion, the precise regulation of transition metal species can significantly optimize the oxidation activity and stability of Fenton-like SACs. (i) Different transition metal elements anchored in the same coordination environment may lead to different catalytic activity and mineralization mechanisms in Fenton-like SACs. (ii) The unique electronic properties (oxidation state or spin state) of each metal species may also affect the coordination and anchoring mode of coordination atoms provided by ligands, thus forming different geometric structures and affecting the oxidation activity. (iii) The introduction of new transition metal species may significantly optimize the oxidation kinetics of Fenton-like SACs owing to the synergistic effect between different metal species. Noteworthily, the results by the Duan and Miao research groups reflect that the activity order of a single metal center is not consistent, which may be due to the different coordination active sites.99,104 In addition, it also shows that the coordination of nitrogen-doped carbon atoms with transition metal species will also have a great impact on the catalytic performance. Therefore, the coordination environment may also play a key role in regulating the electronic structure and oxidation–reduction activity of Fenton-like catalysts. A detailed description of this aspect is reflected in the next section.

4.2 Coordination environment

The coordination environment (coordination species or bond number) between the active transition metal species and their neighboring atoms is of vital important in the catalytic activity and stability of Fenton-like SACs.105 The oxidation activity and catalytic performance of the so-called heterogeneous monatomic catalysts cannot be completed independently by the separated and anchored transition metal species. Instead, Fenton-like catalytic reactions are completed on the active units formed by coordination between isolated metal atoms and surrounding electronegative heteroatoms (e.g., C, O, N, S and B).106–110 At present, the widely studied active centers (e.g., M–Cx, M–Nx and M–N–C) are composed of transition metal species interacting with heteroatoms in ligands in the form of covalent or ionic bonds.111,112 Importantly, the existence of coordination atoms, especially the electrostatic interaction between anchored non-metallic atoms and transition metal elements, prevents the agglomeration trend of isolated metal elements, thus improving the chemical and structural stability of Fenton-like SACs. Moreover, the electronic properties and geometry of transition metal reaction centers are mainly determined by the local coordination chemistry, which is owing to the coupling effect of the charge density in the center of the metal element regulated by the anchor donor.113–115 Therefore, it is necessary to explore the coordination environment around highly dispersed transition metal catalytic centers, laying a foundation for further research on the design and mechanism of Fenton-like SACs in the future.

Taking the widely studied M–Cx reaction centers as an example, most of them capture and fix transition metal atoms with carbon materials rich in oxygen functional groups (e.g., –COOH and –OH), such as graphene oxide (GO) and carbon nanotubes (CNT).116–119 The coordination of the electronegative C atoms plays the role of fixing and isolating metal atoms. This not only realizes the stability of the chemical structure of Fenton-like SACs and the individual transition metal elements and makes the catalyst more durable and significantly improves the atomic utilization rate, but also improves the overall conductivity of the catalyst, achieving the purpose of rapid electron transfer in Fenton-like catalysis. For example, Li's group designed a Co-based Fenton-like SAC (SACo-NG) with a metal cation–π structure, which was manufactured through enhanced hydrothermal-sintering method with isolated Co atoms highly dispersed on the nanospheric C-based graphene-like structures.120 This Co cation–π (Co2+–N–Cπ) structure has strong polarity due to the existence of rich/poor electronic micro-regions and its π → cation interaction accelerates the interfacial electron transfer cycle, which exists between the pollutants adsorbed on the catalyst surface (due to the π–π stacking between the catalyst aromatic ring and the pollutant aromatic ring) and the electron contribution of the pollutants in the electron deficient center. The orbital interaction (metal cation-π structure) between transition metal species and coordination atoms can successfully regulate the transfer of interfacial electrons, which is an excellent case in the field of rapid and efficient mineralization of pollutants.

Doping other heteroatoms (such as B, O, S, N and P) can significantly improve the electrochemical properties of the graphene carbon layer and produce more structural defects. These defects not only become the new active center of Fenton-like oxidation, but also increase the loading content of metal elements due to their coordination unsaturation. For example, Li et al. reported the synthesis of a Cu-based Fenton-like catalyst (Cu-SA/NGO) by dispersing a single Cu atom in N-doped graphene synthesized via the pyrolysis and freeze-drying of melamine, cyanuric acid and graphene oxide precursors, with a relatively high Cu loading of 5.8 wt% (which was the highest value reported thus far).121 In addition, the N-doped types such as graphitic N, pyrrolic N and pyridinic N will also have a significant impact on the catalytic performance. For example, He and co-workers used nitrogen-doped porous carbon as a ligand anchor to prepare a single Mn atom catalyst (Mn-ISAs@CN) with outstanding catalytic performance.122 A unique perspective for the design of Fenton-like SACs was the dual reaction site strategy to achieve an efficient catalytic performance. Specifically, N-coordinated single Mn element (MnN4) is a highly catalytic active center for PMS activation, while the adjacent pyrrolic N site is the adsorption unit of the target organic molecule. The interaction between the two significantly shortens the migration distance between the free radicals and adsorbed organic molecules. Differently, the high-resolution N 1s XPS analysis of the g-C3N4-anchored single iron atom catalyst (SA Fe-g-C3N4) by Hu et al. confirmed the linear correlation between the pyridine N content (wt%) and iron content (wt%) (R2 = 0.97), that is, strong linear relationship between the pyridinic N and Fe loading content.123 The adsorbate on the catalyst surface can be removed by easy heat treatment, and the catalytic activity of the catalyst can be well restored.

Moreover, in this Fenton-like catalyst, in which the carbon material introduced by the electronegative non-metallic heteroatom N is coordinated and anchored with a 3d metal atom, the electron-rich N dopant can be used as the Lewis basic site, which makes it easier to fix and disperse the metal atom (Lewis acid) and forms a strong M–N (M refers to Fe, Co and Mn) reaction activation unit (M–N–C) in the carbon matrix.124 Although there are relatively few studies on the coordination of S, P and O compared with C and N, the existence of these coordination atoms will also affect the chemical structure or local electronic properties of Fenton-like catalysts. As proposed in the literature, electron-absorbing oxidized S species can weaken the adsorption of intermediates on transition metal species. Du and co-workers successfully developed Fenton-like SACs, which can be extended to various heteroatom doping and coordination in a simple and low-cost way.82 Specifically, biomass-derived biosorbents containing high quaternary ammonium groups can be applied to the adsorption of sulfate or phosphate to realize the co-doping of S or P and N. The scalability and practical industrial application of the synthesis technology lay a foundation for further research.

Overall, the coordination environment around the transition metal species plays an important role in regulating the stability, catalytic activity and selectivity of Fenton-like SACs, which is mainly reflected in the following aspects. (i) The interaction between coordination atoms and isolated single metal elements can improve the stability of the catalytic unit coordination chemical structure and prevent the migration and agglomeration of metal atoms in the Fenton-like catalytic process. (ii) The introduced coordination heteroatoms can produce more unsaturated hanging bonds, thus increasing the position of capturing and anchoring metal species and improving their loading content. (iii) The existence of coordination atoms will also improve the local electronic properties of the Fenton-like SAC catalytic sites. Although numerous reports in the literature have investigated the coordination environment of Fenton-like catalysts, there are relatively few studies on the number of coordination chemical bonds compared with the reports on different coordination heteroatoms. Simultaneously, in the coordination configuration of the same coordination atom, different coordination bond numbers will also significantly affect the stability and electronic structure of the catalytic site. For example, compared with the M–C4 site with a stable carbon double vacancy coordination, the unstable M–C3 site is coordinated by transition metal species with a single carbon vacancy.125 To date, most of the active sites in ligand-anchored Fenton-like catalysts are M–Cx, M–Nx and M–N–C. In contrast, few M–Sx, M–Px and M–Bx active centers have been reported, and thus further experiments, characterization and simulation data are still needed to deeply elucidate the structure and activity of these reaction units.

4.3 Support/carrier

As mentioned above, selecting appropriate supports is an effective strategy to overcome the difficulty of the migration and aggregation of isolated single metal species with high surface energy in the synthesis of Fenton-like SACs. The rigid coordination (such as coordination, electrostatic adsorption and ionic bond) between the single metal reaction sites and rigid supports ensures the durability of the catalyst, inhibits the agglomeration of inactive metal atoms and solves the bottleneck of poor cyclic stability in the Fenton-like oxidation process, even if the catalytic process occurs under extremely bad reaction conditions. In addition, the high specific surface area of the carrier can maximize the exposure of high-density monatomic active sites fixed on the substrate and enhance the mass transfer. Heterogeneous Fenton-like SACs with 0D, 1D, 2D and 3D surface morphologies have been developed to improve the catalytic activity.

The properties of the support (such as conductivity and semiconductor properties) used to stabilize the catalytic unit may also endow the catalyst with high oxidation activity and determine the application direction of the catalyst in the fields of chemistry, photocatalysis and electrocatalysis. At present, the most widely studied carbon-based substrate materials can optimize the conductivity of catalysts, which can accelerate the transfer of electrons from the bulk catalyst excited by light, electricity, heat or radiation to the reaction sites on the catalyst surface and participate in the catalytic reaction.126–128 In addition, the use of carbon-based supports is also an effective strategy to solve the secondary pollution caused by metal ion leaching in heterogeneous Fenton-like SACs. For example, Gao et al. used in situ CVD technology to synthesize Fenton-like SACs with a single iron atom anchored on three-dimensional (3D) N-doped carbon nanosheets (Fe/NC) on a ferrocene-loaded CaO hard template (carbon source and nitrogen source from pyridine).129 Experimental and theoretical calculations showed that electrons can be rapidly transferred from the C to Fe species through the C–N–Fe bond (formed by the combination of a carbon substrate and catalytic site Fe–N4), which ensures that Fe species always maintain a low valence state in the redox process. Meanwhile, the presence of reducing metal species can also significantly inhibit the oxidation of carbon-based substrates.

The choice of semiconductor-based substrate also enables Fenton-like SACs materials to achieve catalytic oxidation, while also having the properties of semiconductors, and determines their application in the fields of light, electricity and chemical catalysis. For example, Wang et al. reported a coupling system in which the nanotube-assembled 3D hierarchical H2-reduced Mn-doped CeO2 micro-flowers (re-Mn–CeO2 NMs) were synthesized as Fenton-like photocatalysts for the activation of PMS.130 The obtained re-Mn–CeO2 NMs can achieve more than 95% organic matter degradation in the re-7Mn-CeO2 NM/PMS/Vis system within only 10 min, which was 1.1-times and 2.0-times higher than that of Fenton-like reaction and photocatalysis alone, respectively. It was confirmed that the synergistic effect of the Fenton-like reaction and photocatalysis in the coupling system has a great application prospect in environmental remediation.

In summary, the construction of strong metal–support interaction is very important to improve the activity and stability of bulk Fenton-like SACs. The effective selection of the support is significant to improve the oxidation activity and stability of Fenton-like SACs and achieve their practical application in environmental remediation. (i) A support with high specific surface area can maximize the exposure of active centers anchored on its surface. (ii) The selection of a carbon-based support can improve the conductivity of the bulk Fenton-like SACs, thus accelerating the electron transfer and optimizing the catalytic oxidation performance. (iii) Semiconductor-based materials such as g-C3N4, TiO2, Fe2O3, WO3 and BiVO4 as anchored single-atom reaction units can give new performances to the whole Fenton-like catalytic system. For example, the coupling and synergy of photocatalysis/electrocatalysis and Fenton-like catalysis reported in the literature can improve the catalytic effect exponentially.

