Next Article in Journal
Does Social Capital Benefit the Improvement of Rural Households’ Sustainable Livelihood Ability? Based on the Survey Data of Jiangxi Province, China
Next Article in Special Issue
Determination of Critical Limit of Zinc for Rice (Oryza sativa L.) and Potato (Solanum tuberosum L.) Cultivation in Floodplain Soils of Bangladesh
Previous Article in Journal
A Study on the Application of the Safety Practice Index to Reduce Safety Accidents in the Manufacturing Industry
Previous Article in Special Issue
Polyhalite Positively Influences the Growth, Yield and Quality of Sugarcane (Saccharum officinarum L.) in Potassium and Calcium-Deficient Soils in the Semi-Arid Tropics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

How Organic Substances Promote the Chemical Oxidative Degradation of Pollutants: A Mini Review

1
State Environmental Protection Key Laboratory of Soil Environmental Management and Pollution Control, Nanjing Institute of Environmental Sciences, Ministry of Ecology and Environment of China, Nanjing 210042, China
2
College of Environment, Hohai University, Nanjing 210098, China
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(19), 10993; https://doi.org/10.3390/su131910993
Submission received: 6 August 2021 / Revised: 26 September 2021 / Accepted: 27 September 2021 / Published: 3 October 2021
(This article belongs to the Special Issue Sustainable Soil Health Management)

Abstract

:
The promotion of pollutant oxidation degradation efficiency by adding organic catalysts has obtained widespread attention in recent years. Studies have shown that organic substances promote the process of traditional oxidation reactions by accelerating the redox cycle of transition metals, chelating transition metals, activating oxidants directly to generate reactive oxygen species such as hydroxyl and sulfate radical, or changing the electron distribution of the target pollutant. Based on the promotion of typical organic functional groups on the chemical oxidative process, a metal-organic framework has been developed and applied in the field of chemical catalytic oxidation. This manuscript reviewed the types, relative merits, and action mechanisms of common organics which promoted oxidation reactions so as to deepen the understanding of chemical oxidation mechanisms and enhance the practical application of oxidation technology.

1. Introduction

Chemical oxidation technology has gradually become one of the mainstream technologies applied in water body and contaminated site restoration engineering due to its good treatment effect and low application cost. Commonly used chemical oxidants include potassium, ozone, Fenton and Fenton-like reagents, activated persulfate (PDS), and etc. The traditional chemical oxidation technology has some deficiencies, such as low oxidation efficiency, pH limitation, and sludge accumulation caused by iron salt [1,2,3]. Therefore, a significant amount of research has begun to focus on exploring the technical means to improve the efficiency of chemical oxidation or broaden the pH scope of application. For example, the Fe3+, Cu2+, Co2+, and Mn2+ were used to replace Fe2+ to active H2O2, which were feasible for broadening the oxidative pH but would often decrease the oxidation rate [4,5], and solid Fe0 or ores could also activate H2O2 or PDS in near-neutral condition [6,7]. In recent years, more and more studies have revealed that the introduction of organic substances into the system can serve such purposes, and some oxidation intermediates of organic pollutants could also self-accelerate the reaction [8,9].
The organic substances which could improve the efficiency of chemical oxidation mainly existed in three forms: organic small molecular form itself, organic chelant, and metal-organic frameworks (MOFs) material. The recent research advances on the types of organic substances, their environmental affinity, and action mechanisms that can effectively improve the efficiency of oxidative degradation of pollutants were summarized in this manuscript, which means to provide a reference for the further study on chemical oxidation kinetics, application conditions, and mechanisms.

2. Reaction Acceleration by Organic Small Molecular Substances

2.1. Accelerated the Metal Redox Cycle

Huang et al. showed that in the Fe2+/Fe3+ cycle, Fe2+ regeneration is the oxidation rate control step [10]. Any method to improve the Fe3+ reduction efficiency will accelerate the formation of hydroxyl radicals (•OH), which is also the key to affect the Fenton/Fenton-like oxidation rate [11,12]. Compounds rich in nucleophilic groups such as carboxyl, hydroxyl, carbonyl, and amino groups can accelerate the Fe2+/Fe3+ cycle due to their strong reducibility, thereby accelerating the generation of free radicals. Subramanian and Madras [13] found that thioglycolic acid can effectively improve the Fe2+/Fe3+ cycle efficiency in the Fenton system under near-neutral pH conditions, and accelerate the degradation efficiency of organic pollutants. Moreover, Chen et al. showed that hydroxylamine can enhance the production of •OH by promoting Fe3+ reduction, and it is still effective when the pH is 5.7. Hydroxylamine was finally oxidized to NO3 and N2O in the system [14].
Advanced oxidation technology based on sulfate radical (SO4•) has similar characteristics [15,16]. For example, Yang et al. found that ascorbic acid (AA) can promote the oxidation effect of the Fe2+/PDS system at a pH range of 2.0 to 6.2 [17]. Garcia et al. studied the difference in the effect of tartaric acid (TA) and hydroxylamine hydrochloride added into the Fe2+/PDS system on the degradation process of bisphenol A [18,19]; the results showed that they both accelerated the efficiency of free radical generation and broadened the pH range of the reaction. The difference is that hydroxylamine hydrochloride only has a short-term promotion effect, while TA can promote the oxidation effect for a long time. The Fe2+ activates PDS to generateSO4•, which reacts with TA to generate organic compound radicals (R•), then Fe3+ reacts with R• slowly to generate Fe2+, forming a chain reaction which promotes the oxidation reaction for a long time. This is consistent with the results of Minisci et al. and Liang et al. [20,21].
Phenolic and quinone compounds have also been found to have the effect of accelerating Fenton/PDS oxidation. Xiao et al. found that phenolic compounds accelerated the Fenton-like degradation of dimethyl phthalate (DMP), with an increased efficiency of H2O2 utilization. The effect of phenolic compounds on the degradation of DMP followed the order: catechol ≈ hydroquinone > resorcinol > phenol, which could be attributed to the interaction between quinone-like substances and iron ions. Hydroquinone-like substances accelerated the Fe(III)/(II) redox cycle. The formation of iron complexes between catechol-like substances and iron ions facilitated the release of H+ and regeneration of Fe(II) [22]. Phenols and quinones are also typical intermediate oxidation products of aromatic compounds [8]. Jiang et al. [9,23] found that hydroquinone and p-benzoquinone, which were the oxidation intermediates of phenol and nitrobenzene, can promote the decomposition of Fe(III)–hydroperoxy complexes, and promote the conversion of Fe(II)/Fe(III). Therefore, the degradation of such aromatic compounds has an autocatalytic effect.
The promotion is also effective in heterogeneous oxidation system. Sun et al. found that AA can significantly increase the Fe2+/Fe3+ cycle on the surface of magnetite (Fe3O4), thereby accelerating the oxidation efficiency of the Fe3O4/H2O2 system on alachlor [24]. A Fenton-like system with MnOx-Fe3O4/biochar composite (FeMn/biochar) was constructed for pollutant degradation. Five well-characterized reducing agents (sodium borohydride (SBH), sodium thiosulfate (STS), AA, hydroxylamine, and oxalic acid (OA)) were added respectively to investigate their performance to the system. The results revealed that only OA and hydroxylamine obviously enhanced the catalytic capacity of the Fenton-like process and HA increased ciprofloxacin degradation efficiency from 38.2% to 92.8%, well in agreement with the accelerated Fe(III/II) cycle and Mn(III/II) cycle in the Fe/Mn/biochar-H2O2-HA system. The accelerated metal redox cycle could enhance the decomposition of H2O2 into •OH and •O2, which were the main reactive oxygen species responsible for ciprofloxacin degradation [25]. Sang et al. found that proper addition of hydroxylamine greatly promoted the degradation of sulfamethoxazole in each of the α-Fe2O3/peroxomonosulfate (PMS), Co2O3/PMS, and CuO/PMS systems. The results suggested that •OH radical was the dominating reactive oxygen species in the α-Fe2O3/hydroxylamine/PMS and Co2O3/hydroxylamine/PMS systems, while in the CuO/hydroxylamine/PMS system, the activated-PMS on the surface of CuO was concluded to be responsible for SMZ degradation instead of free radicals [18].

