Next Article in Journal
Fabrication of Z-Scheme CeVO4/BiVO4 Heterojunction and Its Enhanced Photocatalytic Degradation of NFX
Previous Article in Journal
Marine Mannitol: Extraction, Structures, Properties, and Applications
Previous Article in Special Issue
Increasing Surface Functionalities of FeCl3-Modified Reed Waste Biochar for Enhanced Nitrate Adsorption Property
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances and Prospects in Electrocatalytic Processes for Wastewater Treatment

by
Xince Zhou
1,
Jiajie Yang
1,
Jiahuan Guo
1,
Wei Xiong
1,2,* and
Michael K. H. Leung
2
1
Key Laboratory of Industrial Ecology and Environmental Engineering (Ministry of Education), School of Environmental Sciences and Technology, Dalian University of Technology, Dalian 116024, China
2
Ability R&D Energy Research Centre, School of Energy and Environment, City University of Hong Kong, Hong Kong, China
*
Author to whom correspondence should be addressed.
Processes 2024, 12(8), 1615; https://doi.org/10.3390/pr12081615 (registering DOI)
Submission received: 1 July 2024 / Revised: 20 July 2024 / Accepted: 29 July 2024 / Published: 1 August 2024
(This article belongs to the Special Issue Advanced Oxidation Process for Wastewater Treatment)

Abstract

:
Wastewater pollution is severe, with various refractory compounds extensively used and discharged into sewage, posing risks to the environment and human health. Electrocatalytic technologies including direct and indirect electrocatalytic oxidation, electrocatalytic reduction, and electro-Fenton processes offer advantages such as high efficiency, ease of control, and minimal secondary pollution. This review aims to systematically introduce the principles, current research status, advantages, and disadvantages of various electrocatalytic processes used for wastewater treatment, with a focus on the electrode materials, operational parameters, and cost analysis of various electrocatalytic technologies. It also provides new insights into efficient electrode materials for future electrocatalytic technologies in treating refractory wastewater.

1. Introduction

Freshwater, a vital resource for life and human civilization, constitutes only a minor fraction of the Earth’s total water resources. The total volume of usable freshwater for humans and ecological processes reaches approximately 200,000 cubic kilometers, representing just 1% of the freshwater resources. Currently, global freshwater reserves are rapidly depleting, significantly impacting many densely populated areas. With the rapid development of technology and human society, water resources essential for daily life and agricultural and industrial production are increasingly contaminated by chemical pollutants from human activities. Over 100,000 chemical substances are detected in the water environment, raising widespread concern among the environmental science community and the public. The main types of wastewaters include domestic, industrial, agricultural, and medical wastewater [1]. Chemical substances and excessive nutrients in wastewater lead to ecosystem degradation, affecting soil, freshwater, and marine environments, resulting in food insecurity and other social issues.
Refractory pollutants, including antibiotics, personal care products, dyes, endocrine disruptors, and heavy metals are currently widely used and discharged into the environment [2]. The sources and hazards of these pollutants are shown in Table 1. These pollutants are characterized by persistence, long-range transport, bioaccumulation, and high toxicity. They can travel over long distances, and accumulate in the food chain, affecting organisms at higher trophic levels. These substances can cause carcinogenic, teratogenic, and mutagenic effects. Emerging pollutants like perfluorooctanoic acid pose serious threats to natural ecosystems and biological health, with significant impacts on tissues and organs [3]. Antibiotics such as penicillin and streptomycin can induce hypersensitivity reactions, ranging from mild rashes and fever to hematopoietic suppression and even neurological damage [4]. Endocrine disruptors harm reproductive, developmental, and immune systems in organisms [5]. For example, consumption of fish contaminated with polychlorinated biphenyls by pregnant women may lead to shortened gestation, low birth weight, intelligence deficits, memory issues, and delayed neurodevelopment. Heavy metals such as mercury have led to neurological disorders, exemplified by Minamata disease in Japan. Excessive lead exposure can cause anemia, neurological damage, and memory decline, and, in pregnant women, high levels of lead can result in hearing loss or reduced intelligence in infants.
Common methods for treating refractory pollutants in wastewater include physical, chemical, and biological approaches, and their advantages and disadvantages are shown in Table 2. Physical methods, such as coagulation–precipitation, activated carbon adsorption, and membrane filtration, are highly efficient and widely applicable. Researchers have made a lot of efforts to improve the effectiveness of these methods. For instance, activated carbon fiber composites are developed by introducing new materials into the pores or on the surface of activated carbon to improve adsorption capacity. When activated carbon fibers are catalytically activated with iron and magnesium oxides, π-bonds form between chlorine and metal atoms, significantly increasing the adsorption of dichloroethylene, with equilibrium adsorption reaching up to 7 mmol g−1 [6]. However, physical methods primarily transfer pollutants rather than degrade them completely. In comparison, chemical treatment involves chemical reactions to separate, transform, destroy, or recycle refractory pollutants in wastewater, including neutralization, chemical precipitation, redox reactions, ion exchange, and electrolysis. For example, adding ozone can oxidize and remove phenols and cyanides, while hydroxides or barium salts can precipitate and remove heavy metals and cyanides. Chemical methods offer high degradation efficiency and allow for selective treatment of pollutants. However, they usually require chemical reagents, resulting in high costs and strict operational requirements, often necessitating the combination with other techniques. Recently, photothermal catalysis has emerged as a promising technology for wastewater treatment in recent years. This approach integrates photochemical and thermocatalytic processes, enabling the efficient utilization of the full solar spectrum to drive various chemical reactions for environmental remediation and energy conversion [7]. This integration offers several advantages over traditional methods, potentially leading to significant advancements in wastewater treatment. Biological treatments, such as activated sludge and biofilm processes, provide effective degradation of complex wastewater with various pollutants. For example, the bacterium Pseudomonas aeruginosa NY3 can continuously produce hydrogen peroxide, hydroxyl (•OH) radicals, and superoxide radicals under aerobic conditions, oxidizing phenanthrene in aqueous solutions [8]. However, biological methods have limitations, including long processing times, sensitivity to environmental conditions, and the need for extended periods and sophisticated equipment to effectively remove toxic and non-biodegradable organic pollutants.
In recent years, electrocatalysis has garnered increasing attention as an advanced oxidation process. Electrocatalysis is characterized by its eco-friendliness, lack of secondary pollution, ease of control, and high efficiency. Under an applied electric field, refractory organic pollutants undergo direct or indirect redox reactions on the electrode surface or in the water, leading to their decomposition and water purification. Electrocatalysis typically requires only the addition of electrolyte solutions or specific activators, such as chloride ions for hypochlorous acid generation. This significantly reduces the amount of chemical reagents needed compared to traditional methods. As a result, electrocatalysis minimizes the potential for secondary pollutant formation, contributing to a more environmentally friendly treatment approach. The main techniques include direct electrocatalytic oxidation (DEO), indirect electrocatalytic oxidation (IEO), electrocatalytic reduction (ER), and the electro-Fenton process (EFP). A comparison of the advantages and disadvantages of four electrocatalytic technologies is presented in Table 3. Significant progress has been made in wastewater treatment using electrocatalysis over the past decade [9].
This review aims to provide a systematic overview of electrocatalytic processes for wastewater treatment. It delves into the fundamental principles behind these processes, the current state of research, and the inherent advantages and limitations of various electrocatalytic techniques. The discussion encompasses critical aspects such as the selection of suitable electrode materials, the impact of operational parameters, and a cost analysis of different electrocatalytic technologies. The review also explores promising future directions for the development and implementation of this technology.

2. DEO

DEO includes two pathways, namely, electrocatalytic conversion and electrocatalytic combustion, as illustrated in Figure 1a. In DEO, H2O is electrolyzed on the anode surface to generate adsorbed •OH radicals. When using an active electrode, these •OH interacts with the oxide layer on the anode, transferring oxygen into the oxide lattice to form higher-valent oxides (MOx+1) (Equation (1)). MOx+1 are highly oxidative, selectively oxidizing organic compounds on the electrode surface, primarily resulting in electrochemical conversion where organic pollutants are not fully oxidized. This process degrades adsorbed organic pollutants into smaller, non-toxic, or biodegradable molecules, suitable for further biological treatment, such as oxidizing aromatic compounds to fatty acids. Common active electrodes include Ti/IrO2, Ti/RuO2, and Pt electrodes. Xu et al. studied the oxidation of formic acid on a platinum electrode using cyclic voltammetry, finding that it undergoes direct oxidation without forming formate as an intermediate [10]. Li et al. used Ti/Sb-SnO2, Ti/Sb-SnO2-Pb3O4, and Ti/Sb-SnO2-PbO2 for the direct electrocatalytic oxidation of aniline. [10] After 5 h of electrolysis in 5 g L−2 Na2SO4 solution, the removal rates of aniline were 95.9%, 90.5%, and 88.4%, respectively.
MOx(•OH) → MOx+1 + H+ + e,
When the anode is a non-active electrode, •OH is physically adsorbed on the electrode surface (Equation (2)). The •OH directly reacts with organic compounds, with no selectivity, primarily leading to electrochemical combustion, which deeply oxidizes organic pollutants into stable inorganic substances such as CO2 and H2O. Common non-active electrodes include PbO2, boron-doped diamond (BDD), and SnO2 electrodes. Kapalka et al. studied the oxidation of ammonia on BDD electrodes at high pH (>8) through radical oxidation and found that chlorine radicals enhance ammonia oxidation under low pH conditions [11]. In an electrolytic solution of 0.1 M NaCl with 0.05 M ammonia at pH 5.5 and a current density of 30 mA cm−2, ammonia concentration decreased linearly, and residual active chlorine was very low, indicating that all generated active chlorine reacted with ammonia. Recent progress of DEO has been summarized in Table 4.
MOx + H2O → MOx(•OH) + H+ + e,

3. IEO

IEO utilizes the addition of metal or non-metal ions to the electrolyte, forming a homogeneous electrocatalytic system. These ions undergo oxidation at the anode surface, generating a range of active species including hydroxyl radicals (•OH), sulfate radicals (SO4•−), superoxide radicals (•O2), and reactive chlorine species (RCS) [15]. These powerful oxidants then degrade organic pollutants due to their high reactivity. The generation of these active species is essential for the reaction (Figure 1b, Equations (3)–(5)). Notably, the presence of chloride ions during electrolysis can lead to the formation of hypochlorite, further accelerating the oxidation of organic contaminants (Equations (6)–(8)). IEO leverages a combination of intermediate oxidation and direct anodic oxidation, ultimately enhancing the overall efficiency of the degradation process. Importantly, these reactive species are short-lived, existing only during electrolysis. Once the current is interrupted, they are no longer generated and disappear from the solution. Recent progress of IEO for refractory pollutants is summarized in Table 5.
OH → •OH + e,
S2O82− + Mn+ → M(n+1)+ + SO4•− + SO42−
O2 + e → •O2
2Cl → Cl2 + 2e
Cl2 + H2O → HClO + H+ + Cl
HClO → H+ + ClO