5. Mechanism

Fenton-like SACs with strong metal ligand interaction constructed by the coordination of isolated metal atoms on carefully selected carriers in the form of covalent/ionic bonds have attracted extensive attention of researchers because of their maximum atomic utilization, excellent catalytic activity, and easy recycling.13,51,54,65 However, with the change in the metal-atom catalytic sites designed in the Fenton-like process (such as the type and loading of isolated metal species, as well as the surrounding electronic configuration, chemical structure, coordination environment and bond number), the catalytic oxidation rate will also be enhanced or inhibited. Therefore, the rapid development of various advanced instruments and theoretical calculations makes it clearer and more meaningful to visually define the catalytic sites of a single metal atom and simulate the catalytic mechanism.50 This can deeply clarify the valence state and coordination environment of the reaction unit in the Fenton-like oxidation process, as well as its interaction with oxidants or target pollutants, electron transfer, peroxide activation, and the generation and evolution of different reactive oxygen species (ROS).90 In this section, the catalytic mechanism of Fenton-like SACs involving the adsorption and activation of peroxides (H2O2, PMS, and PDS) to generate radicals (e.g., HO˙, SO4˙, and O2˙) and nonradical (e.g., direct electron transfer, singlet oxygen, and high-valent metal-oxo species) to realize the mineralization of target pollutants in organic wastewater is described in detail. This can provide a theoretical basis for future catalyst design and ensure the high kinetic mineralization of target pollutants in environmental remediation.

5.1 Radical pathways

The radical-dominated catalytic mechanism has been widely studied and recognized in Fenton or Fenton-like systems. In particular, reactive oxygen species (ROS) (e.g., HO˙, SO4˙, and O2˙) have been identified as effective AOPs for mineralizing refractory organic compounds.22,127 The generation of the highly active HO˙ depends on the activation of H2O2 adsorbed on the surface reaction sites of Fenton-like catalysts, while SO4˙ is mainly generated in two main persulfate precursor oxidants excited by light, heat, electricity, ultrasound or transition metals, namely persulfate (PS), PMS or PDS/Fenton-like systems.131 Among them, PMS (also known as Oxone) is a triple salt composed of 2KHSO5, KHSO4, and K2SO4.132 The type of oxidant selected in the reaction system plays an indispensable role in the completion of the oxidative degradation function of the designed Fenton-like catalyst. Therefore, excluding all the external conditions of the catalytic reaction, such as catalyst type, target degradation products, and oxidation conditions (such as pH, temperature, catalyst dosage, and reaction time), by changing only the type of oxidant can enable researchers to obtain the best oxidant species from an intuitive experimental point of view. For example, Liu and colleagues used Fe-encapsulated B/N-doped carbon nanotubes (Fe@C-NB) as a catalyst and compared the differences in catalytic oxidation efficiency caused by different oxidants (H2O2, PMS and PDS) in the treatment of p-arsanilic acid (p-ASA), and the experimental results were ranked as PMS > PDS > H2O2.133 Obviously, in Fenton-like oxidation, PMS exhibits excellent oxidation sensitivity compared with the existing commercial oxidants PDS and H2O2 because PMS molecules (HO3S–O–O–H; IO–O = 1.326 Å) have a longer asymmetric structure of the superoxide O–O bond (PDS: HO3S–O–O–SO3H; IO–O = 1.322 Å).132 Therefore, PMS, which is mild and easy to store and transport, is a promising wastewater oxidation repair agent. Comparing and listing the advantages of the active oxygen radical SO4˙ generated by the activation of PMS/Fenton-like catalyst systems with HO˙: (a) SO4˙ shows a high oxidation potential (SO4˙, 2.5–3.1 V vs. NHE HO˙, 1.8–2.7 V vs. NHE); (b) SO4˙ exhibits better flexibility to a wide pH range (SO4˙, 3–9 vs. HO˙, 2–4); (c) SO4˙ has a longer half-life (SO4˙, 30–40 μs vs. HO˙, 20 ns); and (d) SO4˙ is more selective than HO˙ towards electron-rich organic micropollutants.63,81,86 Therefore, these excellent characteristics of SO4˙ make it possible to act for a longer duration to rapidly degrade toxic organic pollutants, especially persistent organic pollutants, into smaller biodegradable molecules, even H2O and CO2.

Radical quenching experiments can determine the types of active radical species generated during the activation of PMS. Specifically, methanol (MeOH) is a radical scavenger for HO˙ and SO4˙, while tert-butanol (TBA) and isopropyl alcohol (IPA) are used to prove the ˙OH contribution to the reaction.6,9,22,28 Electron paramagnetic resonance (EPR) is a sensitive and convenient technique that uses spin-trapping agents to further verify the ROS species participating in the Fenton-like reaction process.55,63 Taking the Fe@MoS2/PMS system as an example, the active species that play a major role in the Fenton-like oxidative degradation of atrazine (ATZ) were characterized and confirmed by EPR and radical quenching experiments.134 As shown in Fig. 9a, the spin capture EPR experiment using 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as a trapping agent showed that both HO˙ and SO4˙ play a role in the MoS2/PMS and Fe@MoS2/PMS systems, which are marked by pink circles and brown triangles, belonging to the DMPO-HO˙ adducts (four peaks) and DMPO-SO4˙ adducts (six peaks), respectively. Subsequently, a series of quenching experiments in the presence of 100 mM ethanol (EtOH) or TBA showed that they could inhibit the degradation of ATZ (40 μM) by 91.8% and 36.9%, respectively (Fig. 9b). The difference in the inhibition degradation data between the two groups was explained based on the second-order rate constant of the reaction between EtOH with the radicals (kHO˙, EtOH = 1.7–2.2 × 109 M−1 s−1, image file: d2nr02989h-t4.tif, EtOH = 1.6–7.7 × 107 M−1 s−1) and TBA with radicals (kHO˙, TBA = 4.2–7.6 × 108 M−1 s−1, image file: d2nr02989h-t5.tif, TBA = 4–9.1 × 105 M−1 s−1), clearly showing that EtOH can effectively quench HO˙ and SO4˙ radicals. Moreover, although it was considered to be more effective in the quenching HO˙ radical, TBA can only quench about 30% of sulfate, which is mainly due to the instability of the DMPO-SO4˙ adduct and easy conversion to DMPO-HO˙ due to the nucleophilic substitution reaction with OH/H2O. This may be why it was observed that the DMPO-HO˙ signal was larger than the DMPO-SO4˙ signal, similar to the phenomenon in previous studies. Therefore, based on the above-mentioned experimental results, it can be further confirmed that HO˙ and SO4˙ play a significant role in the decomposition of target pollutants into small molecules or even CO2 and H2O.

 
SO4˙ + H2O → SO42− + HO˙ k < 2 × 103 M−1 s−1(6)


image file: d2nr02989h-f9.tif
Fig. 9 Radical quenching experiments and catalytic mechanism of Fe@MoS2: (a) EPR spectra, (b) inhibition of different free radical scavengers on ATZ decomposition in Fe@MoS2/PMS system, and (c) schematic diagram of possible oxidative degradation on Fe@MoS2.134 ©2021, Elsevier B.V. All rights reserved. Mechanism characterization results: (d) ΔG on Au/VO2 material and possible adsorption units in the form of an asterisk (*).90 ©2021, Wiley-VCH GmbH. (e) In situ Raman spectra of SACo-NGs/PMS/BPA system.120 ©2021, Elsevier B.V. All rights reserved.

Previous studies have shown that the heterogeneous transition metal sulfide MoS2 activates MoS2/PMS through the oxidation of Mo(IV) to Mo(V) and further to Mo(VI), and then decomposes to the high-activity SO4˙.135 According to this degradation experiment, the PMS activation pathway was proposed in Fig. 9c. Specifically, the oxidation reaction process involves three steps, as follows: (i) the Mo atoms on the edge and basal surface contact with PMS and adsorb to form Mo-PMS ligands, which can be verified according to the low adsorption energy of PMS on Mo and the increased length of the peroxy bonds of PMS after the formation of the Mo-PMS ligand in DFT calculations and (ii) the Mo(VI)/Mo(V)/PMS system generates HO˙ and SO4˙. According to the XPS characterization of Fe@MoS2 before and after the catalytic experiment, a new small peak of Mo(VI) appeared after the cyclic experiment, which may be due to the electron exchange between Mo(IV) and PMS. Additionally, (iii) the generated HO˙ and SO4˙ radicals play an indispensable role in the treatment of ATZ to achieve its decomposition and mineralization. Importantly, the loaded Fe atoms may also contribute to the activation of PMS and can be reduced by Mo(IV), and the Fe atoms not only activate the adjacent Mo atoms, but also the Mo atoms not adjacent to Fe, thus adsorbing and activating the PMS molecules. This result was consistent with the EPR experimental results, that is, the intensity of the radicals, especially HO˙ in the Fe@MoS2/PMS system was much higher than that in the MoS2/PMS system. It was further confirmed that the Fe atom anchored on the MoS2 support plays an important role in the Fenton-like process. The Fe atom with high atomic utilization can activate the oxidant and produce more reactive oxygen species to degrade ATZ, thus effectively improving the catalytic activity of Fenton-like reactions.

The reaction Gibbs free energy (ΔG) is a thermodynamic function introduced in chemical thermodynamics to determine the direction of reaction process. The ΔG of the Au/VO2/PDS Fenton-like system was calculated, as shown in Fig. 9d.90 It can be concluded that the adsorption and decomposition of persulfate were both spontaneous in the reaction process because the formation of S2O82−* and SO4˙* was an exothermic reaction. Moreover, due to the anchoring of isolated Au atoms, it was more conducive for electrons to transfer to the PDS adsorbed on the surface. Therefore, in the persulfate decomposition step, the reaction decomposition energy of Au/VO2 was 1.83 eV, which was much less than that of the other two catalysts (5.30 eV for Au (cluster)/VO2(Au(C)/VO2) and 7.41 eV for VO2(B)). More importantly, the rate-limiting step, that is, the desorption reaction of SO4˙*, is also significantly accelerated, which was reflected by the fact that the desorption energy of Au/VO2 was much lower than that of Au(C)/VO2 and VO2. Therefore, the acceleration of the rate-limiting step can greatly improve the desorption capacity of the persulfate radical, which is beneficial for the subsequent oxidative decomposition reaction between ROS and organic macromolecules and acceleration of the whole sewage purification process.