2.2. Activated H2O2 or PDS to Construct Advanced Oxidation System

In addition to accelerating the transition metal cycle in the advanced oxidation system, organic matter can also directly activate the oxidant to generate free radicals to build an advanced oxidation system.

2.2.1. Organic Matters Activate H2O2

Studies have shown that organic matter can directly activate H2O2 to produce oxidizing active substances, such as •OH, O2− and 1O2, which strengthen the oxidation capacity. The activation is generally related to the organic free radicals generated during the reaction (see Figure 1) [26,27].
In addition to being used as a promoter to improve oxidation efficiency, hydroxylamine can also directly activate H2O2 to form an advanced oxidation system. Chen et al. proposed that hydroxylamine activating H2O2 to produce •OH may be divided into two steps: the first step is hydroxylamine ions to activate H2O2 to produce •OH [27]; the second step is to produce •OH by the reaction of H2O2 with the protonated amino radical generated in the first step. They also proposed that the reaction between hydroxylamine and H2O2 to generate •OH may be related to the -OH group in hydroxylamine.
The quinone structure can also activate H2O2 and promote the generation of •OH. Zhu et al. studied the production mechanism of •OH during the activation of H2O2 by haloquinone, and the results showed that •OH is generated by tetrachloro-1,4-benzoquinone and H2O2 through a mechanism that has nothing to do with metals: the nucleophilic attack of H2O2 on tetrachloro-1,4-benzoquinone forms a trichlorohydroperoxy-1,4-benzoquinone (TrCBQ-OOH) intermediate, which is further cracked to produce •OH [28,29].

2.2.2. Organic Matter Activates PDS

With the rapid development of PDS activation technology, the new type of activation technology for organic activation of PDS has received widespread attention. Studies have shown that organic compounds such as quinone compounds, AA, hydroxylamine, phenols, quercetin, and surfactants can activate PDS.
Fang et al. found that both quinone compounds and humic acid can effectively activate PDS to degrade 2,4,4’-trichlorobiphenyl [30]. Its degradation rate could reach 88% in the p-benzoquinone/PDS system, while the degradation rates in a single PDS or p-benzoquinone system were only 20% and 9%, respectively. Zhang et al. explored the effects of different types of anthraquinone dissolved organic matter on the degradation of Rhodamine B by PDS [31]. Since anthraquinone-dissolved organic matter contains an oxidation-sensitive functional group structure, it not only transfers electrons in the PDS activation reaction [32], but also generates reductive semiquinone radicals during the activation process, which reduces S2O82− to SO4•and SO42− and can significantly enhance PDS oxidation capacity [33].
When studying the activation of PDS by AA, Hou et al. used the AA/PDS system to degrade atrazine. Compared with a single PDS oxidation system, after AA, SO4• and •OH are generated in the system, the degradation rate of atrazine is increased by 29 times [34]. Cao et al. showed that when the pH of the AA/PDS system is between 3.5 and 12.5, PDS is mainly activated by AA. When pH > 12.5, PDS is mainly activated by alkali [35].
The activation effect of phenols on PDS is mainly achieved by phenates. When the pH is 8.3, pentachlorophenol degrades pollutants by reducing PDS to generate active free radicals [36]. Quercetin (QCR) is a flavonoid polyphenol organic compound whose structure lacks electron delocalization and can release electrons to activate PDS, and then produce SO4• and •OH. When the pH is 13, compared with a single PDS system, the QCR/PDS system can effectively degrade 1,1-dichloroethane, 1,2-dichloroethane, 1,2-dichloropropane, and dibromomethane [37].
Surfactants are usually used for the desorption of pollutants and the dissolution of non-aqueous liquids in in-situ chemical oxidation processes. The results of studies have shown that anionic, nonionic, and cationic surfactants (docusate sodium, polyethylene glycol 400, and N-tallow propylene diamine polyoxyethylene ether) can effectively activate PDS [38]. Among them, the cationic surfactant N-tallow-based propylene diamine polyoxyethylene ether shows the strongest activation effect. It can generate •OH under alkaline conditions, and can generate reducing or nucleophilic groups (superoxide radicals, hydroperoxide anions, alkyl radicals, etc.) under acidic and alkaline conditions.

2.3. Other Effects

Except for Fenton, Fenton-like, and PDS, organic small molecular substances can also enhance the oxidizing ability of other oxidants. Yang et al. showed that the removal efficiency of phenol and bisphenol A (BPA) permanganate has a synergistic effect under weak acid conditions (pH 4.0~6.0) [39]. Among them, the removal effect of phenol increased with the increase of the initial concentration of BPA, but decreased with the pH. However, under weakly alkaline (pH 7.5~8.5) conditions, the two had a competitive effect on the degradation of permanganate. That is, the degradation of phenol was inhibited in the presence of BPA, but the degradation efficiency of BPA was slightly improved in the presence of phenol. The study speculated that the reason for the synergy is that bisphenol A induced the production of manganese oxide in the system, and the reason for the competition may be the formation of reactive manganese intermediates Mn5+ or Mn4+ in the oxidation system.
Degradation of pollutants by ozonation in the presence of hydroxylamine has been investigated in previous studies [40,41]. The results showed that the degradation rate could be improved obviously in the presence of HA, which was attributed to the production of •OH and singlet oxygen.

3. The Promotion of Organic Chelating Agents

Organic complexing agents can effectively prevent the precipitation of transition metals in a non-acidic environment, and the formed coordination field often affects the redox characteristics of Fe2+/Fe3+, thereby promoting the oxidation effect of the Fenton (Fenton-like) system [42,43]. The main action mechanism of organic small molecular substances and complexing agents are displayed in Figure 2. Commonly used organic complexing agents include humic substances, carboxylic acid compounds, and amino carboxylic acid compounds.

3.1. Humic Substances

Humus widely exists in natural water bodies, soils, and sediments [44]. Metal ions can form chelates with humic acid (HA) or fulvic acid (FA) by binding to carboxyl groups, phenolic groups, and nitrogen-containing sites [45], and HAs and FAs can effectively reduce Fe3+, thereby accelerating the Fenton system oxidation process [46,47]. Research results show that different concentrations of humic acid in the pH range of 5~7 often have a good effect on Fenton (Fenton-like) reaction [42,48]. However, when the pH is low, humic acid often has no obvious effect on the Fenton (Fenton-like) reaction, and even exerts a certain inhibitory effect. For example, Lipczynska-Kochany and Kochany found that adding 3000 mg L−1 humate to the Fenton system of pH 7 could greatly improve the removal efficiency of pollutants. However, when the pH was 3.5, the addition of humic acid salt actually inhibited the degradation [49]. FA presents similar performance [49,50]. Lindsey et al. showed that the inhibitory effect of humus on the degradation of pollutants under acidic pH conditions is related to the hydrophobicity of the substrate [51].
Humic acid can also promote the oxidation efficiency of permanganate and ozone. He et al. found that humic acid could promote the removal of phenol by potassium permanganate under the condition of pH 4~8, while it would inhibit the oxidation of phenol by potassium permanganate under the condition of pH 9~10 [52]. Potassium permanganate is an electrophilic reagent, and the oxidation rate increases with the increase of the electron cloud density on the aromatic ring of the target pollutant [53]. The aromatic ring structure in humic acid forms a π-π interaction with phenol, which increases the electron cloud density of phenol, thereby promoting the oxidation capacity of potassium permanganate [26,54]. The macromolecular humic acid has a high C = C content, and the π-π interaction has a good positive correlation with the C = C content. Therefore, the macromolecular humic acid further promotes the oxidation ability of potassium permanganate. Yang et al. found that adding low concentration (such as 1 mg L−1) humus to the manganese-catalyzed ozone oxidation system can help improve the degradation rate of atrazine [55]. However, when the concentration of humus is further increased, the degradation of atrazine is inhibited. The reason is that the low concentration of humic acid helps to stimulate the formation of •OH in the system, while the quenching of free radicals under the condition of high concentration of humic acid plays a leading role.