3.1. •OH-Based IEO

Hydroxyl radicals (•OH) are highly reactive oxygen species with strong oxidizing potential (2.8 V) and high electronegativity. These properties enable them to react rapidly with various biomolecules, organic pollutants, and even inorganic species like ammonia and nitrogen, resulting in a wide range of chemical reactions. These radicals can oxidize pollutants through addition or hydrogen abstraction, decomposing them into CO2, H2O, N2, and other inorganic compounds [33]. However, their usefulness is limited by their short lifespan (less than 1 μs) and restricted diffusion distance. Cho et al. addressed these limitations by developing a three-dimensional electrode reactor (3DER) using granular activated carbon (GAC) in place of a traditional two-dimensional electrode reactor (2DER) [34]. The GAC particles function as microelectrodes, enhancing ibuprofen (IBP) degradation compared to a 2D system. This improvement is attributed to the increased surface area promoting earlier adsorption of IBP and facilitating its indirect oxidation by dissolved radicals. Additionally, the 3DER design catalyzes the formation of hydrogen peroxide (H2O2) from dissolved oxygen (O2), which in turn reacts to generate •OH. This enhanced •OH production contributes to the impressive 98% IBP removal rate achieved by the 3DER within 4 h, exceeding the 2DER by a factor of 2.5. Arias et al. studied the anodic oxidation (AO) of volatile organic compounds (VOCs) at high current densities, showing that boron-doped diamond (BDD) electrodes produced sufficient reactive species, including •OH radicals, leading to an 80% mineralization rate of VOCs [35]. Yang et al. doped La3+ into the SnO2 crystal structure of the anode, employing it for a high-performance electrochemical wastewater treatment. The resultant SnOx/La-Sb anode significantly enhanced the ability to degrade moxifloxacin, achieving nearly 100% removal within 30 min of treatment. Notably, the process also exhibited low energy consumption, requiring only 0.09 kWh m−3. The improved performance could be attributed to that the La3+ doping-induced oxygen vacancy activated the anode surface by boosting the interfacial electron transfer and •OH generation (Figure 2a) [26].
The presence of •OH radicals in various systems can be confirmed through spectroscopic analysis. The basic principle involves the reaction of •OH with coumarin (C9H6O2) to produce the highly fluorescent compound 7-hydroxycoumarin (C9H6O3), which can be detected under fluorescence spectroscopy [36]. The fluorescence intensity is linearly related to the concentration of 7-HC, allowing for the construction of a standard curve. Experimental data show that the reaction yields 29% of total hydroxyl radicals as 7-HC. The generation of •OH radicals can be calculated based on the concentration of coumarin and fluorescence intensity. The •OH radicals can also be detected using electron spin resonance (ESR), a technique for qualitative analysis of radicals by transferring them into a magnetic field. However, due to the instability and rapid reaction of radicals in electrocatalytic oxidation, direct detection with a spectrometer is not feasible. A spin trap must be used to form stable adducts for analysis. DMPO is commonly used for detecting •OH, with a characteristic peak ratio of 1:2:2:1 for the DMPO-OH adduct. Interestingly, Wang et al. investigated the sources of 5,5-dimethyl-1-pyrroline N-oxide (DMPO) adducts in systems containing Cu(II), Cu(II)/H2O2, Cu(II)/peroxymonosulfate (PMS), and Cu(II)/persulfate (PDS), finding that DMPO-•OH signals could be observed in most Cu(II) systems even in the absence of peroxides, indicating alternative sources for these adducts beyond the intended spin trapping of •OH by DMPO [37].

3.2. SO4•−-Based IEO

Sulfate radicals (SO4•−) possess a high oxidation potential of 2.5~3.1 V and a broad pH application range, with a lifespan of 30~40 μs in water. In AO processes utilizing sulfate electrolytes, SO4•− and S2O82− are the primary active species generated. The dominant transformation pathways involve the direct oxidation of SO42− at the anode surface to form S2O82−, which is subsequently activated into SO4•− through radical or non-radical mechanisms. Additionally, •OH produced from water oxidation also contribute to the formation of SO4•−. Chen and Saha revealed that the strong oxidative ability of •OH facilitates the generation of SO4•−, and the production rate of SO4•− is correlated to that of •OH [38,39].
However, it is crucial to note that while sulfate ions act as precursors for SO4•−, they also compete with pollutants for •OH. Consequently, increasing sulfate concentration does not always enhance pollutant removal. Moreover, excessive sulfate concentrations can hinder electrocatalytic performance by blocking active sites on the anode. To address these limitations, Wang et al. developed an Fe/Co-N/P-9 electrode (carbonized at 900 °C), where graphite carbon nanofibers provided good conductivity (Figure 2b) [40]. Iron and cobalt served as catalytic centers, and heteroatoms influenced the electronegativity of neighboring carbon atoms, affecting PMS adsorption. This design offers a large surface area for efficient mass transfer and utilizes a graphitic carbon layer for enhanced stability. The catalyst showed remarkable performance in degrading Rhodamine B. The presence of both SO4•− and •OH in the system was confirmed through radical quenching experiments and ESR spectroscopy. Additionally, Gao et al. employed the ESR secondary radical spin trapping method utilizing DMPO and dimethyl sulfoxide as a typical •OH scavenger. Their findings revealed the generation of •OH across a range of conditions, from weakly acidic (pH = 5.5) to strongly alkaline (optimal pH = 13.0). However, SO4•− was identified as the dominant radical species at pH < 5.5 [41].

3.3. •O2-Based IEO

The •O2 exhibits lower reactivity compared to other reactive oxygen species. However, its advantages lie in its extended lifespan (around 1 ms in aqueous solutions) and ability to travel longer distances, allowing it to effectively reach and degrade target pollutants. Zhang et al. employed a three-dimensional reactor containing ZnAl-layered double hydroxide/activated carbon (ZnAl-LDH/AC) as particle electrodes for the electrocatalytic degradation of nitrosopyrrolidine (NPYR) in water, achieving an impressive removal rate of 83.78% under optimal conditions (Figure 2c) [42]. ESR measurements utilizing DMPO and TEMP as spin traps revealed the presence of both DMPO-•OH and DMPO-•O2 adducts. Subsequent radical quenching experiments confirmed that •OH and •O2 were the primary active species, while singlet oxygen (1O2) also contributed to the degradation of NPYR, albeit to a lesser extent. Notably, model optimization and DFT calculations indicated that the nitroso oxygen atom in NPYR is more susceptible to electrophilic and radical attacks compared to the nitrogen atom. Consequently, radical attack preferentially targets and breaks the N-N and N-O bonds within the NPYR molecule. A BDD anode achieved complete removal (100%) of acid orange 74 (AO74) and an 84.3% reduction in chemical oxygen demand (COD) within 2 h under specific operating conditions (40 mA cm−2 current density and 2.5 g L−1 Na2SO4 electrolyte) [43]. Interestingly, the removal efficiency of AO74 decreased by 7.94% when p-benzoquinone, a superoxide scavenger, was introduced. This observation suggests the involvement of •O2 produced during BDD anode oxidation. These superoxide radicals can further react to form H2O2, which contributes to enhanced AO74 oxidation.
It is important to note that DMPO can also be used for ESR detection of •O2. The characteristic ESR spectrum of the DMPO-•O2 adduct exhibits four lines with an equal intensity ratio (1:1:1:1). Interestingly, Wang et al. investigated the sources of DMPO adducts in various systems containing Cu(II) ions with different oxidants (H2O2, PMS, and PDS) [37]. They found that DMPO-•O2 adducts only appeared under highly alkaline conditions within the Cu(II)/PDS system. This suggests that the formation of •O2 in this specific case might be linked to a unique reaction pathway.

3.4. RCS-Based IEO

Chlorides are ubiquitous electrolytes in wastewater, significantly influencing the generation of active species during electrochemical oxidation processes. In chloride-containing media, the primary RCS includes Cl2, HClO, ClO, ClO2, •Cl, and •ClO. The specific composition of RCS depends on the prevailing pH and Cl concentration. Chloride ions undergo a two-electron transfer process to form Cl2, which subsequently reacts with H2O to generate the weak acid HClO. Depending on the pH, HOCl reaches an equilibrium with the ClO. Additionally, under specific conditions, some AO systems can produce •Cl and •ClO [44,45,46]. Notably, the formation of chlorinated byproducts should be minimized in practical applications of electrochemical oxidation with Cl.
The speciation of RCS is heavily influenced by pH. At pH values below 1, Cl2 and HClO coexist, with Cl2 dominating. As the pH increases from 1 to 4, HClO becomes the predominant species. Between pH 4 and 7.5, both HClO and ClO exist, with HClO prevailing. Finally, at pH exceeding 7.5, ClO becomes the primary form, acting as a weaker oxidant compared to HClO [47]. Importantly, Zheng et al. demonstrated that the presence of Cl and HCO3 in electrochemical systems promotes the formation of ClO species, consequently facilitating the simultaneous removal of refractory organics and nitrogen from secondary coking wastewater [48]. Their study achieved an impressive 87.8% COD removal and 86.5% total nitrogen removal using a PbO2/Ti anode and a Cu/Zn cathode at a current density of 37.5 mA cm−2 for 6 h.
Due to their strong oxidizing properties, RCS find applications in various fields. Wang et al. employed anodic oxidation in Cl-containing electrolytes to generate RCS [49]. The presence of these radicals was confirmed using probe experiments, and they demonstrated superior performance in self-cleaning ultrafiltration membranes compared to other reactive oxygen species systems. In the medical field, reactive radicals offer a promising approach to combating tumors without inducing drug resistance, a limitation associated with traditional therapies. However, the efficacy of existing ROS-based cancer treatments is hampered by their low production rates. Song et al. discovered that •Cl possessing strong nucleophilicity can effectively target and attack intracellular biomolecules, triggering apoptosis in cancer cells [50]. This approach offers advantages in terms of high efficiency and low cost. Photocatalysis provides another avenue for RCS utilization. Zhou et al. employed •Cl to rapidly convert amino groups in urine to N2, achieving selective urea denitrification [51]. Additionally, RCS also demonstrated its applications in sterilization [52]. The RCS was generated through photolysis and the presence was confirmed via quenching and kinetic simulations.
RCS plays a crucial role in electrochemical oxidation processes due to their unique combination of properties such as low oxidation potential, short lifespan, and strong oxidizing power. This combination makes them a cost-effective and efficient method for degrading pollutants in chloride-containing electrolytes [53]. The primary reactions leading to RCS formation involve the oxidation of Cl ions, generating •Cl radicals alongside Cl2 and HClO (Equations (9)–(11)). Subsequently, ClO can further react to form •ClO (Equations (12) and (13)), and Cl can also produce •Cl2 radicals (Equations (14)–(16)).
Cl → •Cl + e
Cl + •Cl → Cl2 + e
Cl2 + H2O → HClO + Cl + H+
ClO + •OH → •ClO + OH
ClO + •Cl → •ClO + Cl
H2O → •OH + H+ + e
Cl + •OH → •Cl +OH
Cl + •Cl → •Cl2
Currently, chlorine radical detection methods primarily fall into two categories: qualitative detection and quantitative detection. Qualitative detection utilizes ESR technology, where DMPO is added as a radical trapping agent. The detection spectrum typically shows a seven-line pattern, corresponding to oxidized derivatives of DMPO-X [54]. For example, Wang et al. successfully utilized a mixed cobalt–nickel electrocatalytic system with NaCl as the electrolyte for selective methane conversion [55]. Employing EPR analysis, they detected strong signals corresponding to •Cl and •ClO radicals, confirming their role as the primary reactive species during the electrocatalytic process. A novel electrocatalytic oxidation process employing a Fe-DSA dual anode configuration had been established for leachate pretreatment. In this process, chloride in the leachate was converted to active chlorine by the DSA anode, while Fe(OH)3 was generated from the Fe anode. Consequently, Fe(VI) was formed in situ through the reactions between active chlorine and Fe(OH)3. The decomposition of Fe(VI) produced iron oxide, amorphous iron-oxygen hydroxide, and amorphous Fe(OH)3, which enhanced the removal performance through coagulation (Figure 2d) [30].
Quantitative detection involves free radical quenching experiments. Here, specific chemicals are used to scavenge and react with the radicals of interest. Table 6 provides details on commonly used RCS quenching agents, their types, and their corresponding rate constants. Li et al. constructed a CoFe2O4 anodic oxidation system, employing a heterogeneous electro-Fenton/Cl system for the degradation of norfloxacin (NOR) [56]. Through free radical quenching experiments using different scavengers like benzoic acid (BA), carbamazepine (CBZ), and 1-4Dioxomethylbenzene (DMOB), they were able to quantitatively determine the steady-state concentrations of •Cl, •ClO, and •Cl2 during the degradation process. These concentrations were found to be 2.50 × 10−12 M, 2.72 × 10−13 M, and 1.39 × 10−11 M, respectively.