In situ characterization is an important means to study the catalytic mechanism. Here, the adsorption, activation and evolution of PMS on the catalyst surface are clearly illustrated by in situ Raman spectroscopy. The adsorption of PMS on the catalyst surface is a prerequisite for starting all oxidation reactions, and thus it is particularly important to study the initial state (Fig. 9e). The in situ Raman spectrum of the pure PMS system has three main peaks at 878 cm−1, 973 cm−1 and 1052 cm−1, representing the stretching vibration of the O–O bond and the symmetrical stretching vibration of SO42− and SO3 in HSO5, respectively.136 In the NG/PMS and NG/PMS/BPA systems, only the peak representing O–O bond was different from the pure PMS solution and shifted to a lower wavenumber. It is worth noting that in SACo-NG/PMS solution, the peak of the O–O bond splits at 866 cm−1 and 882 cm−1, indicating that two different species in PMS adsorbed on the surface of SACo-NGs react with the O and Co sites, respectively.120 The peak intensity corresponding to HSO5 decreased significantly, indicating the consumption of PMS. After the addition of BPA to the SACo-NGs/PMS/BPA system, the three characteristic peaks and splitting peaks of PMS became very weak, indicating that PMS reacted with the Co site and decomposed. Interestingly, the presence of BPA in solution accelerates the oxidation of PMS, which was reflected by the emergence of a new peak corresponding to some peroxy species, such as SO5˙, at 803 cm−1.

5.2 Nonradical pathways

Non-radical oxidation has been recognized in many catalytic PMS oxidations, where the formation of singlet oxygen (1O2) and mediated electron transfer are the typical mechanisms.137 The oxidation degradation pathway dominated by nonradicals has wide applicability and adaptability for practical use in environmental remediation because it plays a significance role in real wastewater systems with anions such as humic acid (HA), Cl, HCO3, H2PO4, NO3 and SO42− compared with the traditional active radicals.82,134
5.2.1 Singlet oxygen. Singlet oxygen (1O2), a type of molecular oxygen in the excited state, belongs to reactive oxygen.138 It has the characteristics of zero total spin and no paramagnetism. Although singlet oxygen is not a free radical, its reaction activity is much higher than that of ordinary oxygen due to the lifting of the spin restriction. After the ground-state oxygen atom (triplet oxygen molecule, 3O2) is excited, two electrons with parallel spins in the original two 2pπ* orbitals can occupy one 2pπ* orbitals simultaneously with opposite spins or occupy two 2pπ* orbitals, respectively, with opposite spins. The two excited states S = 0 and 2S + 1 = 1, that is, their spin multiplicity is 1, are a singlet state (represented by 1Δg and 1Σg+, respectively).139 Moreover, the energy of singlet oxygen is higher than that of triplet oxygen, and thus it needs to absorb some energy to change. In Fenton-like reaction systems, this process is realized by various types of catalysts (enzymes). This process directly induces the formation of the superoxide ion O2˙. The superoxide ion O2˙ is a superoxide anion radical because it has a single electron in the π* orbital. This radical has strong oxidation ability because it has the tendency to seize electrons and pair single electrons. Thus, 1O2, a type of molecular oxygen in the excited state, is similar to reactive oxygen species such as superoxide anion radical, hydroxyl radical and hydrogen peroxide, and also plays an important role in catalytic oxidation.

The main way to verify the singlet oxygen (1O2) in the catalytic system is radical quenching experiments using NaN3, KI, azide, L-histidine and furfuryl alcohol (FFA) as scavengers and EPR analysis using DMPO and TEMP as trapping agents. As is well known, KI is usually used as a strong quencher of surface-bound free radicals in scavenging tests.82 Moreover, the reaction between the capture reagent 5,5-dimethyl-1-pyrroline N-oxide (DMPO) and strong oxidizing agent 1O2 generates 5,5-dimethyl-1-pyrrolidone-2-oxyl (DMPOX, seven-line spectrum and the characteristic signals with aN = 7.2 G and arH = 4.1 G) adduct through the secondary oxidation–reduction procedure instead of the DMPO-SO4˙ and DMPO-˙OH signals.123 The distinguishable and typical 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 triplet signal was detected in the EPR spectrum with TEMP as a capture agent, which represents the TEMP-1O2 adduct.

Thus far, research has been proven that there are four main mechanisms for the generation and evolution of 1O2, which are described in detail as follows: (i) the self-decomposition of PMS, (ii) originating from O2, which is activated by oxygen vacancy defects, and this pathway can be verified by controlling gas experiments (nitrogen and air bubbling) in the reaction solution and the released O2 receives electrons and reduced to O2˙. (iii) From OL (oxygen in lattice).140 For example, the injection of nitrogen gas into the mixed solution of 5%Ca–Fe2O3/PMS and RhB molecules proves that 1O2 comes from the lattice oxygen in the 5%Ca–Fe2O3 complex rather than the dissolved oxygen in the solution.141 The formation of lattice oxygen comes from Ca doping, which leads to the partial substitution of strong bound oxygen by relatively weak bound oxygen, which is released into the solution and transformed into active oxygen species (O*), and then O* is converted to 1O2 in the presence of PMS, as shown in the following eqn (7)–(11):

 
HSO5 → SO52− + H+(7)
 
SO52− + HSO5 → HSO4 + SO42− + 1O2(8)
 
O2 + e → O2˙(9)
 
Ovac → O*(10)
 
O* + HSO5 → HSO4 + 1O2(11)

Moreover, the optimized electron conduction of Ca doping is also beneficial for the generation of 1O2, which is consistent with the previous research conclusions. As shown in Fig. 10, according to the Bader method, the electrons transferred from the α-Fe2O3 and 5%Ca–Fe2O3 surfaces to PMS was calculated to be 0.54 and 0.63, respectively, verifying the optimization of Ca doping on the electron transfer ability of α-Fe2O3. In addition, enhancing the adsorption energy (Eads) on the catalyst surface is more favorable for the adsorption and activation of PMS on the 5%Ca–Fe2O3 surface (−2.17 eV for α-Fe2O3 and −2.56 eV for 5%Ca–Fe2O3). It is of great significance to activate PMS to produce 1O2 by optimizing the adsorption and electron transfer capacity, which has been proven both experimentally and theoretically.


image file: d2nr02989h-f10.tif
Fig. 10 Electron density difference (EDD) diagram of Ca atom catalyst material coordinated and dispersed with α-Fe2O3 as a carrier: (a) 5%Ca–Fe2O3 and (b) with PMS adsorbed on the (110) crystal plane.141 ©2019, Elsevier B.V. All rights reserved. Schematic diagram of DFT theoretical characterization of electrostatic potential distribution on various graphene fragments: (c) pure graphene, (d) N-doped graphene, (e) PMS molecule and (f) PMS-adsorbed graphene.137 ©2019, Elsevier B.V. All rights reserved.

(iv) From O2˙ oxidized to 1O2 by h+, and O2˙ is disproportionate to 1O2. Also, the existence of O2˙ can be verified via the scavenger BQ.

It is well known that the electron-rich ketonic groups (C[double bond, length as m-dash]O) located on the boundaries of the sp-conjugated carbon lattice play a significant role in interacting with PMS through nucleophilic addition and mediating dioxirane intermediates to generate 1O2. For example, the radical pathways of O2˙ and non-radical 1O2 have been proven to play a major role in the catalytic oxidation of PCM in the Fe–N@C/PMS system.137 Between them, the mechanism that the formation of 1O2 comes from the carbonyl active sites on the surface was experimentally verified and theoretically calculated. This proved the charge conduction mechanism of 1O2 generation between PMS (electron donor) and the electrophilic C[double bond, length as m-dash]O group (electron acceptor), which was also proposed in previous studies. Specifically, PMS releases electrons to generate anionic radicals (SO5˙), which then further combine to form S2O82− or SO42− ions, and 1O2 is also produced in this step.138 In addition, O2˙ also plays a key role in the formation of 1O2 in the Fe–N@C/PMS system. For example, in the charge transfer mechanism, there are two ways to generate 1O2, which are related to O2˙[thin space (1/6-em)]:[thin space (1/6-em)]O2˙ oxidized to 1O2 by h+ and O2˙ was disproportioned to 1O2. Both approaches are highly dependent on O2˙.140 Moreover, the recombination and self-decomposition of the superoxide anion radical and hydroxyl radical (O2˙/HO˙) (or hyper-hydroxyl radical, O2˙/HO2˙) can also lead to the formation of 1O2 (eqn. (12)–(18)). In conclusion, based on the research mechanism of PMS activation and degradation of pollutants, it was clarified that 1O2 has multiple evolution pathways, mainly including the self-oxidation of PMS, the combination of PMS with carbonyl parts or carbonyl on carbonaceous materials, and the evolution of O2˙.

 
HSO5 → H+ + SO5˙ + e(12)
 
SO5˙ + SO5˙ → 2 SO42− + 1O2(13)
 
SO5˙ + SO5˙ → S2O82− + 1O2(14)
 
O2˙ + HO˙ → 1O2 + OH(15)
 
2O2˙ + 2H+ → H2O2 + 1O2(16)
 
2O2˙ + 2H2O → H2O2 + 2OH + 1O2(17)
 
HO2˙ + O2˙1O2 + HO2(18)

The electrostatic potential (ESP) diagram of reduction and oxidation active sites represented by positive/negative regions was used to study the potential distribution of graphene fragments doped with graphitic N atoms before and after the adsorption of PMS molecules.137 After the introduction of graphitic N on the surface of graphene fragments, ESP becomes uneven, which is opposite to the uniform distribution of pure graphene fragments on each C atom (Fig. 10c–f). Moreover, the positive potential was mainly concentrated around the N atom of graphite, while the other atoms were not significantly affected, which indicates that the uniform carbon network was severely damaged and has a high chemical potential. Free PMS molecules show a very negative potential around their atoms and the possible adsorption model of PMS molecules on graphitic N atoms is shown in Fig. 10d. The ESP spectra show that the most electronegative part of PMS adsorbed on the catalyst surface was the area around the binding of PMS and graphitic N atoms, and the adsorption energy of graphitic N and PMS molecules (ΔEads) was calculated to be −2.62 eV, which was conducive to promoting PMS activation and nonradical oxidation on the catalyst surface due to the strong binding affinity of graphitic N to PMS molecules (Fig. 10f). As described above, it can be concluded that the doped graphitic N atom is the generation center of the nonradical 1O2, which plays a leading role in the Fe–N@C treatment PCM reaction.

1O2 can be formed in many free radical reactions. However, there is a shortage of methods that can be used for the study and analysis of singlet oxygen. This is mainly because to distinguish 1O2 from other reactive oxygen species, on the one hand, it requires analytical methods with high selectivity. On the other hand, due to the short life (∼4.2 μs) and low yield of 1O2 in aqueous solution, the method used should also have high sensitivity.142 Additionally, the interference and even inhibition of catalytic activity by anions (such as CO32−, Cl, SO42− or NO3) and some natural organic compounds (Nom) can be alleviated by the singlet oxygen nonradical process during sewage purification. Therefore, the development and utilization of excellent PMS activators based on 1O2 can fully meet the requirements of real wastewater treatment.