3.2. Carboxylic Acid

Molecular molar ratio and pH are important factors that affect the catalytic effect of Fe-carboxylic acid chelating agents. Generally, when the pH is in the range of acidic to neutral, the Fe-carboxylic acid chelating agent has a better catalytic effect. Citric acid (CA) is a commonly used carboxylic acid Fenton reaction complexing agent. Studies have shown that when the molar ratio of Fe and CA molecules is 1:1, there are three main forms, namely, [Fe(Cit)]°, [Fe(Cit)]+, and [Fe(Cit)(OH)] [56,57]. Studies have also reported the formation of a complex with a molar ratio of Fe and CA of 2:2 [58]. Lewis et al. showed that the Fenton reaction modified by CA chelation can effectively reduce the usage amount of Fe2+ under near-neutral (pH 6–7) conditions [59]. In addition, when the molar ratio of CA and Fe molecules is between 1:1 and 4:1, increasing the molar ratio of molecules will reduce the decomposition efficiency of H2O2. In addition, Li et al. pointed out that even when the pH is greater than 5, the Fenton system with CA as a chelating agent can still efficiently degrade 2,4,6-trichlorophenol. However, when pH > 8, a large amount of iron precipitation is formed [60]. Trovo et al. showed that CA can extend the application range of the photo-Fenton system to neutral pH conditions, and the degradation rate heavily depends on the initial concentration of citrate [61]. When the pH is 5–8, the degradation efficiency of diclofenac (DCF) in the photofenton method decreases with the increase of pH.
Except for CA, in recent years, some new carboxylic acid natural organic complexing agents have also been found to not only promote the Fe2+/Fe3+ cycle, but also effectively prevent the precipitation of iron ions. For example, Qin et al. found that adding protocatechuic acid (PCA) to the Fe3+/H2O2 Fenton-like system could effectively increase the degradation rate of alachlor when the pH was less than 7, but the degradation efficiency decreased with the increase of pH; when pH ≥ 7, PCA no longer increased the degradation rate of alachlor [62]. Ren et al. showed that rosmarinic acid (RA) could greatly enhance the oxidation efficiency of 2,4-dichlorophenol in the Fe3+/H2O2 system when the pH is was the range of 3–6 [63]. However, when the pH increased to 6.4 and 7.2, the promotion effect gradually weakened. Generally, both the target pollutant and the organic complexing agent can be effectively mineralized in the Fenton system, and this type of organic complexing agent has good environmental friendliness [62,63].

3.3. Amino Carboxylic Acid

Because of its strong complexing ability, amino carboxylic acid can prevent iron precipitation even under neutral pH conditions, and it is also a common organic complexing agent [10,64]. Ethylenediamine tetraacetic acid (EDTA) is the most commonly used aminocarboxylic acid complexing agent. The 1:1 organic ligand Fe2+ (EDTA) formed by EDTA-2Na and Fe2+ can not only extend the time of Fe2+ participating in the Fenton reaction, but also activate the dissolved oxygen in the system and spontaneously generate H2O2 [65]. Brian et al. showed that when the molar ratio of EDTA-2Na to Fe2+ was 1:1, the degradation effect of benzene and 1,2-dichlorobenzene was the best [66]. When the concentration of EDTA-2Na is further increased and exceeds the requirement for complexation, it will consume •OH by itself to reduce the degradation rate. In the EDTA-Fe3+-H2O2 system, as the molar ratio of EDTA: Fe3+ (range 1:1~5:1) increases, the decolorization efficiency of malachite green gradually increases. The degradation mechanism does not follow the simple hydroxyl radical mechanism, and the intermediate valence iron (four or five valence) that exists at the same time plays a major role in oxidation [67]. In recent years, EDTA has also been found to promote the activation of H2O2 by heterogeneous oxidation systems such as zero-valent iron or iron oxides [68,69,70]. However, EDTA has a strong capacity to chelate heavy metals and poor biodegradability, which may exert an adverse effect on the environment [71,72]. Therefore, in recent years, the search for biodegradable EDTA substitutes has gradually attracted the attention of researchers [73,74,75], such as N, N’-ethylenediamine disuccinic acid (EDDS), which is a structural isomer of EDTA and can exist in [S,S’], [S,R/R,S], and [R,R’] configurations (see Figure 3). Among them, the [S,S] configuration can be quickly and completely mineralized, and the other two configurations can be partially biodegraded [76]. The isomeric mixture of EDDS contains statistically 25% of (S,S)-EDDS, 25% of (R,R)-EDDS, and 50% of (R,S)/(S, R)-EDDS. Orama et al. proposed that the suitable pH range of EDDS as a Fe3+ chelating agent is 3–9 [77]. When pH ≤ 7, Fe3+-EDDS complex mainly exists in the form of Fe3+-EDDS-; when pH > 7, the Fe3+-EDDS complex mainly exists in the form of Fe(OH)EDDS2− and Fe(OH)2EDDS3−. Huang et al. found that in the Fenton reaction driven by EDDS, under neutral or alkaline conditions, due to the generation of •HO2 or •O2 radicals and the presence of various forms of complex iron, the oxidation efficiency was much higher than that of acidic conditions [8]. EDDS can not only keep iron in a soluble form, but also promote the generation of superoxide radicals, thereby promoting the generation of Fe2+ and •OH. Nitrilotriacetic acid (NTA) is also an amino carboxylic acid complexing agent that has been frequently studied in recent years. Sun and Pignatello tested 50 kinds of structurally diverse organic and inorganic polydentate chelators for their ability to solubilize Fe(III) at pH 6 and catalyzed the oxidation of 0.1 mM 2,4-dichlorophenoxyacetic acid by 10 mM H2O2 in aerated aqueous solution [78]. The result showed that NTA was one of the most active chelates that promote the degradation. The reactivity of the Fe3+-NTA system would not be affected by excessive NTA [79]. There is only one N atom in the NTA molecule, and it could be degraded by microorganisms under hypoxic conditions, which has a small risk of causing environmental problems [80].
Moreover, some other N-containing organic complexing agents have also been proven to have better Fe3+ complexing and the capacity to promote H2O2 decomposition. In addition to the production of hydroxyl radicals, the oxidation process is often accompanied by the production of high-valent iron compounds [78,81].
The promoting effect of organic matters on the Fenton-like system is also related to the type of transition metal. For example, Ghiselli et al. found that at pH 5.5, EDTA, NTA, and CA could inhibit the degradation of organic pollutants in the Fenton-like reaction catalyzed by copper [82], while TA could promote the degradation. The reason may be that EDTA, NTA, and CA have strong complexing capacity with Cu2+ and high stability constants, which prevented the interaction between the effective sites of Cu ions and H2O2. TA itself has strong reducibility and can reduce Cu2+ existing in a complex state to Cu+. The reaction of Cu+ with H2O2 can increase the production of •OH and promote the oxidative degradation of organic pollutants.
The characteristics of different types of chelating agents are listed in Table 1, and the application of typical organic small molecular substances and chelating agents in promoting oxidation reaction is summarized in Table 2.

4. Metal-Organic Framework Materials

In recent years, solid-phase catalytic materials have attracted extensive attention. The heterogeneous oxidation system composed of solid-phase catalytic materials and oxidants overcomes the disadvantages of homogeneous Fenton catalyst, such as difficult separation and repeated use, easy formation of iron sludge by Fe2+, increase of color in water, and narrow reaction pH range. Among them, metal-organic frameworks (MOFs) which have intramolecular voids formed by self-assembly of organic ligands and metal ions or clusters through coordination bonds [83,84,85,86], have a higher catalytic activity than traditional solid catalytic material and have been used in advanced oxidation systems. MOFs combines the excellent properties of organic and inorganic substances, so its effectiveness exceeds the performance of simple mixing [87,88]. Due to the diversity of metal ion species and the selection of organic ligands, MOFs materials have many advantages: high specific surface area, functional groups such as -NH2, -OH, and -COOH which can be introduced to the pore surface of MOFs material, high thermal and chemical stability, and etc. The performance can be adjusted through the synergistic effect of metal ions and organic functional groups, and post-synthetic modification can be performed to make it have special properties [89]. A typical mechanism for pollutant degradation catalyzed by MOF is shown in Figure 4.