4. ER

Electrocatalytic reduction offers a promising approach for treating oxyanions associated with heavy metals, halogens, nitrate ions, and other similar contaminants. This technology leverages an externally applied electric field to drive the reduction of high-valence cations present in these pollutants at the cathode surface. Through this electron gain, the reduction process transforms the pollutants into non-toxic elemental substances or into less toxic intermediates that are more readily biodegradable (Figure 1c). The recent research on electrocatalytic reduction for the degradation of pollutants is shown in Table 7.

4.1. ER of Heavy Metals

Due to the non-degradability and bio-toxicity of heavy metals (HMs), as well as their threat to public health and ecosystems, these pollutants have recently become a global focus of concern. ER technology involves the transformation of high-valence HMs into low-valence or zero-valent states by accepting electrons. Therefore, in this method, active sites are released by converting HMs to their zero-valent state. This technology is categorized into two types: direct reduction, where HMs directly accept electrons from the cathode surface, and indirect reduction, where electrons are accepted through the help of a reducing agent and in-situ-generated active radicals. In the direct reduction of heavy metals, HMs directly accept electrons from the cathode surface, followed by nucleation to form metal oxides or elemental metals. In the indirect reduction of HMs, electrons are transferred through a mediator, such as a reducing agent or in-situ-generated reactive radicals (e.g., •H, e, H2O2), to achieve HM reduction [66].
ER can also address the challenge of heavy metal complexes. However, achieving high selectivity of complex reduction remains a significant challenge. Qin et al. used a molybdenum disulfide nanosheets/graphite felt (MoS2/GF) cathode, achieving an average Faradaic efficiency of 29.6% for Cu-EDTA at −0.65 V (vs. SCE), with a specific removal rate (SRR) of 0.042 mol cm−2 h−1 [66]. Detoxification of metal complexes is crucial in the electrochemical reduction of HMs. Pan et al. designed BaTiO3@graphene electrodes and fabricated them into three-dimensional millimeter-scale spheres [67]. They demonstrated that 100% of Cu-EDTA was in situ dissociated by the piezoelectric potential of barium titanate@graphene. The addition of graphene significantly enhanced the surface potential of barium titanate@graphene through its flexoelectric effect (increasing from 19.8 ± 0.97 to 96.8 ± 1.48 mV), effectively promoting piezoelectric catalysis. Additionally, Cu(II) released from Cu-EDTA dissociation was simultaneously recovered via interactions with graphene groups. Yang employed inert electrodes (IrO2–Ta2O5/Ti anode and graphite cathode) to remove mixed heavy metal ions extracted from soils by citrate. This combined cathodic precipitation process decreased the concentrations of toxic heavy metals by over 99.4% after 12 h of electrolysis at a current density of 26 mA cm−2. The low potential of the cathode facilitated the electrodeposition of Cd, while the local alkaline environment, created by electro-mediated water reduction, led to the hydrolytic precipitation of Zn and Pb. Additionally, the precipitation of Fe from Fe-rich soil resulted in the coprecipitation of As on the cathode surface (Figure 3a) [58].

4.2. ER for the Dehalogenation

ER of halogens involves utilizing hydrogen protons generated from the electrocatalytic water splitting process to produce •H. •H can attack the carbon–halogen bonds, replacing the halogen atoms and hydrogenating the carbon–halogen bonds, thereby achieving halogen reduction. Taking the reduction of chlorinated organic compounds on metallic Pd as an example, the reduction process is illustrated by Equations (17)–(20).
Pd + RCl + H2O → Pd-RCl + Pd-H2O
Pd-H2O + e → Pd-•H + OH
Pd-RCl + Pd-•H + e → Pd-RH + Cl
Pd-RH → Pd + RH
Electrocatalytic reduction of halogens commonly employs cathode materials such as transition metals including Pd, Fe, Co, Ni, Ag, and Cu. These metals form M-H bonds that selectively absorb and store •H, effectively preventing •H from escaping as H2. They exhibit high activity in breaking carbon–halogen bonds. Jiang et al. utilized supermolecular coordination cages (SCCs) as templates and prepared Pd nanoparticle catalysts on carboxymethyl cellulose (CMC) hydrogel supports via the impregnation-reduction method [68]. Their study demonstrated that SCCs not only significantly reduced the aggregation tendency of Pd nanoparticles but also played a crucial role in adjusting their size and size distribution, providing guidance for the construction of electrode catalysts.
Transition metal phosphides are also common cathode materials where transition metal and phosphorus sites serve as proton acceptors and hydride acceptor centers, thereby enhancing electrocatalytic efficiency. Yao et al. reported a novel amorphous nickel phosphide catalyst (ANP) for removing trichloroacetic acid (TCAA) from water [69]. The superhydrophilic surface of ANP facilitated the contact between the electrocatalyst and the electrolyte, reducing impedance and promoting rapid electron transfer from the electrode to water, thereby enhancing atomic hydrogen production. ANP exhibited electrochemical dechlorination efficiency comparable to Pd-based catalysts, with an apparent reaction rate constant of 0.0283 min−1. Liu et al. prepared crystalline cobalt phosphide nanosheets arrayed on titanium plates (C-CoP/Ti) for cathodic dechlorination of fluorophenylkone (FLO) [70]. Under −1.2 V conditions, the C-CoP/Ti electrode achieved a dehalogenation rate of 97.4% for 20 mg L−1 FLO within 30 min, with nearly complete degradation of Cl and 20% of F in 120 min.
Carbonaceous materials such as graphene electrodes, activated carbon electrodes, and graphite felt electrodes are also viable cathode materials. King et al. utilized activated carbon cathodes in 100 mM phosphate electrolyte for electrocatalytic reduction of halogenated alkanes, observing degradation consistent with first-order kinetics with degradation rates reaching 90% [71]. Higher substitution of halogen groups correlated with slower degradation rates. Carbon materials are also common electrode modification materials, enhancing the dispersion of catalytic metals and preparing efficient and stable composite electrodes. Lou et al. fabricated a novel Ag@Ni/GF electrode with silver nanoparticles deposited on nickel-coated graphite felt via the self-deposition method [72]. This electrode exhibited high silver content (13.7 mg cm−3), large surface area (26 m2 g−1), and excellent electrocatalytic performance toward chloroacetanilide herbicides like metolachlor. It achieved a dechlorination conversion rate of 96%~99% for an initial concentration of 50 mg L−1 metolachlor, with a current efficiency of 36.6%, and selectively generated 76.1% dechlorinated metolachlor.

4.3. ER of Nitrate

ER technology has emerged as a promising approach for nitrate removal due to its efficiency and tunability. Non-metallic electrodes, particularly boron-doped diamond (BDD) electrodes, have garnered significant interest for their high wear and corrosion resistance. García et al. demonstrated complete nitrate removal from slaughterhouse wastewater within 180 min at 35.7 mA cm−2 using BDD electrodes without additional reagents, primarily converting nitrate to N2 gas [73]. However, the high cost of chemical vapor deposition limits the widespread application of BDD electrodes. Other non-metallic electrodes often face limitations such as insufficient activity, slow kinetics, and low reduction efficiency.
Metal oxide electrodes, particularly those composed of transition metals like Cu, Co, Ag, Ti, Nb, Bi, and Ru, have become a major focus in ER research. These materials offer diverse interfaces, structures, defects, and oxygen vacancies, leading to high nitrate conversion efficiency, Faradaic efficiency, and selective ammonia formation. Wang et al. utilized CuO nanowire arrays, achieving a remarkable nitrate conversion efficiency of 97.0% and an ammonia yield of 0.2449 mmol h−1 cm−2 at an optimal potential of −0.85 V (vs. RHE). Their work also demonstrated a Faradaic efficiency of 95.8% and ammonia selectivity of 81.2% in a 0.5 M Na2SO4 electrolyte [74]. CoOx nanosheets with abundant adsorbed oxygen effectively suppressed hydrogen evolution on the electrode. This approach yielded an ammonia production rate of 82.4 ± 4.8 mg h−1 mg−1 and a Faradaic efficiency of 93.4 ± 3.8% at −0.85 V (vs. RHE) [75]. Jia et al. explored the role of oxygen vacancies in nitrate reduction by heating TiO2 in an H2 atmosphere, generating TiO2-x [76]. These oxygen vacancies facilitated the incorporation of oxygen atoms from nitrate, weakening N-O bonds and minimizing byproduct formation. Their work achieved a nitrate conversion rate of 95.2%, ammonia selectivity of 87.1%, and a yield of 0.045 mmol h−1 mg−1 at −1.6 V (vs. SCE) in 0.5 M Na2SO4 electrolyte solution, with a Faradaic efficiency of 85%. Wan et al. pioneered the use of niobium oxides (NbOx) with oxygen vacancies for nitrate electroreduction to NH3 [77]. Prepared hydrothermally, the NbOx catalyst achieved a Faradaic efficiency of 94.5% and an ammonia production rate of 55.0 μg h−1 mg−1 at −1.15 V (vs. RHE) in a 0.5 M K2SO4 electrolyte.
Recent advancements in ER technology highlight the potential of bimetallic alloy electrodes. These electrodes leverage synergistic effects between constituent metals to enhance performance. Xu et al. designed a low-palladium content alloy with Cu2O using a layered defect method (Figure 4b) [78]. The combination of cavity defects and surface oxygen vacancies in this bimetallic catalyst synergistically promoted nitrate adsorption, weakened N-O bonds, and limited byproduct formation. Their work achieved an ammonia yield of 925.11 µg h−1 mg−1, Faradaic efficiency of 96.56%, and maximum ammonia selectivity of 95.31% at −1.3 V (vs. SCE) in a 0.5 M K2SO4 electrolyte solution. Similarly, Wang et al. electrochemically deposited Cu50Ni50 alloy nanoparticles, achieving a near-perfect Faradaic efficiency of 99 ± 1% at a low potential of −0.15 V (vs. RHE) in a 1.0 M KOH electrolyte [79]. This efficiency improvement was attributed to the d-band center shift of the Cu50Ni50 alloy catalyst toward the Fermi level, enhancing the adsorption of intermediate species. Wang et al. further explored PtxRuy alloy nanoparticles prepared via NaBH4 reduction, demonstrating a Faradaic efficiency exceeding 93% for nitrate reduction to ammonia in a 1.0 M H2SO4 electrolyte solution [80]. This ongoing research on ER technology suggests a promising future for nitrate removal, with bimetallic alloy electrodes offering particularly high efficiency and selectivity.