5.2.2 Electron-transfer regime. The direct electron transfer pathway at the Fenton-like catalyst surface/interface that activates PMS and decomposes organic pollutants can be characterized and confirmed by cyclic voltammetry (CV), linear sweep voltammetry (LSV), electrochemical impedance spectroscopy (EIS), and in situ measurement of the open circuit potential, which are common electrochemical research methods. Among them, cyclic voltammetry recording the current potential curve can be employed to determine the reversibility of the electrode reaction, the possibility of intermediates, phase boundary adsorption or new phase formation, and the properties of coupling chemical reactions according to the shape of the curve.143–145 Electrochemical impedance spectroscopy involves the application of a small amplitude AC potential wave with different frequencies to an electrochemical system and measuring the change in the ratio of AC potential to current signal (this ratio is the impedance of the system) with the sine wave frequency (ω), or the change in the phase angle (Φ) of impedance with ω. The principle of this technology is to regard the electrochemical system as an equivalent circuit composed of basic elements such as resistance (R), capacitance (c) and inductance (L) in series and parallel. The EIS curve can reflect the composition of the equivalent circuit and the size of each element, enabling the further analysis of the structure of the electrochemical system and the properties of the electrode process.146

Taking the Fenton-like oxidation of the phenol molecule attacked and decomposed by the Co@NG-900 catalyst as an example, a series of the above-mentioned electrochemical analyses and characterizations was carried out to confirm its nonradical oxidation performance.147 Among them, the LSV could can be divided into three stages, as follows: (i) when the electrode potential (φ) was in the range of 0.40–0.57 V, the j value remained almost unchanged (fluctuating around 1.5 mA cm−2), and the oxidation reaction of phenol could not be carried out, indicating that the oxidation of phenol can only occur above this electrode potential; (ii) in the range of 0.57–0.82 V, the oxidative decomposition of phenol was carried out on the anode and produced a sharply increased current; and (iii) the current density showed a downward trend when the potential exceeded 0.82 V, which may be due to the rapid consumption of phenol, resulting in the oxidation rate on the anode in the reaction system being much higher than the mass transfer rate of phenol molecules. The corresponding Tafel curve was drawn according to the LSV curve, and the Tafel slope was obtained by fitting the LSV data to the Tafel curve. The Tafel slope is usually determined by the rate-limiting step in electrocatalysis, which can reflect the intrinsic characteristics of a catalyst. The Tafel slopes of Co@NG-900 and NG-900 were 266 and 188 mV dec−1, respectively (inset in Fig. 11a), indicating that Co@NG-900 exhibits a faster reaction rate and higher application potential than NG-900, which may be due to the strong interface interaction between the Co and NG shells.


image file: d2nr02989h-f11.tif
Fig. 11 Schematic diagram of the electron transfer results: (a) LSV curve and Tafel slope (inset image) of the Co@NG-900 and NG-900 catalysts in phenol solution and (b) changes in the open-circuit potential curves of the NG-PMS complexes on the Co@NG-900 and NG-900 electrodes.147 ©2021, The Royal Society of Chemistry. (c) CVs curves (scan rate = 50 mV s−1) and (d) EIS Nyquist plots of different single-atom Co Fenton-like catalysts.82 ©2019, Elsevier B.V. All rights reserved.

In situ characterization is an indispensable means for the study of the micro mechanism. The addition of PMS and phenol to the solution at different times can lead to a significant increase in the current, indicating the electron transfer between the organic matter and the catalyst surface. Moreover, the measurement of the in situ open circuit potential of HSO5* on the surface of the Co@NG-900 and NG-900 catalysts clarified the formation of NG-PMS*. In the presence of PMS, it was obvious that the open circuit potentials Co@NG-900 and NG-900 in the two reaction solution systems increased immediately (Fig. 11b), which verified the generation of the NG-PMS* complex. It was observed that the potential change of Co@NG-900/PMS (0.64 V) was higher than that of NG-900/PMS (0.37 V), which further proved that Co@NG-900 has a higher inherent redox potential of NG-PMS*, which may be because the anchored isolated Co atoms enhance the interaction between PMS and NG. However, with the subsequent addition of phenol, the potential on the surface of Co@NG-900 and NG-900 catalyst first decreased rapidly to about +0.60 V, and then decreased slowly. These results demonstrate the electron transfer interaction between the catalyst surface and PMS/organic pollutants.

Similar to the above-mentioned catalytic system, the S-doped Co–S@NC catalyst/PMS system was applied for the degradation of dinotefuran (DIN). Similarly, the increase in Co–S@NC current shown by the CV curve proves that the catalyst has a higher current density and greater reduction ability to coordinate the redox process, and the combination of doped S and Co3O4 can form a metastable reaction complex to further optimize the interaction between PMS and the surface of Co–S@NC (Fig. 11c).82 In addition, Co–S@NC has a low charge transfer resistance to illustrate its good electron transfer path (Fig. 11d). Importantly, the researchers in this group simulated the tolerance of the Fenton-like catalyst to treat a real wastewater environment with the environment-friendly Co–S@NC catalyst, that is, the catalyst in wastewater solution containing different concentrations of inorganic anions (such as Cl, HCO3 and H2PO4) adsorbs active substances and generates free radicals to realize the mineralization of organic substances. As shown in Fig. 12a, the dose of 2 mM Cl inhibited the degradation of DIN and reduced the rate constant from 0.054 to 0.005 min−1. This can be explained by the fact that Cl quenches SO4˙ (k SO4˙ + Cl = 2.0 × 108 M−1 s−1) at a lower concentration to form chlorine-containing radicals (e.g., ˙Cl, ˙Cl2 and ˙ClOH) with an oxidation potential lower than SO4˙. Wang et al. showed the same Cl inhibition effect in their experiments on the activation of PMS to degrade BPA with single-atom dispersed Ag modified mesoporous graphitic carbon nitride (Ag/mpg-C3N4).148 However, they also found that when Cl exceeded a critical concentration, there was an acceleration effect due to the large amount of Cl2 and HOCl produced in the Ag/mpg-C3N4/PMS system.

 
SO4˙ + Cl → SO42− + Cl˙(19)
 
Cl + Cl˙ → Cl2˙(20)
 
Cl2˙ + H2O → ClOH˙ + H+ + Cl(21)
 
Cl + ˙OH → ClOH˙(22)
 
HSO5 + Cl → SO42− + HOCl(23)
 
HSO5 + 2Cl + H+ → SO42− + Cl2 + H2O(24)


image file: d2nr02989h-f12.tif
Fig. 12 Degradation results of DIN in Co–S@NC/PMS system with various coexisting anions: (a) Cl, (b) HCO3, (c) H2PO4 and (d) HA (catalyst dose = 0.1 g L−1, [DIN]0 = 10 mg L−1, [PMS]0 = 0.65 mM).82 ©2019, Elsevier B.V. All rights reserved.

The presence of HCO3 (10 mM) also inhibited the degradation of DIN (Fig. 12b). The main reason for this was that HCO3 reacts with SO4˙ and OH˙ to form HCO3˙ and CO3˙ (k SO4˙ + HCO3 = 2.8 × 106 M−1 s−1), which significantly reduce the oxidation capacity of DIN. In contrast to the inhibitory effects of Cl and HCO3, the addition of 10 mM H2PO4 to the reaction solution could greatly improve the degradation rate of (Fig. 12c). The asymmetric structure of PMS makes it vulnerable to attack by nucleophiles such as H2PO4 and rapidly decomposed, leading to the degradation of DIN by more reactive groups such as SO4˙ and OH˙. There are three reasons for the decrease in catalytic activity caused by HA in the Co–S@NC/PMS system, as follows: firstly, HA in the electron-rich part is equivalent to quencher, which can consume electrophilic SO4˙ and OH˙ radicals; secondly, HA is attached to the surface of the heterogeneous Co–S@NC, which limits the interaction between PMS and catalytic sites and interferes with the mass transfer process in the degradation process. Finally, the ineffective utilization of PMS is due to the reaction of hydroquinone, quinones and phenols in HA with PMS to form low-potential semiquinone free radicals.

5.2.3 High-valent metal-oxo species. Recently, the high-valence metal-oxo species-mediated Fenton-like catalytic pathway has been gradually developed and verified, and the most widely studied catalyst system is the Fe(III) reaction unit coordinated with the N-ligand to activate PMS and H2O2.149 For example, a series of Fe–Nx catalysts doped with isolated atomic iron was designed on the zeolite imidazolite skeleton (ZIF-8 and ZIF-67), dicyandiamide (DCD), 2-methylimidazole (2-MI) and nitrogen-rich biomass carbon precursor substrate via a chemical or high-energy ball milling method.150–152 In the Fenton-like/PMS reaction system, the formation of high-valent iron-oxo species and the degradation process of harmful organic substances were shown in Fig. 13a. Specifically, the original valence state of the Fe center (Fe(II) or Fe(III)) in the Fe–Nx complex usually determines the valence state of the generated heterogeneous high-valent iron-oxo species (FeIV[double bond, length as m-dash]O or FeV[double bond, length as m-dash]O).153–156 In the process of catalytic oxidation, the O atom originates from the cleavage of the O–O bond in PMS and coordinates with a single iron atom site in the Fe(III)–Nx coordination to form a metastable Fe(III)OOSO3 intermediate complex. Then, the intermediate complex Fe(III)OOSO3 has two evolution pathways, as follows: through (i) Fe(III)–O bond dissociation, returning to its original state and (ii) the heterolytic cleavage transition of the O–O bond in the metastable complex converted to [triple bond, length as m-dash]FeV[double bond, length as m-dash]O. Finally, the rapid decomposition of [triple bond, length as m-dash]FeV[double bond, length as m-dash]O into Fe(III) occurs between [triple bond, length as m-dash]FeV[double bond, length as m-dash]O and the adsorbed target organic matter through the “electron acceptor-donor complex” mechanism, accompanied by the mineralization of pollutants. Moreover, the evolution process of high-valence iron-oxo species can be clearly reflected by the following equation:123
 
[triple bond, length as m-dash] Fe(III) + HSO5 → [Fe(III)OOSO3]+ + H+(25)
 
[Fe(III)OOSO3]+[triple bond, length as m-dash] FeIV = O + SO4˙(26)
 
[triple bond, length as m-dash] FeIV = O + target pollution → products + [triple bond, length as m-dash] Fe(III)(27)

Moreover, the EPR experiment verified that the high-valent iron-oxo species is the main active center by uniquely using dimethyl sulfoxide (DMSO), which was completely different from the traditional free radical-based oxidation. Also, it can be oxidized by the high-valent iron-oxo species to form the corresponding dimethyl sulfone (DMSO2) via an oxygen-atom-transfer reaction. In addition, gas chromatography-mass spectrometry (GC-MS) can also confirm the change in DMSO and DMSO2 concentration during the oxidation process.157 For example, in Fig. 13b, the EPR experiment for the treatment of pharmaceutical contaminants with an Enteromorpha-derived Fe–N–C catalyst showed that the DMSO concentration was negatively correlated with the degradation rate of paracetamol (kobs decreased from 0.1194 min−1 to 0.0210 min−1, 20 mM DMSO).158 Furthermore, after acid treatment to remove the iron clusters, the single Fe atom in Fe–N–C was speculated to be closely related to the evolution of the high-valence FeIV[double bond, length as m-dash]O and FeV[double bond, length as m-dash]O species (Fig. 13c). Therefore, to further verify that the formation of high-valent iron-oxo species comes from the evolution of isolated iron atoms in Fe–Nx coordination compounds, Hu's team synthesized an Fe-g-C3N4 Fenton-like catalyst to illustrate the role of the Fe sites in the evolution of high-valent iron-oxo species, mainly through the addition of oxalate and SCN to the reaction system to form a metal carboxyl complex and to realize the inactivation of Fe centers.83Fig. 13d and e show that the increase in oxalate and NaSCN concentration significantly inhibited the quenching reaction and degradation of tetracycline (TC), indicating the strong contribution of iron sites to the formation of high valence metal sites and their participation in the oxidation reaction. Therefore, the nonradical pathways dominated through high-valent metal-oxo species are an effective way to mineralize pollutants and the direct activation of PMS via a single metal atom in the metal–Nx complex can be transformed into a key active nonradical species, namely high-valent metal-oxo species.


image file: d2nr02989h-f13.tif
Fig. 13 (a) Formation mechanism of high-valence iron oxygen species in PMS/CNF3/4-CP system.156 ©2018, the American Chemical Society. Using Enteromorpha-derived Fe–N–C as a catalyst, the effects of (b) different coexisting quenchers on the degradation performance and (c) reaction activity of the catalyst after thermal acid treatment and the quenching effect of DMSO (PMS: 0.5 mM; catalyst: 0.1 g L−1; pH: 6.0; paracetamol concentration: 10 mg L−1).83 ©2021, Elsevier B.V. All rights reserved. In SA Fe-g-C3N4 (600)/PMS system: (d) effect of oxalic acid and citrate coexisting in the reaction system on the decomposition of TC and (e) kobs of different pollutants by different catalysts (PMS: 0.5 mM; catalyst: 0.1 g L−1; pH: 6.5; TC concentration: 10 mg L−1).123 ©2020, Elsevier B.V. All rights reserved.