4.1. Fe-Based MOFs Catalytic Materials

Iron-based MOFs are porous materials assembled from iron ions or iron clusters and organic ligands, which can promote the Fe3+/Fe2+ conversion rate in MOFs by means of metal doping and organic ligand modification and improve the heterogeneity catalytic H2O2 degradation efficiency of pollutants [91]. Fe-based MOFs were widely studied because Fe was eco-friendly and cost efficient. So far, a large number of iron-based MOFs using different types of organic ligands have been found. For example, MIL-53 (Fe), MIL-88B(Fe), and MIL-100 (Fe) have been reported as Fenton-like catalysts to degrade pollutants [92,93,94]. Li et al. prepared Fe2+-containing MOFs using 2, 2′-bipyridine-5,5′-dihydroxy acid as ligand [95]. The material activated H2O2 to degrade organic pollutants in near neutral conditions, and it had an excellent activity and stability. A high H2O2 utilization efficiency was also achieved. Lv et al. used Fe2+@MIL-100(Fe) Fenton catalyst to study the effect of MB degradation. The experimental results showed that Fe2+@MIL-100 (Fe) demonstrated the highest Fenton catalytic ability compared with MIL-100 (Fe) and Fe2O3 [94]. The Fe2+ and Fe3+ ions in Fe2+@MIL-100 (Fe) have a synergistic effect on the production of •OH, thereby promoting oxidation efficiency.

4.2. Copper- and Cobalt-Based MOFs Catalytic Materials

Other than Fe-based MOF, other metals such as copper and cobalt were also found to have the ability to catalyze H2O2 to generate •OH [96,97]. Because of the appropriate price and wide distribution, Cu was usually used in catalytic material. Lyn et al. synthesized Cu-doped mesoporous silica microspheres (Cu-MSM) by the hydrothermal method, and the Fenton-like process catalyzed by Cu-MSMs showed excellent performance in the degradation of phenytoin sodium (PHT) and diphenhydramine (DP). The reaction mechanism is that H2O2 was converted to •OH by the framework Cu+ in Cu-MSMs, and Cu+ was simultaneously oxidized to Cu2+. The generated •OH could cause the decomposition of PHT and DP, while the generated phenolic intermediates could be adsorbed on the surface of Cu-MSMs, complexed with the framework Cu2+ and form copper complexes. The complex can interact with H2O2 and promote Cu2+ reduction, accelerate the Cu+/Cu2+ circulation, generate •OH more efficiently, and thus promote the degradation of organic pollutants.
Cobalt ion is also a common heterogeneous Fenton oxidant, which can be used to catalyze oxidants such as H2O2, PDS, and peroxymonosulfate [91]. Co has a higher catalytic activity than Fe and Cu. For example, Racles et al. synthesized two kinds of MOFs containing Cu and Co respectively at room temperature, and added them to H2O2 to degrade the azo dye congo red (CR) [98]. The results showed that the degradation efficiency of the cobalt-based MOF material to CR reached 90% after reaction for 30 min, which was greater than that of the copper-based MOFs material. This indicates that to a certain extent the catalytic capacity of the cobalt-based MOFs material is greater than that of the copper-based MOFs material.

4.3. Multi-Core MOFs Catalytic Materials

In recent years, MOFs doped with multiple metals have been widely concerned [99,100]. Li et al. studied and synthesized Fe-Co Prussian blue complexes as photo-Fenton catalysts, and this material has high degradation efficiency for rhodamine B under the conditions of pH 3.0–8.5 [101]. During the entire photo-Fenton reaction process, H2O2 molecules replaced water molecules coordinated with Fe, and the resulting Fe2+–peroxide complex could generate •OH. At the same time, Fe3+ in the catalyst was reduced by H2O2, and the generated HOO• and •OH reacted to form 1O2, which directly participated in the degradation of rhodamine B. Wang et al. showed that under the condition of pH 5, the removal rate of 20 mg·L−1 MB by the MIL-101(Fe,Cu)/H2O2 system was 100% when the reaction time was 20 min [102]. Compared with MIL-101(Fe)/H2O2 and H2O2 alone, it increased by 43.1% and 88.9%, respectively. This is because Cu2+ doping introduces new active sites, where Cu2+/Cu+ can synergize with Fe3+/Fe2+ cycles to produce more •OH to improve the Fenton-like degradation effect.

5. Conclusions and Prospects

The oxidative degradation of pollutants in the water can be enhanced by adding organic substances or preparing solid materials with organic substances. Organics containing nucleophilic functional groups such as carboxyl, quinone, hydroxyl, and amino groups can effectively accelerate the generation of free radicals and broaden the pH range in Fenton, Fenton-like, and activated PDS systems by promoting the reduction of transition metals, complexing transition metals, or inhibiting the hydrolysis of transition metals. Some quinones, phenols and carboxylic acids can even directly activate H2O2/PDS to generate free radicals. These organic substances can also improve the oxidation efficiency of permanganate by changing the valence state of manganese. In addition, based on the characteristics of organic compounds which promote chemical oxidation reactions, MOFs materials are created to improve the oxidation effect of the advanced oxidation system, which constructed a heterogeneous system to release the oxidizing species steadily or re-utilize the catalysts.
However, at present, except traditional complexing agents such as CA and EDTA which are often used to enhance the oxidation effect in actual water bodies, most organics and MOFs materials are applied only in the laboratories. As a result, for future development of organics or organic-containing material in real environmental applications, more efforts can be made on the following aspects: (1) to decrease the acquisition cost of the organic materials; (2) to investigate their environmental risks; and (3) to clarify whether the organic matter unremoved during the chemical oxidation can offer available carbon for further microbiological deterioration of pollutants.