5. EFP

In 1876, the use of a mixture of H2O2 and Fe2+ to degrade tartaric acid marked the beginning of the Fenton process and its related chemical reactions. Fenton reaction technology offers advantages such as rapid reaction rates, low toxicity, and environmental compatibility, playing a crucial role in water pollution treatment. However, traditional Fenton methods suffer from several notable drawbacks including a narrow effective pH range, substantial generation of hazardous iron sludge when pH exceeds four, significant safety risks associated with H2O2 storage and transportation, uncontrollable reaction rates [81]. These limitations have spurred the development of advanced Fenton technologies, such as the sono-Fenton process, photo-Fenton process, and electro-Fenton process (EFP) [82,83].
EFP, gaining significant traction since the 2000s, offers a more controllable and cost-effective approach. EF processes address the limitations of traditional Fenton by generating H2O2 in situ at the cathode through a two-electron oxygen reduction reaction. This eliminates the need for external H2O2 addition, reducing handling complexity and costs. Moreover, EFP facilitates the continuous regeneration of Fe3⁺ back to Fe2⁺ through electrochemical processes, mitigating iron sludge production. The core principle of EFP relies on the generation of highly oxidizing hydroxyl radicals (•OH) from H2O2 and Fe2⁺, which degrade organic pollutants in wastewater to harmless products like CO2 and H2O (Equation (21)).
H2O2 + Fe2+ +→Fe3+ + •OH + OH
EF processes can be categorized into three types based on the source of Fe2⁺ and H2O2: cathodic electro-Fenton process with oxygen reduction (EFP-H2O2), sacrificial anode electro-Fenton process (EFP-Feox), and cathodic and Fe2⁺ recycling electro-Fenton process (EFP-H2O2-Fere). In the EF-H2O2 process, H2O2 is generated at the cathode while Fe2⁺ needs to be added externally to react with H2O2 and produce •OH radicals (Equation (22), Figure 1d). Conversely, the EF-Feox process utilizes a sacrificial anode to generate Fe2⁺ and requires external H2O2 (Equation (23), Figure 1e). Finally, the EF-H2O2-Fere process offers a self-sustaining system where O2 is reduced to H2O2 at the cathode, and Fe3⁺ is electrochemically regenerated to Fe2⁺, eliminating the need for external addition of both reagents (Equation (24), Figure 1f).
O2 + 2H+ + 2e → H2O2
Fe → Fe2+ + 2e
Fe3+ + e → Fe2+
The versatility of the EFP allows for its application in various scenarios, including treating soil-washing wastewater [84], oxidizing arsenite for safer drinking water [85], and reducing ammonia nitrogen in water [86]. Additionally, EFP effectively reduces organic loads in diverse industrial wastewaters like those from mixed industries [87], pharmaceuticals [88], and mining operations [89]. With its clean reaction mechanisms, environmental compatibility, and minimal secondary pollution, EFP holds immense promise for the future of wastewater treatment. Current research efforts by scientists worldwide are focused on optimizing electrode materials, and reactor configurations, and integrating EF with other treatment methods for even greater efficacy [90].

5.1. H2O2 Electrogeneration in EFP

The electrochemical generation of H2O2 at the cathode is a key factor influencing the efficiency of the EFP [91]. The rate and quantity of H2O2 generated during the reaction process are heavily influenced by the selected cathode material and its configuration within the reactor [92].
The cathode material is a critical factor for efficient electrogeneration of H2O2 in the EFP. It affects the H2O2 production rate, current efficiency, and parallel reactions. Moreover, the performance of the cathode determines the energy requirements and costs. Innovative reactor designs using gas pressure or turbulent conditions of the solution can achieve higher O2 mass transfer and H2O2 generation rates. Innovations in cathode design include gas diffusion electrodes and rotating cylindrical electrodes, which pressurize the system and induce turbulence to enhance O2 mass transfer. Li et al. utilized electrode aeration to provide gas directly within the electrode, facilitating easier gas transport within the electrode and enhancing gas–liquid–solid interfaces with higher O2 mass transfer and utilization rates [93]. Recently, natural air diffusion electrodes with superhydrophobic three-phase interfaces were used to rapidly produce H2O2 (101.67 mg h−1 cm−2) with high oxygen utilization rates (44.5–64.9%) [94]. Cathode substrates and diffusion layers enable natural air diffusion into the reduction reaction interface. Gas diffusion electrodes (GDEs) facilitate the gas diffusion process and accelerate oxygen reduction [95]. However, gas diffusion in GDEs depends on O2 supply (pure O2 or air) and electrochemical reactor design. GDEs currently face challenges such as low H2O2 production rates, low oxygen utilization rates, and high energy consumption. Therefore, optimizing GDEs to become efficient and economical cathode O2 supply methods remains a current challenge [96]. Additionally, pressurized reactors in EFP have emerged as a promising strategy to improve the removal efficiency of target compounds. Poza-Nogueiras et al. demonstrated this application by employing a pressurized jet aerator within an EFP system for the elimination of clofibric acid [96]. Their study utilized iron-containing alginate beads as the catalyst and investigated the impact of current intensity (0.12 A and 0.25 A) and pressure (atmospheric vs. 1 bar gauge). Notably, regardless of the applied current, increasing the pressure from atmospheric to 1 bar gauge significantly enhanced the removal of clofibric acid. This improvement was accompanied by a decrease in the specific energy consumption of the electrochemical cell, suggesting a more efficient treatment process under pressurized conditions [97].
To address these challenges, two approaches have been proposed: (i) improving the composition and performance of cathode materials, and (ii) passive diffusion of dissolved O2 from air, present in wastewater. Significant progress has been made in the first approach using nanotechnology, which modifies, synthesizes, or enhances GDE performance through structural gas transport or accelerated O2 conversion to H2O2 to improve production efficiency, such as catalyst morphology control and heteroatom doping. In the latter approach, passive O2 diffusion from air to GDE is limited by the mass transfer of O2 from solution to the cathode and its solubility in water. Research has shown that using natural air diffusion electrodes (NADEs) can enhance H2O2 production and utilization efficiency [98]. H2O2 production on NADEs reached 158% and 188% when currents of 0.2 A and 1.2 A were applied, respectively, with total energy consumption being only 7.2% and 25.4% of traditional GDEs. These results demonstrate the superiority of NADEs over conventional air electrodes due to their extremely high oxygen mass transfer efficiency and reduced energy consumption.

5.2. Activation of H2O2 to Produce •OH

The second crucial stage in EFP involves the activation of generated H2O2 into highly reactive •OH. The efficiency of H2O2 utilization in the •OH generation process depends on the chosen catalyst and reaction conditions. Catalytic decomposition of H2O2 can be achieved using either homogeneous or heterogeneous catalysts, distinguished by their physical state. Homogeneous catalysts typically use soluble iron salts, offering the advantage of simpler operation. In contrast, heterogeneous catalysts employ iron species or other metal species supported on solid carriers, which are advantageous for their ease of separation from the treated system, facilitating catalyst reusability and recyclability, and making them well-suited for industrial applications.
In homogeneous EFP, the activation of H2O2 primarily relies on externally added Fe2+. For instance, researchers achieved a COD removal rate of 56.4% in municipal solid waste leachate (initial COD concentration ~420 mg L−1) after 4 h using Pd/Al2O3 and Fe2+ catalysts (pH = 3.0), while simultaneously introducing H2 and O2 [99].
Zero-valent metals (ZVMs) such as Fe, Al, Zn, Mg, and Cu have been utilized as heterogeneous catalysts in activating EFP [100]. However, for Fenton/Fenton-like processes using ZVMs, they can only convert O2 into H2O2, thus requiring Fenton/Fenton-like catalysts to decompose in situ generated H2O2 into •OH. Consequently, various composite materials have been developed, such as Zn/Fe and Mg/Fe (Figure 4a) [101]. In the activation of O2 by Fe0, in-situ-generated Fe(II) can induce the decomposition of H2O2. However, in an Al0/O2 system, since Al3+ is the only oxidation state accessible in the solution, the formation of •OH is attributed to the transfer of electrons from Al0 to H2O2 [102].
Transition metal composite materials can also serve as heterogeneous catalysts. Due to their partially filled d-orbitals, transition metals exist in multiple oxidation states, enabling them to catalyze the decomposition of H2O2 into •OH in Fenton/Fenton-like processes. Many transition metal ion complexes activate O2 to degrade pollutants in Fenton/Fenton-like processes through similar reaction mechanisms. In aqueous solutions, dissolved O2 is activated by washed-out Fe2+ to form •O2 intermediates, generating H2O2, which then undergoes a Fenton reaction to produce •OH radicals for pollutant degradation. Similarly, surface-bound Fe2+ can activate O2 in a heterogeneous manner to generate H2O2 and •OH. Goethite and magnetite have also been explored as promising heterogeneous alternatives for EFP [103]. Meijide et al. synthesized a composite containing goethite as the iron source in polyvinyl alcohol–alginate beads. This composite catalyst effectively achieved around 94% TOC removal after 8 h of treatment under optimal conditions (catalyst dosage = 2 g L−1 and pH 3) for the mineralization of pyridinium-based ionic liquids via a heterogeneous EFP. Choe et.al developed an EFP system employing a magnetically coated stainless steel mesh as the cathode for methylene blue removal from wastewater [103]. Their findings demonstrated enhanced catalytic activation of in-situ-generated H2O2 when the steel meshes were coated with magnetite particles, highlighting the potential of magnetite as a cathode material in EFP systems [104].
Furthermore, composite materials composed of ZVM oxides and transition metal oxides have garnered special attention. The slow generation rate of Fe(II) is the rate-limiting step in heterogeneous Fenton reactions, where ZVM oxides can accelerate the reduction of Fe(III) to Fe(II), thereby improving the efficiency of H2O2 utilization. For example, Fe@Fe2O3 composed of a Fe0 core and Fe2O3 shell can activate O2 through a two-electron reduction pathway to generate H2O2, and has been employed for the degradation of bromates and 4-CP in engineering applications. The presence of Fe@Fe2O3 nanowires effectively promotes BrO3 removal, facilitating its reduction to bromide (Br) with the generation of HOBr intermediates. Bromide species analyses and characterizations of the nanowires suggested that surface-bound Fe2+ on core–shell Fe@Fe2O3 could effectively promote the EFP. Meanwhile, molecular oxygen competed for electrons from both Fe2+ and Fe0 to generate reactive oxygen species of •O2, H2O2, and •OH during the BrO3 removal process (Figure 4b) [105]. Mu investigated the effect of phosphate on the activation of molecular oxygen with Fe@Fe2O3 nanowires. The oxygen reduction pathway on Fe@Fe2O3 nanowires was gradually shifted from a four-electron reduction pathway to a sequential one-electron reduction one, along with increasing the phosphate concentration from 0 to 10 mmol L−1. This change in ORR greatly enhanced the molecular oxygen activation and ROS generation performances of Fe@Fe2O3 nanowires, and thus increased their aerobic 4-chlorophenol degradation rate by tenfold (Figure 4c) [106]. Similarly, carboxylated Cu⁰/Fe3O4 nanocomposites exploit the synergistic effect of surface carboxyl groups and Cu⁰ to enhance Fe3⁺/Fe2⁺ conversion within Fe3O4, ensuring a sustained supply of Fe2⁺ for H2O2 decomposition (Figure 4d) [107].