6. Summary and perspective

As reported in the literature, single-atom metal catalysts have many advantages, such as outstanding catalytic effect, great atomic utilization efficiency, unique coordination environment and excellent chemical stability. Thus, they have been recognized as ideal candidates with research value to inherit the advantages of traditional heterogeneous and homogeneous catalysts in many important applications. SACs have been developed and successfully applied in the field of sewage purification, especially Fenton-like single- or double-transition metal atomic reaction center catalysts in advanced oxidation processes, which have shown unprecedented chemical reaction activity and cycle stability. According to the Fenton-like SACs/PMS system reported in this review, the Fenton-like SACs realized by metal species such as Fe, Co, Cu, Mn, Mo, Cr, Ni, Ag, Pt, and Au alone or in combination as the redox reaction center have received special attention. Moreover, the active centers of isolated metal atoms synthesized via the “bottom-up” approach are usually coordinated and dispersed on carefully selected carriers in the form of covalent bonds or van der Waals forces. The strong interaction between the monatomic metal and carrier has the following advantages: (i) making charge transfer easier; (ii) preventing metal atoms from aggregating into clusters or particles; and (iii) compared with multi-metal sites, the single-metal reaction unit also has enhanced resistance to harmful side reactions. Notably, SACs possess unique electronic structures and energy level orbitals (different from traditional metal nanoparticles), which can maximize the improvement of metal dispersion and atomic utilization efficiency. Therefore, SACs are the most promising materials to realize the rational utilization of metal resources and atomic economy. Accordingly, developing reasonable strategies for the synthesis of high-quality Fenton-like SACs is significant for their practical application.

In this review, we summarized and showed the recent research progress of Fenton-like SACs used in advanced oxidation process, including the design and synthesis of metal atom reaction sites, the catalytic activity and stability, and the generation and mechanism of active species in the reaction process. Based on this, a basic criterion for Fenton-like SACs in which isolated metal atoms act as active unit and participate in the reaction is provided to explore the relationship between the synthesis principle of catalysts, the coordination configuration of active centers, chemical kinetic constants and catalytic mechanism. Among them, the coordination environment is closely related to the carrier of anchoring and dispersing metal atoms. Herein, we classified them into three main types, as follows: C-coordinated metal (M–C), heteroatom (e.g., N, S, P, and B)-doped carbon material-fixed metal, and semiconductor-dispersed metal atoms. Similarly, the performance of the support itself will also affect the overall performance of Fenton-like SACs. The elaboration of the catalytic mechanism mainly relies on radical capture experiments, electron spin resonance spectroscopy and DFT theoretical calculation. According to the properties of the dominant active species, the reaction mechanism can be divided into two categories, i.e., radical and nonradical pathways (singlet oxygen, electron-transfer regime, and high-valent metal-oxo species). Although much work and attention have been devoted to the research on Fenton-like SACs, there are still considerable challenges to be solved and opportunities for this powerful Fenton-like SACs/PMS system to truly achieve a wide range of practical applications.

(I) Applicability of Fenton-like SACs preparation: nowadays, the most widely used processes for the synthesis of Fenton-like SACs are the hydrothermal method and high-temperature pyrolysis under the protection of an inert atmosphere (N2 or Ar). However, these preparation methods are not feasible for “real-world” applications. On one hand, the thermal movement of metal atoms is more intense at high temperature, and thus the probability of their migration and aggregation into clusters or nanoparticles is greater. This is because the surface energy of isolated metal atoms is higher, and they tend to agglomerate to enhance their stability. On the other hand, they are not applicable to meet the demand of mass production, especially the preparation requirements of Fenton-like SACs above the gram level. Furthermore, the production energy consumption and complex preparation process have become “stumbling blocks” that hinder the practical application of these catalysts. At present, the development of synthetic methods for Fenton-like SACs, especially simple methods that can be employed on an industrial scale, is still lacking. Therefore, significant time and energy need to be invested to explore methods that can realize the large-scale preparation of catalysts (g level or even kg level). To solve the problem of the agglomeration of metal atoms with high surface energy, the usual solution is to reduce the loading of metal atoms, which can realize the preparation of isolated and evenly dispersed catalysts. However, a low loading of metal atoms corresponds to a low density of active sites that can participate in the redox process, which is the main reason for their poor catalytic activity and slow chemical reaction kinetics. In addition, low active centers will also bring potential side reactions in the reaction process, that is, the reaction intermediate of the catalytic process will be accumulated due to the low loading of monatomic sites. Therefore, the low loading of metal atoms is the bottleneck limiting the practical application of catalysts, while a high loading of metal atoms tends to migrate and agglomerate into nanoclusters or even nanoparticles. In summary, it is urgent to develop Fenton-like SACs with stable high-loaded metal atom sites for industrial application. In addition, the design of the single-atom coordination environment in Fenton-like SAC/PMS systems is mostly limited to the M–Nx and M–Nx–C forms, while there are few studies on capturing and anchoring metal atoms using other heteroatoms, such as S, O, P and B. The change in the coordination environment by introducing new heteroatoms may also provide a novel way to regulate the electronic structure of the active sites, resulting in unexpected effects on the catalytic activity.

(II) Surface morphology regulation of Fenton-like SACs: as is known, the number of active sites that can participate in the catalytic process and the intrinsic reaction activity of Fenton-like SACs are the two main factors closely related to the performance of these catalysts. The accessible active reaction center and the prevention of metal atom agglomeration are directly determined and affected by the surface morphology of the catalyst. At present, there are many methods to control the surface morphology of catalysts, such as the template method (soft template method and hard template method). Various morphologies, such as 1D, 2D, 3D and porous morphologies, have been successfully developed and applied to the Fenton-like oxidation process in advanced oxidation processes. Taking platinum-based alloy nanowires as an example, the rapid development of one-dimensional materials can be attributed to their inherent anisotropic structure, excellent specific surface area and outstanding electron conduction ability compared with zero-dimensional nanoparticles. The advantage of 2D catalyst materials is that they can alleviate the aggregation of metal species in the process of high-temperature pyrolysis because the migration distance between metal species is far enough. In the case of porous materials, MOFs are a representative class because of their large surface area and controllable porosity. The most widely studied MOFs in Fenton-like systems belong to the sub-class of zeolitic imidazolate frameworks (ZIFs), which are rich in pore channels and coordination-anchored nitrogen atoms. Their unique porous structure is conducive to the exposure of active sites, and also ensures rapid mass transfer, making the target reaction and the accessibility between reaction units. However, although these materials have been widely studied and achieved considerable catalytic activity, their catalytic performance is still not enough to achieve their industrial-scale application. This is because the pore structure in the catalyst is dominated by micropores, which cannot make full use of its active sites. Therefore, there is an urgent need to design materials with macroporous channel structures, enabling the internal reaction unit of the catalyst to effectively participate in the redox reaction and be effectively utilized. For example, a medium/large pore structure with an average pore size of greater than 2 nm can effectively improve the mass transfer capacity and the utilization of internal active centers. In conclusion, the regulation of the morphology of catalyst materials, especially making full use of the advantages of various morphologies and developing mesoporous materials with accessible internal active centers, is of great significance to improve the stability of catalysts.

(III) Local structure identification and theoretical simulation of Fenton-like SACs: thus far, the characterization and confirmation of the coordination environment of the designed and synthesized monatomic active centers and the valence states of metal atoms still depend on extended X-ray absorption fine structure (EXAFS) measurements, X-ray absorption fine structure spectra (XAFS), and near edge X-ray absorption fine structure (NEXAFS, S L-edge). The XAS of these synchrotron radiation sources can indeed detect and successfully characterize the coordination environment of single metal atoms. However, this technique also has some limitations, that is, it has certain bulk sensitivity and can only provide average information on the bulk phase. Furthermore, for real catalysts, their XAS characterization may be disturbed by the coordination configuration of their different metal atoms, and thus this characterization method results in some distortion to the detected bulk catalyst. Therefore, there is an urgent need to develop more precise characterization methods for dynamic information and complementary verification. In addition, DFT theoretical calculation is an indispensable technical tool to obtain the chemical structure of Fenton-like SACs, exploring the catalytic active center playing a leading role and better describing the properties and reaction mechanism of SACs. Moreover, DFT theoretical calculation can be employed to obtain the reaction free energy of each basic step of a specific reaction, calculate the adsorption energy between the catalyst surface active site and the activated PMS oxidant or intermediate in the Fenton-like reaction process, and establish the relationship between structure and activity at the atomic level. However, not all the catalytic mechanisms obtained by DFT correspond to the experimental conclusions. Therefore, in the process of discussing and verifying the catalytic mechanism, it is necessary to build a more precise theoretical calculation model to obtain experimental evidence between the reaction unit and the kinetic structure in the process of catalytic activation.

(IV) Simulation of catalytic mechanism: in Fenton-like SACs/PMS advanced oxidation systems, the discussion on the catalytic mechanism mainly focuses on: (i) the coordination configuration and electronic structure of the reaction center involved in the adsorption and activation of PMS (transition metal species-electronegative non-metallic element coordination active centers, metal free sites around isolated metal atoms, and various defects); (ii) the reaction unit will spontaneously induce multiple oxidation paths, leading to the generation of radicals (ROS), non-radicals and electrons for the decomposition and mineralization of refractory organics; (iii) the main contribution or synergy of different active species to the decomposition process of organic matter. At present, the catalytic centers involved in the reaction and the active species activated in the degradation pathway are mainly characterized by EXAFS, XAFS, electron paramagnetic resonance (EPR) and free radical capture experiments. In addition, in situ technology is an important means to characterize the mechanism, such as in situ Raman testing technology and in situ Fourier infrared testing technology. However, the application of these in situ technologies is relatively limited. A more in-depth simulation and proof of the above-mentioned characterization come from the support of DFT theoretical calculation, which can calculate the adsorption energy and kinetic energy barrier of active species, intermediates and final products in the reaction center. Unfortunately, due to the limitation of algorithm resources of traditional DFT theory, it cannot calculate and screen a large number of models in a short time. Therefore, the rapid development of computing systems, software and hardware is necessary to establish the structure-performance relationship among the characteristics of catalyst active centers, the formation path of active species and catalytic performance. These advanced calculation methods can enable us to quickly simulate and predict the catalytic decomposition pathway and performance under near real reaction conditions, and provide new insights to understand the evolution of active species and the kinetic and thermodynamic characteristics of organic oxidation on Fenton-like SACs from the molecular to macro level.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No. 5197105852071072), the Fundamental Research Funds for the Central Universities (No. N182312007 and N2023001).