Author Contributions

Z.H. wrote the paper; J.S. conceived and designed the paper; H.Y., J.H. and F.S. collected and analyzed the data. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China, grant number 2018YFC1801100; National Natural Science Foundation of China, grant number 21707041.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available upon request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Matzek, L.W.; Carter, K.E. Activated Persulfate for Organic Chemical Degradation: A Review. Chemosphere 2016, 151, 178–188. [Google Scholar] [CrossRef] [PubMed]
  2. Kumar, J.E. Decolourisation, mineralisation and detoxification of mixture of azo dyes using Fenton and Fenton-type advanced oxidation processes. Chemical. Pap. 2020, 74, 3145–3159. [Google Scholar] [CrossRef]
  3. Chávez, A.M. Simulated solar photocatalytic ozonation of contaminants of emerging concern and effluent organic matter in secondary effluents by a reusable magnetic catalyst. Chem. Eng. J. 2020, 398, 125–642. [Google Scholar] [CrossRef]
  4. Martins, A.F.; Vasconcelos, G.; Wilde, M.L. Influence of variables of the combined coagulation–Fenton-sedimentation process in the treatment of trifluraline effluent. J. Hazard. Mater. 2005, 127, 111–119. [Google Scholar] [CrossRef] [PubMed]
  5. Verma, P.; Baldrian, P.; Gabriel, J.; Trnka, T.; Nerud, F. Copper–ligand complex for the decolorization of synthetic dyes. Chemosphere 2004, 57, 1207–1211. [Google Scholar] [CrossRef]
  6. Costa, R.; Moura, F.; Ardisson, J.D.; Fabris, J.D.; Lago, R.M. Highly active heterogeneous Fenton-like systems based on FeO/Fe3O4 composites prepared by controlled reduction of iron oxides. Appl. Catal. B Environ. 2008, 83, 131–139. [Google Scholar] [CrossRef]
  7. Liu, J.; Du, Y.; Ye, T.; Cao, J.; Peng, C. Removal of orange II using an adsorbent-supported zero-valent iron as a heterogeneous Fenton-like catalyst. Desalination Water Treat. 2020, 175, 273–281. [Google Scholar] [CrossRef]
  8. Chen, R.; Pignatello, J.J. Role of Quinone Intermediates as Electron Shuttles in Fenton and Photoassisted Fenton Oxidations of Aromatic Compounds. Environ. Sci. Technol. 1997, 31, 2399–2406. [Google Scholar] [CrossRef]
  9. Jiang, C.; Pang, S.; Ouyang, F. A new insight into Fenton and Fenton-like processes for water treatment. J. Hazard. Mater. 2010, 174, 813–817. [Google Scholar] [CrossRef]
  10. Huang, W.Y.; Brigante, M.; Wu, F. Assessment of the Fe(III)-EDDS Complex in Fenton-like Processes: From the Radical Formation to the Degradation of Bisphenol, A. Environ. Sci. Technol. 2013, 47, 1952–1959. [Google Scholar] [CrossRef]
  11. Lien, H.L. Bimetallic Fe/Al system: An all-in-one solid-phase Fenton reagent for generation of hydroxyl radicals under oxic conditions. Sci. Total. Environ. 2019, 673, 480–488. [Google Scholar] [CrossRef]
  12. Shen, J.; Hou, L.; Sun, H. Effect of an electro-Fenton process on the biodegradation of lignin by Trametes versicolor. Bioresources 2020, 15, 8039–8050. [Google Scholar] [CrossRef]
  13. Subramanian, G.; Madras, G. Remarkable Enhancement of Fenton Degradation at a Wide pH Range Promoted by Thioglycolic Acid. Chem. Commun. 2017, 53, 1136–1139. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, L.; Ma, J.; Li, X. Strong Enhancement on Fenton Oxidation by Addition of Hydroxylamine to Accelerate the Ferric and Ferrous Iron Cycles. Environ. Sci. Technol. 2011, 45, 3925–3930. [Google Scholar] [CrossRef] [PubMed]
  15. Lin, Y.C.; Chang, F.T. Optimizing Operating Parameters of A Honeycomb Zeolite rotor Concentrator for Processing TFT-LCD Volatile Organic Compounds with Competitive Adsorption Characteristics. J. Hazard. Mater. 2009, 164, 517–526. [Google Scholar] [CrossRef]
  16. Teel, A.L.; Cutler, L.M.; Watts, R.J. Effect of Sorption on Contaminant Oxidation in Activated Persulfate Systems. J. Environ. Sci. Health Part A 2009, 44, 1098–1103. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, L.; Hui, Z.; Wang, J.; Jia, A. Rapid and Continuous Oxidation of Organic Contaminants with Ascorbic Acid and A Modified Ferric/Persulfate System. Chem. Eng. J. 2015, 270, 73–79. [Google Scholar]
  18. Sang, W.; Li, Z.; Huang, M.; Wu, X.; Cui, J. Enhanced transition metal oxide based peroxymonosulfate activation by hydroxylamine for the degradation of sulfamethoxazole. Chem. Eng. J. 2020, 383, 123057. [Google Scholar] [CrossRef]
  19. Minisci, F.; Citterio, A.; Giordano, C. Electron-transfer processes: Peroxydisulfate, a useful and versatile reagent in organic chemistry. Cheminform 1983, 16, 27–32. [Google Scholar] [CrossRef]
  20. Liang, C.; Liang, C.P.; Chen, C.C. pH Dependence of Persulfate Activation by EDTA/Fe(III) for Degradation of Trichloroethylene. J. Contam. Hydrol. 2009, 106, 173–182. [Google Scholar] [CrossRef]
  21. Di, H.; Xiaoiaohong, G.; Jun, M. Influence of different nominal molecular weight fractions of humic acids on phenol oxidation by permanganate. Environ. Sci. Technol. 2009, 43, 8332–8337. [Google Scholar]
  22. Xiao, J.; Wang, C.; Liu, H. Fenton-like degradation of dimethyl phthalate enhanced by quinone species. J. Hazard. Mater. 2019, 382, 121007. [Google Scholar] [CrossRef]
  23. Jiang, C.C.; Gao, Z.; Qu, H.L. A New Insight into Fenton and Fenton-Like Processes for Water Treatment: Part II. Influence of Organic Compounds on Fe(III)/Fe(II) Interconversion and the Course of Reactions. J. Hazard. Mater. 2013, 250–251, 76–81. [Google Scholar] [CrossRef]
  24. Sun, H.; Xie, G.; He, D. Ascorbic Acid Promoted Magnetite Fenton Degradation of Alachlor: Mechanistic Insights and Kinetic Modeling. Appl. Catal. B-Environ. 2020, 267, 118–383. [Google Scholar] [CrossRef]
  25. Li, L.; Liu, S.; Cheng, M.; Lai, C.; Long, M. Improving the Fenton-like catalytic performance of MnOx-Fe3O4/biochar using reducing agents: A comparative study. J. Hazard. Mater. 2020, 406, 124333. [Google Scholar] [CrossRef]
  26. Nanny, M.A.; Maza, J.P. Noncovalent interactions between monoaromatic compounds and dissolved humic acids: A deuterium NMR T1 relaxation study. Environ. Sci. Technol. 2001, 35, 379–384. [Google Scholar] [CrossRef]
  27. Chen, L.W.; Li, X.C.; Zhang, J. Production of hydroxyl radical via the activation of hydrogen peroxide by hydroxylamine. Environ. Sci. Technol. 2015, 49, 10373–10379. [Google Scholar] [CrossRef] [PubMed]
  28. Zhu, B.Z.; Kalyanaraman, B.; Jiang, G.B. Molecular Mechanism for Metal independent Production of Hydroxyl Radicals by Hydrogen Peroxide and Halogenated Quinones. Proc. Natl. Acad. Sci. USA 2007, 104, 17575–17578. [Google Scholar] [CrossRef] [Green Version]