6. Electrocatalytic-Combined Treatment Technology

6.1. Photo-Electrocatalysis

Photo-electrocatalytic (PEC) technology offers a promising approach to environmental remediation by merging the advantages of electrochemical and photocatalytic processes. PEC utilizes a semiconductor photocatalyst to facilitate the separation of photogenerated electron-hole pairs, thereby significantly enhancing catalytic efficiency for various applications. Beyond its high efficiency, PEC technology boasts additional advantages such as low cost, non-toxicity, and remarkable chemical stability. Cerro-Lopez et al. investigated the synergistic effect of PEC technology for pollutant removal from water [108]. The mesostructured PbO2/TiO2 was synthesized and employed in electrocatalytic and photo-electrocatalytic degradation of diclofenac (DCF) with a concentration of 15 ppm in a 0.1 M NaSO4 solution by applying a current of 30 mA cm−2 under different pH conditions. In the electrocatalytic process at pH 6.0, over 50% DCF removal was achieved. Notably, the introduction of UV irradiation during the photo-electrocatalytic experiment led to a significant synergistic effect, enhancing DCF removal efficiency by more than 20% compared to the electrocatalytic process alone. This enhanced degradation was further confirmed by COD analysis, revealing a 76% reduction in COD under PEC conditions compared to a 42% reduction in the electrocatalytic process. Scavenging experiments identified the involvement of photogenerated holes (h+), •OH, and sulfate-based oxidants in the DCF degradation process. Similarly, Umukoro et al. explored the potential of PEC technology for degrading another pharmaceutical contaminant, ciprofloxacin [108]. A pn MoS2-SnO2 heterojunction fixed on expanded graphite (EG) was employed as the photocathode, displaying a remarkable 78.5% PEC removal efficiency for ciprofloxacin, significantly higher than the 55.5% removal achieved by electrocatalysis alone under a current density of 0.010 A·cm−2. The improvement was attributed to the formation of MoS2-SnO2 p-n heterojunctions, which enhanced the light-harvesting capacity and consequently boosted PEC performance [109].

6.2. Microbial Fuel Cells

Electrocatalytic processes combined with microbial fuel cells (MFCs) present another promising strategy for wastewater treatment. MFCs harness the bioelectrochemical activity of microorganisms to decompose organic matter, generating electrical energy as a byproduct. Electrocatalytic components can be further incorporated within the system to facilitate the degradation of pollutants, leading to a synergistic treatment effect. This integrated approach offers several advantages, including enhanced pollutant removal efficiency and the potential for partial energy recovery through MFCs, reducing the overall energy consumption of the treatment process. Hu et al. explored this synergy by employing a Ni/MXene photoelectrochemical cathode within an MFC for chloramphenicol degradation (Figure 5a) [110]. Under optimized conditions (pH = 2), this system achieved a maximum degradation efficiency of 82.62% for an initial chloramphenicol concentration of 30 mg L−1 after 36 h. Yap et al. investigated the impact of aeration and catalysts on both caffeine removal and electricity generation within an MFC (Figure 5b) [111]. Their findings revealed that a CuO/C cathode under aerated conditions achieved a significantly higher caffeine removal rate (52.16%) within 24 h compared to a bare carbon plate (3.41%). Additionally, the CuO/C cathode exhibited superior power and current densities (28.75 mW·m−2 and 253.33 mA·m−2, respectively) under aeration compared to the bare C cathode (9.75 mW·m−2 and 106.67 mA·m−2). Interestingly, connecting the circuit under aerated conditions further enhanced COD removal in the anode chamber, suggesting a complex interplay between aeration, catalyst selection, and MFC operation. Guo et al. explored the method to enhance the electron transfer efficiency and power output of a microbial fuel cell (MFC) by decorating a commercial carbon paper (CP) anode with a composite material of Mo2C/RGO (molybdenum carbide/reduced graphene oxide) [112]. This modification significantly increased the power output density (1747 ± 37.6 mW m−2) compared to the unmodified control (926.8 ± 6.3 mW m−2). The Mo2C/RGO composite also facilitated the formation of a three-dimensional hybrid biofilm and effectively improved the interaction between bacteria and the electrode, leading to a substantial increase in Coulombic efficiency (13.2%) compared to the unmodified control (1.2%).

7. Cost Analysis of Electrocatalytic Technology

The cost-effectiveness of electrocatalytic technology is paramount for its widespread adoption in practical applications. The total operating cost encompasses several factors including electrode preparation and maintenance, electrical energy consumption, electrolyte consumption, and labor costs.
Among these factors, electrical energy consumption is a major hurdle, particularly for processes like electrochemical combustion, hindering their large-scale implementation. In the context of wastewater treatment, the primary cost contributors are typically the electrical energy required for operation and the cost of the electrode materials themselves. Table 8 summarizes operating cost assessments from various studies. Kaur et al. investigated the economic feasibility of Ti/RuO2 electrodes for degrading Ofloxacin (OFLX) antibiotic from synthetic wastewater using the IEO method. Under optimal conditions (pH 2, current 1 A, initial OFLX concentration 50 mg L−1, and NaCl concentration 2 g L−1), an impressive 88.6% degradation of OFLX was achieved within 30 min. The study reported a cost of ~USD 0.50 g−1 of TOC removed for the Ti/RuO2 electrodes. Additionally, the electrical energy consumption at optimal conditions was 541.92 Wh for removing 1 g of TOC in 1 h. Combining the electricity and electrode costs, the total operating cost for TOC removal reached ~USD 0.54 g−1 of TOC removed [113]. Srivastava et al. explored the application of Ti/RuO2 anodes and Fe cathodes for treating real wastewater containing high nitrate concentrations. This system achieved a maximum nitrate reduction efficiency of 46.18% under specific operating conditions (current density = 214.29 A m−2, pH = 7.95, and treatment time = 180 min). Electrical energy consumption for this removal rate was calculated as 108.04 kWh m−3. The cost analysis revealed an electrode cost of 286.25 USD m−3 for the combined anode and cathode setup [114].
In response to these economic challenges, researchers are actively exploring cost-reduction strategies, including choosing methods with low manufacturing costs to produce more efficient electrode catalyst materials, or considering the integration of electrocatalytic technology with biofuel cells and other systems to reduce the consumption of electrical energy.

8. Conclusions

This review focuses on elucidating the mechanisms of electrocatalytic technology for the treatment of refractory pollutants and the current research findings, including the preparation of efficient electrode materials, optimization of operational conditions, and cost analysis. Electrochemical water treatment technology has advanced significantly with the continuous development of electrocatalysis theory. However, it still faces shortcomings and challenges, primarily in two aspects. Firstly, the high cost of electrode materials, limited stability, short lifespan, and high energy consumption remain major concerns. Many new electrodes involve expensive materials such as rare earth elements, precious metals, carbon nanotubes, and graphene. The complex preparation methods restrict their widespread application, largely confining electrode development to the experimental research stage. Secondly, the electrocatalytic degradation mechanism remains imperfect. The degradation of organic pollutants in water via electrocatalysis is a complex process. Actual industrial wastewater contains various substances including organic compounds, heavy metal ions, and inorganic salts, complicating the electrocatalytic process. The generation and transformation processes of active radicals during reactions, as well as the reactions of organic chemical wastewater on electrode surfaces and in solution, are still unclear. These uncertainties limit the application of electrocatalytic technology in degrading organic chemical wastewater.
Considering current research and application trends, the future research directions of electrochemical water treatment technology will mainly focus on three aspects. (1) Design and development of electrode materials: Efficient and stable electrode materials are crucial for electrochemical oxidation water treatment technology. Developing electrodes based on optimized electron transfer and micro-interface regulation will directly impact the efficiency of electrochemical oxidation water treatment. Moreover, developing cost-effective electrodes that can be scaled up for practical engineering applications is a current and future research priority. (2) Optimization of operating parameters: To maximize the performance of anode materials, it is essential to explore key operational parameters during degradation processes. Factors such as supporting electrolytes, pH, properties of organic substances, and mass transfer kinetics within reactors are critical for understanding the dynamics of electrocatalytic oxidation processes. Additionally, establishing kinetic models of substance degradation processes will guide the targeted design of anode materials. (3) Recovery of resources and energy: Wastewater pollutants contain significant chemical energy. Future directions in water pollution control technology involve recovering this chemical energy to simultaneously degrade pollutants and generate energy. By controlling electrode reactions and finely tuning micro-interfaces, efficient separation and targeted conversion of pollutants can be achieved, advancing toward resource recovery and energy utilization in water pollution control technologies.

Author Contributions

Writing—original draft preparation, X.Z., J.Y. and J.G.; review and editing, supervision, W.X. and M.K.H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research is financially supported by the National Natural Science Foundation of China (No. 22006007).