References

  1. H. J. H. Fenton, J. Chem. Soc., Trans., 1894, 65, 899–910 RSC .
  2. M. Xiang, M. Huang, H. Li, W. Wang, Y. Huang, Z. Lu, C. Wang, R. Si and W. Cao, Chem. Eng. J., 2021, 417, 129208 CrossRef CAS .
  3. J. Ding, H. J. Khan, G. V. Sarrigani, P. Fitzgerald, A. E. Ghadi, O. Lefebvre, C. Meng, M. H. Z. M. Harun, Y. Lu, A. Abbas, A. Montoya, D. Wiley and D. K. Wang, Chem. Eng. J., 2022, 432, 134435 CrossRef CAS .
  4. C. Yang, M. Zhou, C. He, Y. Gao, S. Li, X. Fan, Y. Lin, F. Cheng, P. Zhu and C. Cheng, Nano-Micro Lett., 2019, 11, 87 CrossRef CAS PubMed .
  5. M. Moztahida, M. Nawaz, J. Kim, A. Shahzad, S. Kim, J. Jang and D. S. Lee, Chem. Eng. J., 2019, 370, 855–865 CrossRef CAS .
  6. Y. Sun, Z. Yang, P. Tian, Y. Sheng, J. Xu and Y.-F. Han, Appl. Catal., B, 2019, 244, 1–10 CrossRef CAS .
  7. Y. Wang, H. Zhao and G. Zhao, Appl. Catal., B, 2015, 164, 396–406 CrossRef CAS .
  8. X. Li, Z. Ao, J. Liu, H. Sun, A. I. Rykov and J. Wang, ACS Nano, 2016, 10, 11532–11540 CrossRef CAS PubMed .
  9. N. Zhang, Y. Yi, J. Lian and Z. Fang, Chem. Eng. J., 2020, 395, 124897 CrossRef CAS .
  10. H. Xiang, W. Feng and Y. Chen, Adv. Mater., 2020, 32, 1905994 CrossRef CAS PubMed .
  11. M. B. Gawande, K. Ariga and Y. Yamauchi, Small, 2021, 17, 2101584 CrossRef CAS PubMed .
  12. X. Cui, W. Li, P. Ryabchuk, K. Junge and M. Beller, Nat. Catal., 2018, 1, 385–397 CrossRef CAS .
  13. Z. Kou, W. Zang, P. Wang, X. Li and J. Wang, Nanoscale Horiz., 2020, 5, 757–764 RSC .
  14. Y. Liu, S. Liu, Y. Wang, Q. Zhang, L. Gu, S. Zhao, D. Xu, Y. Li, J. Bao and Z. Dai, J. Am. Chem. Soc., 2018, 140, 2731–2734 CrossRef CAS PubMed .
  15. S. Mitchell, E. Vorobyeva and J. Pérez-Ramírez, Angew. Chem., Int. Ed., 2018, 57, 15316–15329 CrossRef CAS PubMed .
  16. Z. Liang, L. Yin, H. Yin, Z. Yin and Y. Du, Nanoscale Horiz., 2022, 7, 31–40 RSC .
  17. S. Tong, B. Fu, L. Gan and Z. Zhang, J. Mater. Chem. A, 2021, 9, 10731–10738 RSC .
  18. M.-M. Millet, G. Algara-Siller, S. Wrabetz, A. Mazheika, F. Girgsdies, D. Teschner, F. Seitz, A. Tarasov, S. V. Levchenko, R. Schlögl and E. Frei, J. Am. Chem. Soc., 2019, 141, 2451–2461 CrossRef CAS PubMed .
  19. X. Xie, L. Shang, X. Xiong, R. Shi and T. Zhang, Adv. Energy Mater., 2022, 12, 2102688 CrossRef CAS .
  20. G. Wan, P. Yu, H. Chen, J. Wen, C. Sun, H. Zhou, N. Zhang, Q. Li, W. Zhao, B. Xie, T. Li and J. Shi, Small, 2018, 14, 1704319 CrossRef PubMed .
  21. Y. Liu, G. Zhu, A. Li, J. Pei, Y. Zheng, W. Chen, J. Ding, W. Wu, T. Wang, D. Wang and J. Mao, Nano Energy, 2021, 83, 105799 CrossRef CAS .
  22. Y. Shang, X. Xu, B. Gao, S. Wang and X. Duan, Chem. Soc. Rev., 2021, 50, 5281–5322 RSC .
  23. H. Yang, Z. Li, S. Kou, G. Lu and Z. Liu, Appl. Catal., B, 2020, 278, 119270 CrossRef CAS .
  24. Y. Wang, P. Zhang, T. Li, L. Lyu, Y. Gao and C. Hu, J. Hazard. Mater., 2021, 408, 124818 CrossRef CAS PubMed .
  25. Y. Gao, X. Duan, B. Li, Q. Jia, Y. Li, X. Fan, F. Zhang, G. Zhang, S. Wang and W. Peng, J. Mater. Chem. A, 2021, 9, 14793–14805 RSC .
  26. Z.-Y. Li, L. Wang, Y.-L. Liu, P.-N. He, X. Zhang, J. Chen, H.-T. Gu, H.-C. Zhang and J. Ma, Water Res., 2021, 195, 116973 CrossRef CAS PubMed .
  27. H. Gao, H. Yang, J. Xu, S. Zhang and J. Li, Small, 2018, 14, 1801353 CrossRef PubMed .
  28. E.-J. Kim, S. Park, S. Adil, S. Lee and K. Cho, Environ. Sci. Technol., 2021, 55, 5301–5311 CrossRef CAS PubMed .
  29. F. Yang, B. Wang, H. Su, S. Zhou and Y. Kong, Mater. Chem. Front., 2017, 1, 2065–2077 RSC .
  30. W. Guo, Z. Wang, X. Wang and Y. Wu, Adv. Mater., 2021, 33, 2004287 CrossRef CAS PubMed .
  31. S. K. Kaiser, Z. Chen, D. F. Akl, S. Mitchell and J. Pérez-Ramírez, Chem. Rev., 2020, 120, 11703–11809 CrossRef CAS PubMed .
  32. Y. Pan, S. Liu, K. Sun, X. Chen, B. Wang, K. Wu, X. Cao, W.-C. Cheong, R. Shen, A. Han, Z. Chen, L. Zheng, J. Luo, Y. Lin, Y. Liu, D. Wang, Q. Peng, Q. Zhang, C. Chen and Y. Li, Angew. Chem., Int. Ed., 2018, 57, 8614–8618 CrossRef CAS PubMed .
  33. B. C. Gates, Chem. Rev., 1995, 95, 511–522 CrossRef CAS .
  34. Q. Shi, T. Yu, R. Wu and J. Liu, ACS Appl. Mater. Interfaces, 2021, 13, 60815–60836 CrossRef CAS PubMed .
  35. J. D. Aiken and R. G. Finke, J. Mol. Catal. A: Chem., 1999, 145, 1–44 CrossRef CAS .
  36. X. Cui, L. Gao, S. Lei, S. Liang, J. Zhang, C. D. Sewell, W. Xue, Q. Liu, Z. Lin and Y. Yang, Adv. Funct. Mater., 2021, 31, 2009197 CrossRef CAS .
  37. S. Liu, Q. Meyer, Y. Li, T. Zhao, Z. Su, K. Ching and C. Zhao, Chem. Commun., 2022, 58, 2323–2326 RSC .
  38. X. Li, X. Yang, L. Liu, H. Zhao, Y. Li, H. Zhu, Y. Chen, S. Guo, Y. Liu, Q. Tan and G. Wu, ACS Catal., 2021, 11, 7450–7459 CrossRef CAS .
  39. M. G. Rinaudo, A. M. Beltrán, M. A. Fernández, L. E. Cadús and M. R. Morales, Mater. Today Chem., 2020, 17, 100340 CrossRef CAS .
  40. H. Yan, Y. Lin, H. Wu, W. Zhang, Z. Sun, H. Cheng, W. Liu, C. Wang, J. Li, X. Huang, T. Yao, J. Yang, S. Wei and J. Lu, Nat. Commun., 2017, 8, 1070 CrossRef PubMed .
  41. J. Y. Jung, J.-H. Jang, J.-G. Kim, K.-S. Lee, H.-K. Lim, P. Kim, R. P. H. Chang, J.-W. Park, S. J. Yoo and N. D. Kim, Small Methods, 2021, 5, 2100239 CrossRef CAS PubMed .
  42. Q. Wang, Z. Zhang, C. Cai, M. Wang, Z. L. Zhao, M. Li, X. Huang, S. Han, H. Zhou, Z. Feng, L. Li, J. Li, H. Xu, J. S. Francisco and M. Gu, J. Am. Chem. Soc., 2021, 143, 13605–13615 CrossRef CAS PubMed .
  43. Y. Chen, Z. Huang, X. Gu, Z. Ma, J. Chen and X. Tang, Chin. J. Catal., 2017, 38, 1588–1596 CrossRef CAS .
  44. L. Xing, Y. Jin, Y. Weng, R. Feng, Y. Ji, H. Gao, X. Chen, X. Zhang, D. Jia and G. Wang, Matter, 2022, 5, 788–807 CrossRef CAS .
  45. J. Liu, ACS Catal., 2017, 7, 34–59 CrossRef CAS .
  46. C. Zhu, S. Fu, J. Song, Q. Shi, D. Su, M. H. Engelhard, X. Li, D. Xiao, D. Li, L. Estevez, D. Du and Y. Lin, Small, 2017, 13, 1603407 CrossRef PubMed .
  47. D. A. Bulushev, M. Zacharska, A. S. Lisitsyn, O. Y. Podyacheva, F. S. Hage, Q. M. Ramasse, U. Bangert and L. G. Bulusheva, ACS Catal., 2016, 6, 3442–3451 CrossRef CAS .
  48. A. Bakandritsos, R. G. Kadam, P. Kumar, G. Zoppellaro, M. Medved, J. Tuček, T. Montini, O. Tomanec, P. Andrýsková, B. Drahoš, R. S. Varma, M. Otyepka, M. B. Gawande, P. Fornasiero and R. Zbořil, Adv. Mater., 2019, 31, 1900323 CrossRef PubMed .
  49. X. Ge, G. Su, W. Che, J. Yang, X. Zhou, Z. Wang, Y. Qu, T. Yao, W. Liu and Y. Wu, ACS Catal., 2020, 10, 10468–10475 CrossRef CAS .
  50. C. Rivera-Cárcamo and P. Serp, ChemCatChem, 2018, 10, 5058–5091 CrossRef .
  51. Y. Peng, B. Lu and S. Chen, Adv. Mater., 2018, 30, 1801995 CrossRef PubMed .
  52. J. Jones, H. Xiong, A. T. DeLaRiva, E. J. Peterson, H. Pham, S. R. Challa, G. Qi, S. Oh, M. H. Wiebenga, X. I. P. Hernández, Y. Wang and A. K. Datye, Science, 2016, 353, 150–154 CrossRef CAS PubMed .
  53. Y. Li, T. Yang, S. Qiu, W. Lin, J. Yan, S. Fan and Q. Zhou, Chem. Eng. J., 2020, 389, 124382 CrossRef CAS .
  54. X. Dong, Z. Chen, A. Tang, D. D. Dionysiou and H. Yang, Adv. Funct. Mater., 2022, 32, 2111565 CrossRef CAS .