  29. Zhu, B.Z.; Zhao, H.T.; Kalyanaraman, B. Metal-independent production of hydroxyl radicals by halogenated quinones and hydrogen peroxide: An ESR spin trapping study. Free. Radic. Biol. Med. 2002, 32, 465–473. [Google Scholar] [CrossRef]
  30. Fang, G.D.; Gao, J.; Dionysiou, D.D. Activation of persulfate by quinones: Free radical reactions and implication for the degradation of PCBs. Environ. Sci. Technol. 2013, 47, 4605–4611. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, G.; Zhou, Y.; Yang, F. FeOOH-catalyzed heterogeneous electro-Fenton system upon anthraquinone@graphene nanohybrid cathode in a divided electrolytic cell: Catholyte-regulated catalytic oxidation performance and mechanism. J. Electrochem. Soc. 2015, 162, H357–H365. [Google Scholar] [CrossRef]
  32. Kang, S.H.; Choi, W. Oxidative degradation of organic compounds using zero-valent iron in the presence of natural organic matter serving as an electron shuttle. Environ. Sci. Technol. 2009, 43, 878–883. [Google Scholar] [CrossRef]
  33. Leng, Y.; Guo, W.; Shi, X. Polyhydroquinone-coated Fe3O4 nanocatalyst for degradation of rhodamine B based on sulfate radicals. Ind. Eng. Chem. Res. 2013, 52, 13607–13612. [Google Scholar] [CrossRef]
  34. Hou, X.; Zhan, G.; Huang, X. Persulfate activation induced by ascorbic acid for efficient organic pollutants oxidation. Chem. Eng. J. 2020, 382, 122355. [Google Scholar] [CrossRef]
  35. Cao, M.; Hou, Y.; Zhang, E. Ascorbic acid induced activation of persulfate for pentachlorophenol degradation. Chemosphere 2019, 229, 200–205. [Google Scholar] [CrossRef]
  36. Ahmad, M.; Teel, A.L.; Watts, R.J. Mechanism of persulfate activation by phenols. Environ. Sci. Technol. 2013, 47, 5864–5871. [Google Scholar] [CrossRef] [PubMed]
  37. Lin, Y.T.; Chiu, Y.T.; Ciou, C. Natural organic activator quercetin for persulfate oxidative degradation of halogenated hydrocarbons. Environ. Sci. Water Res. Technol. 2019, 5, 1064–1071. [Google Scholar] [CrossRef]
  38. Elloy, F.C.; Teel, A.L.; Watts, R.J. Activation of persulfate by surfactants under acidic and basic conditions. Groundw. Monit. Remediat. 2014, 34, 51–59. [Google Scholar] [CrossRef]
  39. Yang, J.; Zhao, J.; Ma, J.; Liu, Z.-Q. Simultaneous oxidation of phenol and bisphenol a by permanganate: Synergetic or competitive effect. Sep. Purif. Technol. 2013, 116, 271–276. [Google Scholar] [CrossRef]
  40. Yang, J.; Li, J.; Dong, W.; Ma, J.; Cao, J.; Li, T. Study on enhanced degradation of atrazine by ozonation in the presence of hydroxylamine. J. Hazard. Mater. 2016, 316, 110–121. [Google Scholar] [CrossRef]
  41. Qin, W.; Yuan, X.; Sun, L.; Qiang, Z.; Xia, D. Insights into the activation of ozonation by hydroxylamine: Influential factors, degradation mechanism and reaction kinetics. J. Hazard. Mater. 2019, 373, 600–607. [Google Scholar] [CrossRef]
  42. Georgi, A.; Schierz, A.; Trommler, U.; Horwitz, C.P.; Collins, T.J.; Kopinke, F.D. Humic Acid Modified Fenton Reagent for Enhancement of the Working pH Range. Appl. Catal. B Environ. 2007, 72, 26–36. [Google Scholar] [CrossRef]
  43. Ying, Z.; Zhou, M.A. Critical Review of the Application of Chelating Agents to Enable Fenton and Fenton-like Reactions at High pH Values. J. Hazard. Mater. 2018, 362, 436–450. [Google Scholar]
  44. Davydova, N.D.; Znamenskaya, T.I. Barrier Functions of Soils of Natural and Technogenic Steppe Landscapes. Geogr. Nat. Resour. 2019, 40, 384–393. [Google Scholar] [CrossRef]
  45. Davies, G.; Fataftah, A.; Cherkasskiy, A.; Ghabbour, E.A.; Radwan, A.; Jansen, S.A. Tight metal binding by humic acids and its role in biomineralization. J. Chem. Soc. Dalton Trans. 1997, 21, 4047–4060. [Google Scholar] [CrossRef]
  46. Li, Z.M.; Shea, P.J.; Comfort, S.D. Fenton Oxidation of 2,4,6-trinitrotoluene in Contaminated Soil Slurries. Environ. Eng. Sci. 1997, 14, 55–66. [Google Scholar] [CrossRef]
  47. Yu, H.; Liu, G.; Jin, R.; Zhou, J. Goethite-humic acid coprecipitate mediated Fenton-like degradation of sulfanilamide: The role of coprecipitated humic acid in accelerating Fe(III)/Fe(II) cycle and degradation efficiency. J. Hazard. Mater. 2021, 403, 124026. [Google Scholar] [CrossRef]
  48. Lipczynsk, A.; Kochany, E.; Kochany, J. Effect of humic substances on the Fenton treatment of wastewater at acidic and neutral pH. Chemosphere 2008, 73, 745–750. [Google Scholar] [CrossRef] [PubMed]
  49. Voelker, B.M.; Sulzberger, B. Effects of fulvic acid on Fe(II) oxidation by hydrogen peroxide. Environ. Sci. Technol. 1996, 30, 1106–1114. [Google Scholar] [CrossRef]
  50. Lindesy, M.E.; Tarr, M.A. Inhibition of Hydroxyl Radical Reaction with Aromatics by Dissolved Natural Organic Matter. Environ. Sci. Technol. 2000, 34, 444–449. [Google Scholar] [CrossRef]
  51. Lindsey, M.E.; Tarr, M.A. Inhibited hydroxyl radical degradation of aromatic hydrocarbons in the presence of dissolved fulvic acid. Water Res. 2000, 34, 2385–2389. [Google Scholar] [CrossRef]
  52. He, D.; Guan, X.; Ma, J. Influence of humic acids of different origins on oxidation of phenol and chlorophenols by permanganate. J. Hazard. 2010, 182, 681–688. [Google Scholar] [CrossRef] [PubMed]
  53. Yan, Y.E.; Schwartz, F.W. Oxidative degradation and kinetics of chlorinated ethylenes by potassium permanganate. J. Contam. Hydrol. 1999, 37, 343–365. [Google Scholar] [CrossRef]
  54. Smejkalova, D.; Spaccini, R.; Fontaine, B. Binding of Phenol and Differently Halogenated Phenols to Dissolved Humic Matter as Measured by NMR Spectroscopy. Environ. Sci. Technol. 2009, 43, 5377–5382. [Google Scholar] [CrossRef] [PubMed]
  55. Ma, J.; Graham, N.J.D. Degradation of atrazine by manganese-catalysed ozonation: Influence of humic substances. Water Res. 1999, 33, 785–793. [Google Scholar] [CrossRef]
  56. Warner, R.C.; Weber, I. The cupric and ferric citrate complexes. J. Am. Chem. Soc. 1953, 75, 5086–5094. [Google Scholar] [CrossRef]
  57. Field, T.B.; Mccourt, J.L.; Mcbryde, W.A.E. Composition and Stability of Iron and Copper Citrate Complexes in Aqueous Solution. Can. J. Chem. 1974, 52, 3119–3124. [Google Scholar] [CrossRef]
  58. Timberlake, C.F. Iron-Malate and Iron-Citrate Complexes. J. Chem. Soc. 1964, 8, 5078–5085. [Google Scholar] [CrossRef]
  59. Lewis, S.; Lynch, A.; Bachas, L. Chelate-modified Fenton Reaction for the Degradation of Trichloroethylene in Aqueous and Two-phase Systems. Environ. Eng. Sci. 2009, 26, 849–859. [Google Scholar] [CrossRef]
  60. Li, Y.; Bachas, L.G.; Bhattacharyya, D. Kinetics Studies of Trichlorophenol Destruction by Chelate-based Fenton Reaction. Environ. Eng. Sci. 2005, 22, 756–771. [Google Scholar] [CrossRef]
  61. Trovo, A.G.; Nogueira, R.F.P. Diclofenac Abatement Using Modified Solar Photo-Fenton Process with Ammonium Iron(III) Citrate. J. Braz. Chem. Soc. 2011, 22, 1033–1039. [Google Scholar] [CrossRef]