Acknowledgments

W. Xiong is grateful for support from the Key Laboratory of Industrial Ecology and Environmental Engineering (Ministry of Education) and Ability R&D Energy Research Centre, School of Energy and Environment, City University of Hong Kong.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, Y.; Wang, Y. Study on the treatment of actual waste-water by CWAO method. IOP Conf. Ser. Earth Environ. Sci. 2020, 514, 032044. [Google Scholar] [CrossRef]
  2. Tian, K.; Hu, L.M.; Li, L.T.; Zheng, Q.Z.; Xin, Y.J.; Zhang, G.S. Recent advances in persulfate-based advanced oxidation processes for organic wastewater treatment. Chin. Chem. Lett. 2022, 33, 4461–4477. [Google Scholar] [CrossRef]
  3. Meng, Z.Y.; Wilsey, M.K.; Cox, C.P.; Müller, A.M. Complete electrocatalytic defluorination of perfluorooctane sulfonate in aqueous solution with nonprecious materials. J. Catal. 2024, 431, 115403. [Google Scholar] [CrossRef]
  4. Gu, H.B.; Xie, W.H.; Du, A.I.; Pan, D.; Guo, Z.H. Overview of electrocatalytic treatment of antibiotic pollutants in wastewater. Catal. Rev. 2023, 65, 569–619. [Google Scholar] [CrossRef]
  5. Sun, J.Y.; Xu, L.P.; Shi, Z.X.; Zhao, Q.Q.; Wang, H.B.; Gan, T. Morphology-tunable hollow Mn2O3 nanostructures: Highly efficient electrocatalysts and their electrochemical sensing for phenolic endocrine disruptors via toughening of graphene oxide. Sens. Actuators B-Chem. 2021, 327, 128889. [Google Scholar] [CrossRef]
  6. Shen, W.Z.; Guo, Q.J.; Zhang, Y.S.; Liu, Y.; Zheng, J.T.; Cheng, J.; Fan, J. The effect of activated carbon fiber structure and loaded copper, cobalt, silver on the adsorption of dichloroethylene. Colloids Surf. A 2006, 273, 147–153. [Google Scholar] [CrossRef]
  7. Zhang, J.Q.; Chen, H.J.; Duan, X.G.; Sun, H.Q.; Wang, S.B. Photothermal catalysis: From fundamentals to practical applications. Mater. Today 2023, 68, 234–253. [Google Scholar] [CrossRef]
  8. Nie, H.Y.; Nie, M.Q.; Diwu, Z.J.; Wang, L.; Qiao, Q.; Zhang, B.; Yang, X.F. Homogeneously catalytic oxidation of phenanthrene by the reaction of extracellular secretions of pyocyanin and Nicotinamide Adenine Dinucleotide. Environ. Res. 2020, 191, 110159. [Google Scholar] [CrossRef]
  9. Martínez-Huitle, C.A.; Rodrigo, M.A.; Sirés, I.; Scialdone, O. A critical review on latest innovations and future challenges of electrochemical technology for the abatement of organics in water. Appl. Catal. B-Environ. 2023, 328, 122430. [Google Scholar] [CrossRef]
  10. Xu, J.; Yuan, D.F.; Yang, F.; Mei, D.; Zhang, Z.B.; Chen, Y.X. On the mechanism of the direct pathway for formic acid oxidation at a Pt(111) electrode. Phys. Chem. Chem. Phys. 2013, 15, 4367–4376. [Google Scholar] [CrossRef]
  11. Kapalka, A.; Joss, L.; Anglada, A.; Comninellis, C.; Udert, K.M. Direct and mediated electrochemical oxidation of ammonia on boron-doped diamond electrode. Electrochem. Commun. 2010, 12, 1714–1717. [Google Scholar] [CrossRef]
  12. Tsai, M.H.; Juang, Y.; Hu, C.C.; Hua, L.C.; Mahata, B.K.; Huang, C.H. The direct electrocatalytic oxidation of ammonia by copper-deposited nickel foam catalysts. Electrochim. Acta 2023, 446, 142130. [Google Scholar] [CrossRef]
  13. Liu, Z.C.; Zhang, G.; Lan, H.C.; Liu, H.J.; Qu, J.H. Optimization of a Hierarchical Porous-Structured Reactor to Mitigate Mass Transport Limitations for Efficient Electrocatalytic Ammonia Oxidation through a Three-Electron-Transfer Pathway. Environ. Sci. Technol. 2021, 55, 12596–12606. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, S.Z.; Mao, J.; Zhang, Y.Q.; Wang, S.R.; Huo, M.J.; Guo, H.Y. Sensitive electrochemical detection of enrofloxacin in eggs based on carboxylated multi-walled carbon nanotubes-reduced graphene oxide nanocomposites: Molecularly imprinted recognition versus direct electrocatalytic oxidation. Food Chem. 2023, 413, 135579. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, X.Y.; Wang, J.L. Selective oxidation of ammonium to nitrogen gas by advanced oxidation processes: Reactive species and oxidation mechanisms. J. Environ. Chem. Eng. 2023, 11, 110263. [Google Scholar] [CrossRef]
  16. Ghani, A.A.; Maile, N.; Tahir, K.; Kim, B.; Lim, Y.; Jang, J.; Lee, D.S. Electrocatalytic oxidation of antidiabetic drug metformin adsorbed on intercalated MXene. Chemosphere 2022, 307, 135767. [Google Scholar] [CrossRef]
  17. Pan, Z.L.; Xu, S.; Xin, H.; Yuan, Y.; Xu, R.S.; Wang, P.C.; Yan, X.Q.; Fan, X.F.; Song, C.W.; Wang, T.H. High performance polypyrrole coated carbon-based electrocatalytic membrane for organic contaminants removal from aqueous solution. J. Colloid Interface Sci. 2022, 626, 283–295. [Google Scholar] [CrossRef] [PubMed]
  18. Li, S.W.; Lin, Y.Z.; Zhu, S.Y.; Liu, G. Electrocatalytic degradation of sulfamethylthiadiazole by GAC@Ni/Fe three-dimensional particle electrode. Environ. Sci. Pollut. Res. 2022, 29, 57112–57126. [Google Scholar] [CrossRef]
  19. Almomani, F.; Bhosale, R.; Khraisheh, M.; Kumar, A.; Tawalbeh, M. Electrochemical oxidation of ammonia on nickel oxide nanoparticles. Int. J. Hydrogen Energy 2020, 45, 10398–10408. [Google Scholar] [CrossRef]
  20. Ma, C.B.; Zhang, Y.Q. Spinel CuxCo1-xMn2O4 electrode for effectively cleaning organic wastewater via electrocatalytic oxidation. Sep. Purif. Technol. 2021, 258, 118024. [Google Scholar] [CrossRef]
  21. Bu, J.Q.; Deng, Z.W.; Liu, H.; Li, T.H.; Yang, Y.J.; Zhong, S. The degradation of sulfamilamide wastewater by three-dimensional electrocatalytic oxidation system composed of activated carbon bimetallic particle electrode. J. Clean. Prod. 2021, 324, 129256. [Google Scholar] [CrossRef]
  22. Sun, Y.J.; Zhu, S.C.; Sun, W.Q.; Zheng, H.L. Degradation of high-chemical oxygen demand concentration pesticide wastewater by 3D electrocatalytic oxidation. J. Environ. Chem. Eng. 2019, 7, 103276. [Google Scholar] [CrossRef]
  23. Yang, C.; Li, X.Y.; Lin, L. Fabrication of a SnO2-Sb nano-pin array anode for efficient electrocatalytic oxidation of bisphenol A in wastewater. J. Hazard. Mater. 2023, 444, 130444. [Google Scholar] [CrossRef]
  24. Niu, Y.X.; Yin, Y.; Xu, R.Y.; Yang, Z.N.; Wang, J.; Xu, D.; Yuan, Y.; Han, J.L.; Wang, H. Electrocatalytic oxidation of low concentration cefotaxime sodium wastewater using Ti/SnO2-RuO2 electrode: Feasibility analysis and degradation mechanism. Chemosphere 2022, 297, 134146. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, L.J.; Zhang, Y.Q.; Ma, C.B. Perovskites SrxLa1-xMnyCo1-yO3-δ coated on Ti as stable non-noble anode for efficient electrocatalytic oxidation of organic wastewater containing ammonia nitrogen. Chem. Eng. J. 2020, 393, 124514. [Google Scholar] [CrossRef]
  26. Yang, C.; Shang, S.S.; Li, X.Y. Oxygen-vacancy-enriched substrate-less SnOx/La-Sb anode for high-performance electrocatalytic oxidation of antibiotics in wastewater. J. Hazard. Mater. 2022, 436, 129212. [Google Scholar] [CrossRef] [PubMed]
  27. Tian, S.C.; Tu, Y.M.; Chen, J.J.; Shao, G.Y.; Zhou, Z.Y.; Ren, Z.Q. Persulfate enhanced electrochemical oxidation of phenol with CuFe2O4/ACF (activated carbon fibers) cathode. Sep. Purif. Technol. 2021, 279, 119727. [Google Scholar] [CrossRef]
  28. Cai, J.J.; Zhou, M.H.; Liu, Y.; Savall, A.; Serrano, K.G. Indirect electrochemical oxidation of 2,4-dichlorophenoxyacetic acid using electrochemically-generated persulfate. Chemosphere 2018, 204, 163–169. [Google Scholar] [CrossRef]
  29. Ghanbari, F.; Wang, Q.L.; Hassani, A.; Waclawek, S.; Rodríguez-Chueca, J.; Lin, K.Y.A. Electrochemical activation of peroxides for treatment of contaminated water with landfill leachate: Efficacy, toxicity and biodegradability evaluation. Chemosphere 2021, 279, 130610. [Google Scholar] [CrossRef] [PubMed]
  30. Jin, X.; Liu, M.W.; Li, S.Y.; Yang, C.; Hu, S.Y.; Shang, Y.B.; Li, K.Q.; Bai, X.; Xu, L.; Shi, X.; et al. Novel use of the electrocatalytic oxidation process with Fe-DSA dual anode configuration for leachate treatment: Significance of in situ ferrate generation. J. Clean. Prod. 2024, 457, 142490. [Google Scholar] [CrossRef]
  31. Salvestrini, S.; Fenti, A.; Chianese, S.; Iovino, P.; Musmarra, D. Electro-Oxidation of Humic Acids Using Platinum Electrodes: An Experimental Approach and Kinetic Modelling. Water 2020, 12, 2250. [Google Scholar] [CrossRef]