  55. J. Li, S. Zhao, L. Zhang, S. P. Jiang, S.-Z. Yang, S. Wang, H. Sun, B. Johannessen and S. Liu, Small, 2021, 17, 2004579 CrossRef CAS PubMed .
  56. B. Zhang, C. Pan, H. Liu, X. Wu, H. Jiang, L. Yang, Z. Qi, G. Li, L. Shan, Y. Lin, L. Song and Y. Jiang, Chem. Eng. J., 2022, 439, 135768 CrossRef CAS .
  57. Y. Gu, T. Xu, X. Chen, W. Chen and W. Lu, Chem. Eng. J., 2022, 427, 131973 CrossRef CAS .
  58. M. Zhang, Y. Liu, H. Zhao, J. Tao, N. Geng, W. Li and Y. Zhai, ACS Appl. Mater. Interfaces, 2021, 13, 19904–19914 CrossRef CAS PubMed .
  59. W. Zhong, G. Zhang, Y. Zhang, C. Jia, T. Yang, S. Ji, O. V. Prezhdo, J. Yuan, Y. Luo and J. Jiang, J. Phys. Chem. Lett., 2019, 10, 7009–7014 CrossRef CAS PubMed .
  60. J. Xin, F. Li, Z. Li, J. Zhao and Y. Wang, Inorg. Chem. Front., 2022, 9, 302–309 RSC .
  61. X. Zhang, L. Shang, Z. Yang and T. Zhang, ChemPlusChem, 2021, 86, 1635–1639 CrossRef CAS PubMed .
  62. Z. Zhao, W. Zhou, D. Lin, L. Zhu, B. Xing and Z. Liu, Appl. Catal., B, 2022, 309, 121256 CrossRef CAS .
  63. S. Zuo, X. Jin, X. Wang, Y. Lu, Q. Zhu, J. Wang, W. Liu, Y. Du and J. Wang, Appl. Catal., B, 2021, 282, 119551 CrossRef CAS .
  64. F. Chen, X.-L. Wu, C. Shi, H. Lin, J. Chen, Y. Shi, S. Wang and X. Duan, Adv. Funct. Mater., 2021, 31, 2007877 CrossRef CAS .
  65. G. Liao, X. Qing, P. Xu, L. Li, P. Lu, W. Chen and D. Xia, Chem. Eng. J., 2022, 427, 132027 CrossRef CAS .
  66. Y. Ha, B. Fei, X. Yan, H. Xu, Z. Chen, L. Shi, M. Fu, W. Xu and R. Wu, Adv. Energy Mater., 2020, 10, 2002592 CrossRef CAS .
  67. T. Shen, X. Huang, S. Xi, W. Li, S. Sun and Y. Hou, J. Energy Chem., 2022, 68, 184–194 CrossRef .
  68. J. N. Tiwari, A. M. Harzandi, M. Ha, S. Sultan, C. W. Myung, H. J. Park, D. Y. Kim, P. Thangavel, A. N. Singh, P. Sharma, S. S. Chandrasekaran, F. Salehnia, J.-W. Jang, H. S. Shin, Z. Lee and K. S. Kim, Adv. Energy Mater., 2019, 9, 1900931 CrossRef .
  69. P. Yin, T. Yao, Y. Wu, L. Zheng, Y. Lin, W. Liu, H. Ju, J. Zhu, X. Hong, Z. Deng, G. Zhou, S. Wei and Y. Li, Angew. Chem., Int. Ed., 2016, 55, 10800–10805 CrossRef CAS PubMed .
  70. Q. Wang and D. Astruc, Chem. Rev., 2020, 120, 1438–1511 CrossRef CAS PubMed .
  71. J. Fonseca, T. Gong, L. Jiao and H.-L. Jiang, J. Mater. Chem. A, 2021, 9, 10562–10611 RSC .
  72. T. Kitao, Y. Zhang, S. Kitagawa, B. Wang and T. Uemura, Chem. Soc. Rev., 2017, 46, 3108–3133 RSC .
  73. X. Li, X. Huang, S. Xi, S. Miao, J. Ding, W. Cai, S. Liu, X. Yang, H. Yang, J. Gao, J. Wang, Y. Huang, T. Zhang and B. Liu, J. Am. Chem. Soc., 2018, 140(39), 12469–12475 CrossRef CAS PubMed .
  74. Y. Gao, C. Yang, M. Zhou, C. He, S. Cao, Y. Long, S. Li, Y. Lin, P. Zhu and C. Cheng, Small, 2020, 16, 2005060 CrossRef CAS PubMed .
  75. Y. Chen, H. Li, W. Zhao, W. Zhang, J. Li, W. Li, X. Zheng, W. Yan, W. Zhang, J. Zhu, R. Si and J. Zeng, Nat. Commun., 2019, 10, 1885 CrossRef PubMed .
  76. H. Zhang, G. Liu, L. Shi and J. Ye, Adv. Energy Mater., 2018, 8, 1701343 CrossRef .
  77. Q. Chen, Y. Liu, Y. Lu, Y. Hou, X. Zhang, W. Shi and Y. Huang, J. Hazard. Mater., 2022, 422, 126929 CrossRef CAS PubMed .
  78. J. Tang and J. Wang, Environ. Sci. Technol., 2018, 52, 5367–5377 CrossRef CAS PubMed .
  79. Y. Zhu, W. Sun, W. Chen, T. Cao, Y. Xiong, J. Luo, J. Dong, L. Zheng, J. Zhang, X. Wang, C. Chen, Q. Peng, D. Wang and Y. Li, Adv. Funct. Mater., 2018, 28, 1802167 CrossRef .
  80. W. Zhao, L.-P. Xiao, G. Song, R.-C. Sun, L. He, S. Singh, B. A. Simmons and G. Cheng, Green Chem., 2017, 19, 3272–3281 RSC .
  81. Y. Qi, J. Li, Y. Zhang, Q. Cao, Y. Si, Z. Wu, M. Akram and X. Xu, Appl. Catal., B, 2021, 286, 119910 CrossRef CAS .
  82. W. Du, Q. Zhang, Y. Shang, W. Wang, Q. Li, Q. Yue, B. Gao and X. Xu, Appl. Catal., B, 2020, 262, 118302 CrossRef CAS .
  83. L. Peng, X. Duan, Y. Shang, B. Gao and X. Xu, Appl. Catal., B, 2021, 287, 119963 CrossRef CAS .
  84. J. Yang, D. Zeng, J. Li, L. Dong, W.-J. Ong and Y. He, Chem. Eng. J., 2021, 404, 126376 CrossRef CAS .
  85. X. An, Q. Tang, H. Lan, H. Liu and J. Qu, Appl. Catal., B, 2019, 244, 407–413 CrossRef CAS .
  86. T. Zhang, L. Zhang and Y. Hou, eScience, 2022, 2, 164–182 CrossRef .
  87. J. Tang, X. Peng, T. Lin, X. Huang, B. Luo and L. Wang, eScience, 2021, 1, 203–211 CrossRef .
  88. H. Song, R. Du, Y. Wang, D. Zu, R. Zhou, Y. Cai, F. Wang, Z. Li, Y. Shen and C. Li, Appl. Catal., B, 2021, 286, 119898 CrossRef CAS .
  89. Y. Wei, J. Gao, G. Xing, G. Li, P. Dang, S. Liang, Y. S. Huang, C. C. Lin, T.-S. Chan and J. Lin, Inorg. Chem., 2019, 58, 6376–6387 CrossRef CAS PubMed .
  90. M. Xie, F. Dai, J. Li, X. Dang, J. Guo, W. Lv, Z. Zhang and X. Lu, Angew. Chem., Int. Ed., 2021, 60, 14370–14375 CrossRef CAS PubMed .
  91. T. Zhang and L. Fu, Chem, 2018, 4, 671–689 CAS .
  92. N. D. Boscher, M. Wang, A. Perrotta, K. Heinze, M. Creatore and K. K. Gleason, Adv. Mater., 2016, 28, 7479–7485 CrossRef CAS PubMed .
  93. Z. Wang, Y. Shen, Y. Ito, Y. Zhang, J. Du, T. Fujita, A. Hirata, Z. Tang and M. Chen, ACS Nano, 2018, 12, 1571–1579 CrossRef CAS PubMed .
  94. X. Li, L. Colombo and R. S. Ruoff, Adv. Mater., 2016, 28, 6247–6252 CrossRef CAS PubMed .
  95. Z. Cai, B. Liu, X. Zou and H.-M. Cheng, Chem. Rev., 2018, 118, 6091–6133 CrossRef CAS PubMed .
  96. D. Grasseschi, W. C. Silva, R. de Souza Paiva, L. D. Starke and A. S. do Nascimento, Coord. Chem. Rev., 2020, 422, 213469 CrossRef CAS .
  97. H. Valencia, A. Gil and G. Frapper, J. Phys. Chem. C, 2010, 114, 14141–14153 CrossRef CAS .
  98. D. Deng, X. Chen, L. Yu, X. Wu, Q. Liu, Y. Liu, H. Yang, H. Tian, Y. Hu, P. Du, R. Si, J. Wang, X. Cui, H. Li, J. Xiao, T. Xu, J. Deng, F. Yang, P. N. Duchesne, P. Zhang, J. Zhou, L. Sun, J. Li, X. Pan and X. Bao, Sci. Adv., 2015, 1, e1500462 CrossRef PubMed .
  99. J. Miao, Y. Zhu, J. Lang, J. Zhang, S. Cheng, B. Zhou, L. Zhang, P. J. J. Alvarez and M. Long, ACS Catal., 2021, 11, 9569–9577 CrossRef CAS .
  100. J.-B. Ma, L.-L. Xu, Q.-Y. Liu and S.-G. He, Angew. Chem., Int. Ed., 2016, 55, 4947–4951 CrossRef CAS PubMed .
  101. J. Köhler, J.-H. Chang and M.-H. Whangbo, J. Am. Chem. Soc., 2005, 127, 2277–2284 CrossRef PubMed .
  102. J. P. Dürholt, B. F. Jahromi and R. Schmid, ACS Cent. Sci., 2019, 5, 1440–1448 CrossRef PubMed .
  103. S. Chen, Z. Kang, X. Hu, X. Zhang, H. Wang, J. Xie, X. Zheng, W. Yan, B. Pan and Y. Xie, Adv. Mater., 2017, 29, 1701687 CrossRef PubMed .
  104. P. Duan, Y. Qi, S. Feng, X. Peng, W. Wang, Y. Yue, Y. Shang, Y. Li, B. Gao and X. Xu, Appl. Catal., B, 2020, 267, 118717 CrossRef CAS .
  105. T. Sun, S. Mitchell, J. Li, P. Lyu, X. Wu, J. Pérez-Ramírez and J. Lu, Adv. Mater., 2021, 33, 2003075 CrossRef CAS PubMed .
  106. J. Ramler and C. Lichtenberg, Dalton Trans., 2021, 50, 7120–7138 RSC .
  107. M. Seitz, A. G. Oliver and K. N. Raymond, J. Am. Chem. Soc., 2007, 129, 11153–11160 CrossRef CAS PubMed .