  62. Qin, Y.; Song, F.; Ai, Z. Protocatechuic Acid Promoted Alachlor Degradation in Fe(III)/H2O2 Fenton System. Environ. Sci. Technol. 2015, 49, 7948–7956. [Google Scholar] [CrossRef]
  63. Ren, H.; Jin, X.; Li, C.G. Rosmarinic Acid Enhanced Fe(III)-mediated Fenton Oxidation Removal of Organic Pollutants at Near Neutral pH. Sci. Total. Environ. 2020, 736, 139528. [Google Scholar] [CrossRef] [PubMed]
  64. Hu, Y.; Li, Y.L.; He, J.Y. EDTA-Fe(III) Fenton-like Oxidation for the Degradation of Malachite Green. J. Environ. Manag. 2018, 226, 256–263. [Google Scholar] [CrossRef] [PubMed]
  65. Seibig, S.; Eldik, R.V. Kinetics of [FeII(EDTA)] Oxidation by Molecular Oxygen Revisited. New Evid. A Multistep Mechanism. Inorg. Chem. 1997, 36, 4115. [Google Scholar]
  66. Brian, K.; Gullett, L.; Oudejans, D. Near-real-time combustion monitoring for PCDD/PCDF indicators by GC-REMPI-TOFMS. Environ. Sci. Technol. 2011, 46, 923–928. [Google Scholar]
  67. Tweedy, S.E.; Benitez, A.R.; Narayan, A.; Zimmerman, P.M.; Wymore, T. The hydroxyl radical coupled electron transfer mechanism of flavin-dependent hydroxylases. J. Phys. Chem. B 2019, 123, 8065–8073. [Google Scholar] [CrossRef]
  68. Banavath, R.; Srivastava, R.; Bhargava, P. Improved non-enzymatic H2O2 sensors using highly electroactive cobalt hexacyanoferrate nanostructures prepared through EDTA chelation route. Mater. Chem. Phys. 2021, 267, 124593. [Google Scholar] [CrossRef]
  69. Feng, D.; Hong, C.; Sida, Z. Effect of chelating agents on reactive Green 19 decolorization through Fe0-activated persulfate oxidation process. J. Environ. Manag. 2017, 200, 325–334. [Google Scholar]
  70. Pan, Y.; Bu, Z.; Sang, C. EDTA enhanced pre-magnetized Fe0/H2O2 process for removing sulfamethazine at neutral pH. Sep. Purif. Technol. 2020, 250, 117281. [Google Scholar] [CrossRef]
  71. Englehardt, J.D.; Meeroff, D.E.; Echegoye, N.L. Oxidation of aqueous EDTA and associated organics and coprecipitation of inorganics by ambient iron-mediated aeration. Environ. Sci. Technol. 2007, 41, 270–276. [Google Scholar] [CrossRef] [PubMed]
  72. Kari, F.G.; Giger, W. Speciation and Fate of Ethylenediaminetetraacetate (EDTA) in Municipal Wastewater Treatment. Water Res. 1996, 30, 122–134. [Google Scholar] [CrossRef]
  73. Sun, S.P.; Zeng, X.; Li, C. Enhanced heterogeneous and homogeneous Fenton-like degradation of carbamazepine by nano-Fe3O4/H2O2 with nitrilotriacetic acid. Chem. Eng. J. 2014, 244, 44–49. [Google Scholar] [CrossRef]
  74. Ye, Z.; Brillas, E.; Centellas, F. Expanding the Application of Photoelectro-Fenton Treatment to Urban Wastewater Using the Fe(III)-EDDS Complex. Water Res. 2020, 169, 115219. [Google Scholar] [CrossRef] [PubMed]
  75. Miralles-cuevas, S.; Oller, I.; Ruiz-delgado, A. EDDS as complexing agent for enhancing solar advanced oxidation processes in natural water: Effect of iron species and different oxidants. J. Hazard. Mater. 2019, 372, 129–136. [Google Scholar] [CrossRef]
  76. Schowanek, D.; Feijtel, T.C.J.; Perkins, C.M. Biodegradation of [s,s], [r,r] and Mixed Stereoisomers of Ethylene Diamine Disuccinic Acid (EDDS), A Transition Metal Chelator. Chemosphere 1997, 34, 2375–2391. [Google Scholar] [CrossRef]
  77. Orama, M.; Saarinen, H.; Aksela, R. Complexation of [s,s] and Mixed Stereoisomers of N,N’-Ethylenediaminedisuccinic Acid (EDDS) with Fe(III), Cu(II), Zn(II) and Mn(II) Ions in Aqueous Solution. J. Chem. Soc. Dalton Trans. 2002, 24, 4644–4648. [Google Scholar] [CrossRef]
  78. Sun, Y.; Pignatello, J.J. Chemical Treatment of Pesticide Wastes. Evaluation of Fe(III) Chelates for Catalytic Hydrogen Peroxide Oxidation of 2,4-D at Circumneutral pH. J. Agric. Food Chem. 1992, 40, 74–81. [Google Scholar] [CrossRef]
  79. Bates, G.W.; Billups, C.; Saltman, P. The Kinetics and Mechanism of Iron(III) Exchange Between Chelates and Transferrin, I. The Complexes of Citrate and Nitrilotriacetic Acid. J. Biol. Chem. 1971, 246, 2810. [Google Scholar] [CrossRef]
  80. White, V.E.; Knowles, C.J. Degradation of copper-NTA by Mesorhizobium sp. NCIMB 13524. Int. Biodeterior. Biodegrad. 2003, 52, 143–150. [Google Scholar] [CrossRef]
  81. Canals, M.; Gonzalez-olmos, R.; Costas, M. Robust Iron Coordination Complexes with N-Based Neutral Ligands as Efficient Fenton-Like Catalysts at Neutral pH. Environ. Sci. Technol. 2013, 47, 9918–9927. [Google Scholar] [CrossRef]
  82. Ghiselli, G.; Jardim, W.F.; Litter, M.I.; Mansilla, H.D. Destruction of EDTA using Fenton and photo-Fenton-like reactions under UV-A irradiation. J. Photochem. Photobiol. A Chem. 2004, 167, 59–67. [Google Scholar] [CrossRef]
  83. Chang, Z.; Yang, D.H.; Xu, J. Flexible Metal–Organic Frameworks: Recent Advances and Potential Applications. Adv. Mater. 2015, 27, 5432–5441. [Google Scholar] [CrossRef] [PubMed]
  84. Chang, Z.; Yang, D.H.; Xu, J.; FDecoste, J.B.; Peterson, G.W. Metal–Organic Frameworks for Air Purification of Toxic Chemicals. Chem. Rev. 2014, 114, 5695. [Google Scholar]
  85. Zhu, L.; Meng, L.; Shi, J. Metal-organic frameworks/carbon-based materials for environmental remediation: A state-of-the-art mini-review. J. Environ. Manag. 2019, 232, 964–977. [Google Scholar] [CrossRef]
  86. Cheng, M.; Hu, L.; Huan, G.D. Metal-organic frameworks for highly efficient heterogeneous Fenton-like catalysis. Coord. Chem. Rev. 2018, 368, 80–92. [Google Scholar] [CrossRef]
  87. Toyao, T.; Saito, M.; Dohshi, S. Development of a Ru Complex-Incorporated MOF Photocatalyst for Hydrogen Production under Visible-light Irradiation. Chem. Commun. 2014, 50, 6779–6781. [Google Scholar] [CrossRef]
  88. Xuan, W.; Zhu, C.; Liu, Y. Mesoporous Metal–organic Framework Materials. Chem. Soc. Rev. 2012, 41, 1677–1695. [Google Scholar] [CrossRef]
  89. Jiang, L.; Cui, Z.; Yu, H.; Hou, X. Synergistic effect of a magnetic field and carbon nanotubes on the adsorption performance of metal ions and phenols. Desalination Water Treat. 2019, 151, 295–306. [Google Scholar] [CrossRef]
  90. Yao, Y.; Hao, C.; Chao, L.; Wei, F.; Wang, S. Fe, Co, Ni nanocrystals encapsulated in nitrogen-doped carbon nanotubes as Fenton-like catalysts for organic pollutant removal. J. Hazard. Mater. 2016, 314, 129–139. [Google Scholar] [CrossRef] [PubMed]
  91. Yu, D.; Wang, L.; Yang, T.; Yang, G.; Wu, M. Tuning lewis acidity of iron-based metal-organic frameworks for enhanced catalytic ozonation. Chem. Eng. J. 2021, 404, 127075. [Google Scholar] [CrossRef]