  32. Deng, Y.; Chen, N.; Feng, C.P.; Chen, F.X.; Wang, H.S.; Kuang, P.J.; Feng, Z.Y.; Liu, T.; Gao, Y.; Hu, W.W. Treatment of organic wastewater containing nitrogen and chlorine by combinatorial electrochemical system: Taking biologically treated landfill leachate treatment as an example. Chem. Eng. J. 2019, 364, 349–360. [Google Scholar] [CrossRef]
  33. Fu, R.; Zhang, P.S.; Jiang, Y.X.; Sun, L.; Sun, X.H. Wastewater treatment by anodic oxidation in electrochemical advanced oxidation process: Advance in mechanism, direct and indirect oxidation detection methods. Chemosphere 2023, 311, 136993. [Google Scholar] [CrossRef]
  34. Cho, S.; Kim, C.; Hwang, I. Electrochemical degradation of ibuprofen using an activated-carbon-based continuous-flow three-dimensional electrode reactor (3DER). Chemosphere 2020, 259, 127382. [Google Scholar] [CrossRef] [PubMed]
  35. Arias, A.N.; de Mello, R.; Lobato, J.; Motheo, A.J.; Rodrigo, M.A. Electrolytic removal of volatile organic compounds: Keys to understand the process. J. Electroanal. Chem. 2022, 912, 116259. [Google Scholar] [CrossRef]
  36. Maezono, T.; Tokumura, M.; Sekine, M.; Kawase, Y. Hydroxyl radical concentration profile in photo-Fenton oxidation process: Generation and consumption of hydroxyl radicals during the discoloration of azo-dye Orange II. Chemosphere 2011, 82, 1422–1430. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, L.L.; Fu, Y.; Li, Q.C.; Wang, Z.H. EPR Evidence for Mechanistic Diversity of Cu(II)/Peroxygen Oxidation Systems by Tracing the Origin of DMPO Spin Adducts. Environ. Sci. Technol. 2022, 56, 8796–8806. [Google Scholar] [CrossRef]
  38. Chen, L.C.; Lei, C.J.; Li, Z.J.; Yang, B.; Zhang, X.W.; Lei, L.C. Electrochemical activation of sulfate by BDD anode in basic medium for efficient removal of organic pollutants. Chemosphere 2018, 210, 516–523. [Google Scholar] [CrossRef] [PubMed]
  39. Saha, P.; Wang, J.M.; Zhou, Y.N.; Carlucci, L.; Jeremiasse, A.W.; Rijnaarts, H.H.M.; Bruning, H. Effect of electrolyte composition on electrochemical oxidation: Active sulfate formation, benzotriazole degradation, and chlorinated by-products distribution. Environ. Res. 2022, 211, 113057. [Google Scholar] [CrossRef]
  40. Wang, L.Y.; Di, J.; Nie, J.; Ma, G.P. Multicomponent Doped Sugar-Coated Nanofibers for Peroxymonosulfate Activation. ACS Appl. Nano Mater. 2019, 2, 6998–7007. [Google Scholar] [CrossRef]
  41. Gao, H.Y.; Huang, C.H.; Mao, L.; Shao, B.; Shao, J.; Yan, Z.Y.; Tang, M.; Zhu, B.Z. First Direct and Unequivocal Electron Spin Resonance Spin-Trapping Evidence for pH-Dependent Production of Hydroxyl Radicals from Sulfate Radicals. Environ. Sci. Technol. 2020, 54, 14046–14056. [Google Scholar] [CrossRef]
  42. Zhang, H.Y.; Di, H.C.; Jiang, Z.W.; Sun, F.Y.; Pan, C.T.; Yang, J.H.; Zhang, F.Y.; Bai, X. Enhanced degradation of nitrogen-containing disinfection by-products via radical attack in ZnAl particle electrode system: Mechanism and density functional theory calculations. J. Clean. Prod. 2024, 449, 141720. [Google Scholar] [CrossRef]
  43. Li, A.Y.; Weng, J.Q.; Yan, X.M.; Li, H.; Shi, H.B.; Wu, X.D. Electrochemical oxidation of acid orange 74 using Ru, IrO2, PbO2, and boron doped diamond anodes: Direct and indirect oxidation. J. Electroanal. Chem. 2021, 898, 115622. [Google Scholar] [CrossRef]
  44. Thiam, A.; Brillas, E.; Centellas, F.; Cabot, P.L.; Sirés, I. Electrochemical reactivity of Ponceau 4R (food additive E124) in different electrolytes and batch cells. Electrochim. Acta 2015, 173, 523–533. [Google Scholar] [CrossRef]
  45. Nidheesh, P.V.; Zhou, M.H.; Oturan, M.A. An overview on the removal of synthetic dyes from water by electrochemical advanced oxidation processes. Chemosphere 2018, 197, 210–227. [Google Scholar] [CrossRef] [PubMed]
  46. Ganiyu, S.O.; Martínez-Huitle, C.A. Nature, Mechanisms and Reactivity of Electrogenerated Reactive Species at Thin-Film Boron-Doped Diamond (BDD) Electrodes During Electrochemical Wastewater Treatment. Chemelectrochem 2019, 6, 2379–2392. [Google Scholar] [CrossRef]
  47. Dionisio, D.; Rodrigo, M.A.; Motheo, A.J. Electrochemical degradation of a methyl paraben and propylene glycol mixture: Interference effect of competitive oxidation and pH stability. Chemosphere 2022, 287, 132229. [Google Scholar] [CrossRef] [PubMed]
  48. Zheng, W.X.; Zhu, L.Y.; Liang, S.; Ye, J.S.; Yang, X.; Lei, Z.C.; Yan, Z.; Li, Y.D.; Wei, C.H.; Feng, C.H. Discovering the Importance of ClO• in a Coupled Electrochemical System for the Simultaneous Removal of Carbon and Nitrogen from Secondary Coking Wastewater Effluent. Environ. Sci. Technol. 2020, 54, 9015–9024. [Google Scholar] [CrossRef]
  49. Wang, X.X.; Sun, M.; Zhao, Y.M.; Wang, C.; Ma, W.; Wong, M.S.; Elimelech, M. In Situ Electrochemical Generation of Reactive Chlorine Species for Efficient Ultrafiltration Membrane Self-Cleaning. Environ. Sci. Technol. 2020, 54, 6997–7007. [Google Scholar] [CrossRef]
  50. Song, R.X.; Wang, H.; Zhang, M.; Liu, Y.Y.; Meng, X.F.; Zhai, S.J.; Wang, C.C.; Gong, T.; Wu, Y.L.; Jiang, X.W.; et al. Near-Infrared Light-Triggered Chlorine Radical (•Cl) Stress for Cancer Therapy. Angew. Chem. Int. Ed. 2020, 59, 21032–21040. [Google Scholar] [CrossRef] [PubMed]
  51. Li, L.; Li, J.H.; Fang, F.; Zhang, Y.; Zhou, T.S.; Zhou, C.H.; Bai, J.; Zhou, B.X. Efficient H2O2 production from urine treatment based on a self-biased WO3/TiO2-Si PVC photoanode and a WO3/CMK-3 cathode. Appl. Catal. B-Environ. 2023, 333, 122776. [Google Scholar] [CrossRef]
  52. Wang, Y.Y.; Ma, B.; Zhao, J.; Tang, Z.Y.; Li, W.X.; He, C.; Xia, D.H.; Linden, K.G.; Yin, R. Rapid Inactivation of Fungal Spores in Drinking Water by Far-UVC Photolysis of Free Chlorine. Environ. Sci. Technol. 2023, 57, 21876–21887. [Google Scholar] [CrossRef]
  53. Yan, Z.; Dai, Z.R.; Zheng, W.X.; Lei, Z.C.; Qiu, J.W.; Kuang, W.J.; Huang, W.J.; Feng, C.H. Facile ammonium oxidation to nitrogen gas in acid wastewater by in situ photogenerated chlorine radicals. Water Res. 2021, 205, 117678. [Google Scholar] [CrossRef]
  54. Li, T.; Jiang, Y.; An, X.Q.; Liu, H.J.; Hu, C.; Qu, J.H. Transformation of humic acid and halogenated byproduct formation in UV-chlorine processes. Water Res. 2016, 102, 421–427. [Google Scholar] [CrossRef]
  55. Wang, Q.H.; Li, T.F.; Yang, C.; Chen, M.H.; Guan, A.X.; Yang, L.; Li, S.; Lv, X.M.; Wang, Y.H.; Zheng, G.F. Electrocatalytic Methane Oxidation Greatly Promoted by Chlorine Intermediates. Angew. Chem. Int. Ed. 2021, 60, 17398–17403. [Google Scholar] [CrossRef]
  56. Li, X.; Ma, S.Q.; Hu, Y.Y.; Zhang, C.Y.; Xiao, C.; Shi, Y.Y.; Liu, J.Y.; Cheng, J.H.; Chen, Y.C. Degradation of norfloxacin in a heterogeneous electro-Fenton like system coupled with sodium chloride as the electrolyte. Chem. Eng. J. 2023, 473, 145202. [Google Scholar] [CrossRef]
  57. Tran, T.K.; Chiu, K.F.; Lin, C.Y.; Leu, H.J. Electrochemical treatment of wastewater: Selectivity of the heavy metals removal process. Int. J. Hydrogen Energy 2017, 42, 27741–27748. [Google Scholar] [CrossRef]
  58. Yang, X.; Liu, L.H.; Wang, Y.; Lu, T.; Wang, Z.W.; Qiu, G.H. Sustainable and reagent-free cathodic precipitation for high-efficiency removal of heavy metals from soil leachate*. Environ. Pollut. 2023, 320, 121002. [Google Scholar] [CrossRef]
  59. Feng, Y.; Yang, S.M.; Xia, L.; Wang, Z.W.; Suo, N.; Chen, H.; Long, Y.Y.; Zhou, B.; Yu, Y.Z. In-situ ion exchange electrocatalysis biological coupling (i-IEEBC) for simultaneously enhanced degradation of organic pollutants and heavy metals in electroplating wastewater. J. Hazard. Mater. 2019, 364, 562–570. [Google Scholar] [CrossRef] [PubMed]
  60. Li, J.J.; Wang, H.; Qi, Z.Y.; Ma, C.; Zhang, Z.H.; Zhao, B.; Wang, L.; Zhang, H.W.; Chong, Y.T.; Chen, X.; et al. Kinetics and mechanisms of electrocatalytic hydrodechlorination of diclofenac on Pd-Ni/PPy-rGO/Ni electrodes. Appl. Catal. B-Environ. 2020, 268, 118696. [Google Scholar] [CrossRef]
  61. Almassi, S.; Samonte, P.R.V.; Li, Z.; Xu, W.Q.; Chaplin, B.P. Mechanistic Investigation of Haloacetic Acid Reduction Using Carbon-Ti4O7 Composite Reactive Electrochemical Membranes. Environ. Sci. Technol. 2020, 54, 1982–1991. [Google Scholar] [CrossRef]