  108. Y. Wang, Y. Liu, W. Liu, J. Wu, Q. Li, Q. Feng, Z. Chen, X. Xiong, D. Wang and Y. Lei, Energy Environ. Sci., 2020, 13, 4609–4624 RSC .
  109. Y. Xu, M. Chu, F. Liu, X. Wang, Y. Liu, M. Cao, J. Gong, J. Luo, H. Lin, Y. Li and Q. Zhang, Nano Lett., 2020, 20, 6865–6872 CrossRef CAS PubMed .
  110. X. Zhou, M. Ke, G. Huang, C. Chen, W. Chen, K. Liang, Y. Qu, J. Yang, Y. Wang, F. Lia, H. Yu and Y. Wu, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2119492119 CrossRef CAS PubMed .
  111. C. Zhang, S. Qin, B. Li and P. Jin, J. Mater. Chem. A, 2022, 10, 8309–8323 RSC .
  112. X.-M. Hu, H. H. Hval, E. T. Bjerglund, K. J. Dalgaard, M. R. Madsen, M.-M. Pohl, E. Welter, P. Lamagni, K. B. Buhl, M. Bremholm, M. Beller, S. U. Pedersen, T. Skrydstrup and K. Daasbjerg, ACS Catal., 2018, 8, 6255–6264 CrossRef CAS .
  113. R. Qin, K. Liu, Q. Wu and N. Zheng, Chem. Rev., 2020, 120, 11810–11899 CrossRef CAS PubMed .
  114. Y. Ren, Y. Tang, L. Zhang, X. Liu, L. Li, S. Miao, D. S. Su, A. Wang, J. Li and T. Zhang, Nat. Commun., 2019, 10, 4500 CrossRef PubMed .
  115. Y. Gu, B. Xi, W. Tian, H. Zhang, Q. Fu and S. Xiong, Adv. Mater., 2021, 33, 2100429 CrossRef CAS PubMed .
  116. B. Liu, J. Xie, H. Ma, X. Zhang, Y. Pan, J. Lv, H. Ge, N. Ren, H. Su, X. Xie, L. Huang and W. Huang, Small, 2017, 13, 1601001 CrossRef PubMed .
  117. E. Yazici, S. Yanik and M. B. Yilmaz, Carbon, 2017, 111, 822–827 CrossRef CAS .
  118. A.-B. Bornhof, M. Vázquez-Nakagawa, L. Rodríguez-Pérez, María Ángeles Herranz, N. Sakai, N. Martín, S. Matile and J. López-Andarias, Angew. Chem., Int. Ed., 2019, 58, 16097–16100 CrossRef CAS PubMed .
  119. M. Dasbach, M. Pyschik, V. Lehmann, K. Parey, D. Rhinow, H. M. Reinhardt and N. A. Hampp, ACS Nano, 2020, 14, 8181–8190 CrossRef CAS PubMed .
  120. H. Zhang, L. Lyu, Q. Fang, C. Hu, S. Zhan and T. Li, Appl. Catal., B, 2021, 286, 119912 CrossRef CAS .
  121. Q. Wu, J. Wang, Z. Wang, Y. Xu, Z. Xing, X. Zhang, Y. Guan, G. Liao and X. Li, J. Mater. Chem. A, 2020, 8, 13685–13693 RSC .
  122. J. Yang, D. Zeng, Q. Zhang, R. Cui, M. Hassan, L. Dong, J. Li and Y. He, Appl. Catal., B, 2020, 279, 119363 CrossRef CAS .
  123. X. Peng, J. Wu, Z. Zhao, X. Wang, H. Dai, L. Xu, G. Xu, Y. Jian and F. Hu, Chem. Eng. J., 2022, 427, 130803 CrossRef CAS .
  124. O. Tutusaus, C. Ni and N. K. Szymczak, J. Am. Chem. Soc., 2013, 135, 3403–3406 CrossRef CAS PubMed .
  125. H. Zhang, W. Tian, X. Duan, H. Sun, S. Liu and S. Wang, Adv. Mater., 2020, 32, 1904037 CrossRef CAS PubMed .
  126. J. Hu, W. Liu, C. Xin, J. Guo, X. Cheng, J. Wei, C. Hao, G. Zhang and Y. Shi, J. Mater. Chem. A, 2021, 9, 24803–24829 RSC .
  127. M. B. Gawande, P. Fornasiero and R. Zbořil, ACS Catal., 2020, 10, 2231–2259 CrossRef CAS .
  128. T. Wang, X. Cao and L. Jiao, Small, 2021, 17, 2004398 CrossRef CAS PubMed .
  129. C. Du, Y. Gao, J. Wang and W. Chen, J. Mater. Chem. A, 2020, 8, 9981–9990 RSC .
  130. A. Wang, Z. Zheng, H. Wang, Y. Chen, C. Luo, D. Liang, B. Hu, R. Qiu and K. Yan, Appl. Catal., B, 2020, 277, 119171 CrossRef CAS .
  131. X. Lai, X. Ning, J. Chen, Y. Li, Y. Zhang and Y. Yuan, J. Hazard. Mater., 2020, 398, 122826 CrossRef CAS PubMed .
  132. S. Liu, Z. Zhang, F. Huang, Y. Liu, L. Feng, J. Jiang, L. Zhang, F. Qi and C. Liu, Appl. Catal., B, 2021, 286, 119921 CrossRef CAS .
  133. X. Liu, T. Liu, H. Zhou, W. Hu, B. Yao, J. Li, Y. Yang, D. Zhi, S. Du and Y. Zhou, Sci. Total Environ., 2021, 785, 147152 CrossRef CAS PubMed .
  134. Y. Sun, R. Li, C. Song, H. Zhang, Y. Cheng, A. Nie, H. Li, D. D. Dionysiou, J. Qian and B. Pan, Appl. Catal., B, 2021, 298, 120537 CrossRef CAS .
  135. J.-C. E. Yang, M.-P. Zhu, X. Duan, S. Wang, B. Yuan and M.-L. Fu, Appl. Catal., B, 2021, 297, 120460 CrossRef CAS .
  136. X. Cheng, Y. Lu, L. Zheng, M. Pupucevski, H. Li, G. Chen, S. Sun and G. Wu, Mater. Today Energy, 2021, 20, 100653 CrossRef CAS .
  137. C. Chen, T. Ma, Y. Shang, B. Gao, B. Jin, H. Dan, Q. Li, Q. Yue, Y. Li, Y. Wang and X. Xu, Appl. Catal., B, 2019, 250, 382–395 CrossRef CAS .
  138. L. Yang, H. Yang, S. Yin, X. Wang, M. Xu, G. Lu, Z. Liu and H. Sun, Small, 2022, 18, 2104941 CrossRef CAS PubMed .
  139. R. Ossola, O. M. Jönsson, K. Moor and K. McNeill, Chem. Rev., 2021, 121, 4100–4146 CrossRef CAS PubMed .
  140. F. Chen, X.-L. Wu, L. Yang, C. Chen, H. Lin and J. Chen, Chem. Eng. J., 2020, 394, 124904 CrossRef CAS .
  141. S. Guo, H. Wang, W. Yang, H. Fida, L. You and K. Zhou, Appl. Catal., B, 2020, 262, 118250 CrossRef CAS .
  142. K. Chaisiwamongkhol, C. Batchelor-McAuley, R. G. Palgrave and R. G. Compton, Angew. Chem., Int. Ed., 2018, 57, 6270–6273 CrossRef CAS PubMed .
  143. W.-Y. Tsai, R. Wang, S. Boyd, V. Augustyn and N. Balke, Nano Energy, 2021, 81, 105592 CrossRef CAS .
  144. R. Zheng, H. Yu, X. Zhang, Y. Ding, M. Xia, K. Cao, J. Shu, A. Vlad and B. Su, Angew. Chem., Int. Ed., 2021, 60, 18430–18437 CrossRef CAS PubMed .
  145. C. Huang, Y. Liu, R. Zheng, Z. Yang, Z. Miao, J. Zhang, X. Cai, H. Yu, L. Zhang, J. Shu and J. Mater, Sci. Technol., 2022, 107, 64–69 Search PubMed .
  146. M. A. Blommaert, D. A. Vermaas, B. Izelaar, B. in ‘t Veen and W. A. Smith, J. Mater. Chem. A, 2019, 7, 19060–19069 RSC .
  147. X. Yu, L. Wang, X. Wang, H. Liu, Z. Wang, Y. Huang, G. Shan, W. Wang and L. Zhu, J. Mater. Chem. A, 2021, 9, 7198–7207 RSC .
  148. Y. Wang, X. Zhao, D. Cao, Y. Wang and Y. Zhu, Appl. Catal., B, 2017, 211, 79–88 CrossRef CAS .
  149. J. Hu, Y. Li, Y. Zou, L. Lin, B. Li and X. Li, Chem. Eng. J., 2022, 437, 135428 CrossRef CAS .
  150. A. E. Platero-Prats, A. B. League, V. Bernales, J. Ye, L. C. Gallington, A. Vjunov, N. M. Schweitzer, Z. Li, J. Zheng, B. L. Mehdi, A. J. Stevens, A. Dohnalkova, M. Balasubramanian, O. K. Farha, J. T. Hupp, N. D. Browning, J. L. Fulton, D. M. Camaioni, J. A. Lercher, D. G. Truhlar, L. Gagliardi, C. J. Cramer and K. W. Chapman, J. Am. Chem. Soc., 2017, 139, 10410–10418 CrossRef CAS PubMed .
  151. A. W. Stubbs, L. Braglia, E. Borfecchia, R. J. Meyer, Y. Román-Leshkov, C. Lamberti and M. Dincă, ACS Catal., 2018, 8, 596–601 CrossRef CAS .
  152. J. Tang and J. Wang, Chem. Eng. J., 2018, 351, 1085–1094 CrossRef CAS .
  153. Z. Zhu, W. Lu, N. Li, T. Xu and W. Chen, Chem. Eng. J., 2017, 321, 58–66 CrossRef CAS .
  154. A. S. Rosen, J. M. Notestein and R. Q. Snurr, Angew. Chem., Int. Ed., 2020, 59, 19494–19502 CrossRef CAS PubMed .
  155. C. Qiao, Z. Usman, T. Cao, S. Rafai, Z. Wang, Y. Zhu, C. Cao and J. Zhang, Chem. Eng. J., 2021, 426, 130873 CrossRef CAS .
  156. H. Li, C. Shan and B. Pan, Environ. Sci. Technol., 2018, 52, 2197–2205 CrossRef CAS PubMed .
  157. H. Zhang, C. Li, L. Lyu and C. Hu, Appl. Catal., B, 2020, 270, 118874 CrossRef CAS .
  158. L. Chen, K. Xing, Q. Shentu, Y. Huang, W. Lv and Y. Yao, Chemosphere, 2021, 280, 130911 CrossRef CAS PubMed .

This journal is © The Royal Society of Chemistry 2022
Click here to see how this site uses Cookies. View our privacy policy here.