  92. Ai, L.; Zhang, C.; Li, L. Iron Terephthalate Metal–organic Framework: Revealing the Effective Activation of Hydrogen Peroxide for the Degradation of Organic Dye under Visible Light Irradiation. Appl. Catal. B Environ. 2014, 148–149, 191–200. [Google Scholar] [CrossRef]
  93. Gao, C.; Chen, S.; Quan, X. Enhanced Fenton-like Catalysis by Iron-based Metal Organic Frameworks for Degradation of Organic Pollutants. J. Catal. 2017, 356, 125–132. [Google Scholar] [CrossRef]
  94. Lv, H.; Zhao, H.; Cao, T. Efficient Degradation of High Concentration Azo-dye Wastewater by Heterogeneous Fenton Process with Iron-based Metal-organic Framework. J. Mol. Catal. A Chem. 2015, 400, 81–89. [Google Scholar] [CrossRef]
  95. Li, Y.; Liu, H.; Li, W.J. A Nanoscale Fe(II) Metal-organic Framework with a Bipyridinedicarboxylate Ligand as a High Performance heterogeneous Fenton catalyst. Rsc. Adv. 2016, 6, 6756–6760. [Google Scholar] [CrossRef]
  96. Choi, K.; Lee, W. Enhanced Degradation of Trichloroethylene In Nano-scale zero-Valent Iron Fenton System with Cu(II). J. Hazard. Mater. 2012, 211–212, 146–153. [Google Scholar] [CrossRef]
  97. Guimaraes, I.R.; Giroto, A.; Oliveira, L.C.; Guerreiro, M.C.; Lima, D.Q.; Fabris, J.D. Synthesis and Thermal Treatment of Cu-doped Goethite: Oxidation of Quinoline through Heterogeneous Fenton Process. Appl. Catal. B Environ. 2009, 91, 581–586. [Google Scholar] [CrossRef]
  98. Racles, C.; Zaltariov, M.F.; Iacob, M. Siloxane-based Metal–Organic Frameworks with Remarkable Catalytic Activity in Mild Environmental Photodegradation of Azo Dyes. Appl. Catal. B: Environ. 2017, 205, 78–92. [Google Scholar] [CrossRef]
  99. Pariyar, A.; Yaghoobnejad, A.H.; Choundhury, A. Tetragonal Versus Hexagonal: Structure-dependent Catalytic Activity of Co/Zn Bimetallic Metal-Organic Frameworks. Inorg. Chem. 2016, 55, 9250–9257. [Google Scholar] [CrossRef]
  100. Sun, Q.; Liu, M.; Li, K. Synthesis of Fe/M (M=Mn, Co, Ni) bimetallic metal organic frameworks and their catalytic activity for phenol degradation at mild conditions. Inorg. Chem. Front. 2017, 4, 144–153. [Google Scholar] [CrossRef]
  101. Li, X.; Liu, J.; Rykov, A.I.; Han, H.; Jin, C.; Liu, X.; Wang, J. Excellent Photo-Fenton Catalysts of Fe-Co Prussian Blue Analogues and Their Reaction Mechanism Study. Appl. Catal. B Environ. 2015, 179, 196–205. [Google Scholar] [CrossRef]
  102. Wang, Z.; Zhao, H.; Qi, H.; Liu, X.; Yang, L. Free radical behaviours during methylene blue degradation in Fe2+/H2O2 system. Environ. Technol. 2017, 40, 1–25. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The mechanism of H2O2/PDS activation by organics.
Figure 1. The mechanism of H2O2/PDS activation by organics.
Sustainability 13 10993 g001
Figure 2. The main action mechanism of small molecular substances and chelator in Fe-mediated systems.
Figure 2. The main action mechanism of small molecular substances and chelator in Fe-mediated systems.
Sustainability 13 10993 g002
Figure 3. Chemical structure of [S,S], [S,R/R,S], and [R,R] configurations of EDDS.
Figure 3. Chemical structure of [S,S], [S,R/R,S], and [R,R] configurations of EDDS.
Sustainability 13 10993 g003
Figure 4. Mechanisms for dye degradation by the Co@N-C/PMS system. Reproduced with permission from Yao [90].
Figure 4. Mechanisms for dye degradation by the Co@N-C/PMS system. Reproduced with permission from Yao [90].
Sustainability 13 10993 g004
Table 1. The merits and demerits of different chelators.
Table 1. The merits and demerits of different chelators.
ChelatorMeritDemerit
HumusUbiquitous in natureNo obvious influence and even inhibition when pH is low
Carboxylic acidSuitable for acid to neutral range; Chelating agent can effectively degradeAffected by pH and molar ratio to transition metal
Amino carboxylic acidThe time of Fe2+ participation in Fenton reaction can be extended; It can activate the dissolved oxygen in the system and produce H2O2 spontaneously; The effect is better under neutral neutral or alkaline conditionsThe molar ratio of transition metal has influence on the promoting effect; Environmental degradation is poor
Table 2. Chemical oxidation degradation of pollutants by organics.
Table 2. Chemical oxidation degradation of pollutants by organics.
Organics ClassOrganic SubstanceStructureOxidantReaction pHTarget PollutantSource
PhenolsHydroquinone Sustainability 13 10993 i001H2O2 (10 mmol L−1), Fe3+ (0.5 mmol L−1) and hydroquinone (0.1 mmol L−1)3.1~3.2DMP
(1 mmol L−1)
[22]
Pentachlorophenol Sustainability 13 10993 i002Na2S2O8 (0.5 mol L−1) and pentachlorophenol (1 mmol L−1)6.5~10.5Nitrobenzene
(1 mmol L−1)
[36]
AA Sustainability 13 10993 i003H2O2 (1 mmol L−1), AA (0.5 mmol L−1) and Fe3O4 (1 g L−1)4Alachlor
(20 mg L−1)
[24]
Na2S2O8 (40 mmol L−1) and AA (1.0 mmol L−1)7.2Pentachlorophenol (10 mg L−1)[35]
HumusHumic Acid/H2O2 (130 mmol L−1), Fe2+ (30 μmol L−1) and HA (50~100 mg L−1)5~7Benzene
(25 μmol L−1)
[42]
H2O2 (50 mmol L−1), Fe2+ (5 mmol L−1) and HA (10 mg L−1)6.515 organic compounds[48]
Carboxylic Acid CompoundsCA Sustainability 13 10993 i004H2O2 (50 mmol L−1), Fe2+ (10 mmol L−1) and CA (10 mmol L−1)5~72,4,6–trichlorophenol
(1.5 mmol L−1)
[60]
Gallic Acid Sustainability 13 10993 i005H2O2 (8 mmol L−1) and Fe3+ (0.1 mmol L−1)3.6gallic acid
(0.11 mmol L−1)
[62]
Amino Carboxylic AcidsEDTA Sustainability 13 10993 i006Na2S2O8 (5 mmol L−1), Fe0(1.0 g L−1) and EDTA (1 mmol L−1)6.0Reactive Green 19 (0.05 mmol L−1)[69]
EDDSa mixture consisted of different configurationsH2O2 (5 mol L−1) and Fe3+-EDDS (1 mol L−1)6.2bisphenol A
(20 μmol L−1)
[10]
NTA Sustainability 13 10993 i007H2O2 (100 mmol L−1), Fe3O4 (1.0 g L−1) and NTA (0.5 mmol L−1)7Carbamazepine (63.5 μmol L−1)[73]
OthersHydroxylamine Sustainability 13 10993 i008H2O2 (0.4 mmol L−1), Fe2+ (10.0 μmol L−1) and NH2OH (0.4 mmol L−1)2.0~5.7benzoic acid (40.0 μmol L−1)[14]
1,4-benzoquinone Sustainability 13 10993 i009Persulfate (5 mmol L−1)7.4PCB 28
(0.5 mg L−1)
[30]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hu, Z.; Shi, J.; Yang, H.; Huang, J.; Sheng, F. How Organic Substances Promote the Chemical Oxidative Degradation of Pollutants: A Mini Review. Sustainability 2021, 13, 10993. https://doi.org/10.3390/su131910993

AMA Style

Hu Z, Shi J, Yang H, Huang J, Sheng F. How Organic Substances Promote the Chemical Oxidative Degradation of Pollutants: A Mini Review. Sustainability. 2021; 13(19):10993. https://doi.org/10.3390/su131910993

Chicago/Turabian Style

Hu, Zhewei, Jiaqi Shi, Hao Yang, Jianbo Huang, and Feng Sheng. 2021. "How Organic Substances Promote the Chemical Oxidative Degradation of Pollutants: A Mini Review" Sustainability 13, no. 19: 10993. https://doi.org/10.3390/su131910993

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

Article Metrics

Back to TopTop