  62. Chen, Y.J.; Feng, C.; Wang, W.H.; Liu, Z.; Li, J.X.; Liu, C.G.; Pan, Y.; Liu, Y.Q. Electronic structure engineering of bimetallic Pd-Au alloy nanocatalysts for improving electrocatalytic hydrodechlorination performance. Sep. Purif. Technol. 2022, 289, 120731. [Google Scholar] [CrossRef]
  63. Yao, F.B.; Jia, M.C.; Yang, Q.; Chen, F.; Zhong, Y.; Chen, S.J.; He, L.; Pi, Z.J.; Hou, K.J.; Wang, D.B.; et al. Highly selective electrochemical nitrate reduction using copper phosphide self-supported copper foam electrode: Performance, mechanism, and application. Water Res. 2021, 193, 116881. [Google Scholar] [CrossRef]
  64. Yu, Y.; Wang, C.H.; Yu, Y.F.; Wang, Y.T.; Zhang, B. Promoting selective electroreduction of nitrates to ammonia over electron-deficient Co modulated by rectifying Schottky contacts. Sci. China Chem. 2020, 63, 1469–1476. [Google Scholar] [CrossRef]
  65. Guo, Y.; Zhang, R.; Zhang, S.C.; Zhao, Y.W.; Yang, Q.; Huang, Z.D.; Dong, B.B.; Zhi, C.Y. Pd doping-weakened intermediate adsorption to promote electrocatalytic nitrate reduction on TiO2 nanoarrays for ammonia production and energy supply with zinc-nitrate batteries. Energy Environ. Sci. 2021, 14, 3938–3944. [Google Scholar] [CrossRef]
  66. Qin, H.H.; Liu, X.R.; Liu, X.Y.; Zhao, H.Y.; Mao, S. Highly Selective Electrocatalytic CuEDTA Reduction by MoS2 Nanosheets for Efficient Pollutant Removal and Simultaneous Electric Power Output. Nano-Micro Lett. 2023, 15, 193. [Google Scholar] [CrossRef] [PubMed]
  67. Pan, M.L.; Zhang, C.; Wang, J.; Chew, J.W.; Gao, G.D.; Pan, B.C. Multifunctional Piezoelectric Heterostructure of BaTiO3@Graphene: Decomplexation of Cu-EDTA and Recovery of Cu. Environ. Sci. Technol. 2019, 53, 8342–8351. [Google Scholar] [CrossRef]
  68. Jiang, W.L.; Shen, J.C.; Peng, Z.Y.; Wu, G.Y.; Yin, G.Q.; Shi, X.L.; Yang, H.B. Controllable synthesis of ultrasmall Pd nanocatalysts templated by supramolecular coordination cages for highly efficient reductive dehalogenation. J. Mater. Chem. A 2020, 8, 12097–12105. [Google Scholar] [CrossRef]
  69. Yao, Q.F.; Zhou, X.F.; Xiao, S.Z.; Chen, J.B.; Abdelhafeez, I.A.; Yu, Z.J.; Chu, H.Q.; Zhang, Y.L. Amorphous nickel phosphide as a noble metal-free cathode for electrochemical dechlorination. Water Res. 2019, 165, 114930. [Google Scholar] [CrossRef]
  70. Liu, H.L.; Han, J.L.; Yuan, J.L.; Liu, C.B.; Wang, D.; Liu, T.; Liu, M.J.; Luo, J.M.; Wang, A.J.; Crittenden, J.C. Deep Dehalogenation of Florfenicol Using Crystalline CoP Nanosheet Arrays on a Ti Plate via Direct Cathodic Reduction and Atomic H. Environ. Sci. Technol. 2019, 53, 11932–11940. [Google Scholar] [CrossRef]
  71. King, J.F.; Mitch, W.A. Electrochemical Reduction of Halogenated Alkanes and Alkenes Using Activated Carbon-Based Cathodes. Environ. Sci. Technol. 2022, 56, 17965–17976. [Google Scholar] [CrossRef]
  72. Lou, Y.Y.; He, W.Y.; Verlato, E.; Musiani, M.; Floner, D.; Fourcade, F.; Amrane, A.; Li, C.Y.; Tian, Z.Q.; Merdrignac-Conanec, O.; et al. Ni-coated graphite felt modified with Ag nanoparticles: A new electrode material for electro-reductive dechlorination. J. Electroanal. Chem. 2019, 849, 113357. [Google Scholar] [CrossRef]
  73. García-Gómez, C.; Drogui, P.; Zaviska, F.; Seyhi, B.; Gortáres-Moroyoqui, P.; Buelna, G.; Neira-Sáenz, C.; Estrada-Alvarado, M.; Ulloa-Mercado, R.G. Experimental design methodology applied to electrochemical oxidation of carbamazepine using Ti/PbO2 and Ti/BDD electrodes. J. Electroanal. Chem. 2014, 732, 1–10. [Google Scholar] [CrossRef]
  74. Wang, Y.; Zhou, W.; Jia, R.; Yu, Y.; Zhang, B. Unveiling the Activity Origin of a Copper-based Electrocatalyst for Selective Nitrate Reduction to Ammonia. Angew. Chem. Int. Edit 2020, 59, 5350–5354. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, J.; Cai, C.; Wang, Y.A.; Yang, X.M.; Wu, D.J.; Zhu, Y.M.; Li, M.H.; Gu, M.; Shao, M.H. Electrocatalytic Reduction of Nitrate to Ammonia on Low-Cost Ultrathin CoOx Nanosheets. ACS Catal. 2021, 11, 15135–15140. [Google Scholar] [CrossRef]
  76. Jia, R.R.; Wang, Y.T.; Wang, C.H.; Ling, Y.F.; Yu, Y.F.; Zhang, B. Boosting Selective Nitrate Electroreduction to Ammonium by Constructing Oxygen Vacancies in TiO2. ACS Catal. 2020, 10, 3533–3540. [Google Scholar] [CrossRef]
  77. Wan, X.F.; Guo, W.W.; Dong, X.; Wu, H.H.; Sun, X.F.; Chu, M.G.; Han, S.T.; Zhai, J.X.; Xia, W.; Jia, S.Q.; et al. Boosting nitrate electroreduction to ammonia on NbOx via constructing oxygen vacancies. Green Chem. 2022, 24, 1090–1095. [Google Scholar] [CrossRef]
  78. Xu, Y.; Ren, K.L.; Ren, T.L.; Wang, M.Z.; Wang, Z.Q.; Li, X.N.; Wang, L.; Wang, H.J. Ultralow-content Pd in-situ incorporation mediated hierarchical defects in corner-etched Cu2O octahedra for enhanced electrocatalytic nitrate reduction to ammonia. Appl. Catal. B-Environ. 2022, 306, 121094. [Google Scholar] [CrossRef]
  79. Wang, Y.H.; Xu, A.; Wang, Z.Y.; Huang, L.S.; Li, J.; Li, F.W.; Wicks, J.; Luo, M.C.; Nam, D.H.; Tan, C.S.; et al. Enhanced Nitrate-to-Ammonia Activity on Copper-Nickel Alloys via Tuning of Intermediate Adsorption. J. Am. Chem. Soc. 2020, 142, 5702–5708. [Google Scholar] [CrossRef]
  80. Wang, Z.X.; Young, S.D.; Goldsmith, B.R.; Singh, N. Increasing electrocatalytic nitrate reduction activity by controlling adsorption through PtRu alloying. J. Catal. 2021, 395, 143–154. [Google Scholar] [CrossRef]
  81. Oturan, M.A.; Aaron, J.J. Advanced Oxidation Processes in Water/Wastewater Treatment: Principles and Applications. A Review. Crit. Rev. Environ. Sci. Technol. 2014, 44, 2577–2641. [Google Scholar] [CrossRef]
  82. Heidari, Z.; Pelalak, R.; Alizadeh, R.; Oturan, N.; Shirazian, S.; Oturan, M.A. Application of Mineral Iron-Based Natural Catalysts in Electro-Fenton Process: A Comparative Study. Catalysts 2021, 11, 57. [Google Scholar] [CrossRef]
  83. Zhang, Y.; Zhou, M.H. A critical review of the application of chelating agents to enable Fenton and 1 Fenton-like reactions at high pH values. J. Hazard. Mater. 2019, 362, 436–450. [Google Scholar] [CrossRef] [PubMed]
  84. Ganiyu, S.O.; Martínez-Huitle, C.A.; Rodrigo, M.A. Renewable energies driven electrochemical wastewater/soil decontamination technologies: A critical review of fundamental concepts and applications. Appl. Catal. B-Environ. 2020, 270, 118857. [Google Scholar] [CrossRef]
  85. Nidheesh, P.V.; Babu, D.S.; Dasgupta, B.; Behara, P.; Ramasamy, B.; Kumar, M.S. Treatment of Arsenite-Contaminated Water by Electrochemical Advanced Oxidation Processes. Chemelectrochem 2020, 7, 2418–2423. [Google Scholar] [CrossRef]
  86. Menon, P.; Singh, T.S.A.; Pani, N.; Nidheesh, P.V. Electro-Fenton assisted sonication for removal of ammoniacal nitrogen and organic matter from dye intermediate industrial wastewater. Chemosphere 2021, 269, 128739. [Google Scholar] [CrossRef] [PubMed]
  87. Popat, A.; Nidheesh, P.V.; Singh, T.S.A.; Kumar, M.S. Mixed industrial wastewater treatment by combined electrochemical advanced oxidation and biological processes. Chemosphere 2019, 237, 124419. [Google Scholar] [CrossRef] [PubMed]
  88. Ganzenko, O.; Trellu, C.; Oturan, N.; Huguenot, D.; Péchaud, Y.; van Hullebusch, E.D.; Oturan, M.A. Electro-Fenton treatment of a complex pharmaceutical mixture: Mineralization efficiency and biodegradability enhancement. Chemosphere 2020, 253, 126659. [Google Scholar] [CrossRef] [PubMed]
  89. Olvera-Vargas, H.; Gore-Datar, N.; Garcia-Rodriguez, O.; Mutnuri, S.; Lefebvre, O. Electro-Fenton treatment of real pharmaceutical wastewater paired with a BDD anode: Reaction mechanisms and respective contribution of homogeneous and heterogeneous •OH. Chem. Eng. J. 2021, 404, 126524. [Google Scholar] [CrossRef]
  90. Nidheesh, P.V.; Ganiyu, S.O.; Martínez-Huitle, C.A.; Mousset, E.; Olvera-Vargas, H.; Trellu, C.; Zhou, M.H.; Oturan, M.A. Recent advances in electro-Fenton process and its emerging applications. Crit. Rev. Environ. Sci. Technol. 2023, 53, 887–913. [Google Scholar] [CrossRef]
  91. Ganiyu, S.O.; Martínez-Huitle, C.A.; Oturan, M.A. Electrochemical advanced oxidation processes for wastewater treatment: Advances in formation and detection of reactive species and mechanisms. Curr. Opin. Electrochem. 2021, 27, 100678. [Google Scholar] [CrossRef]
  92. Zhou, W.; Xie, L.; Gao, J.H.; Nazari, R.; Zhao, H.Q.; Meng, X.X.; Sun, F.; Zhao, G.B.; Ma, J. Selective H2O2 electrosynthesis by O-doped and transition-metal-O-doped carbon cathodes via O2 electroreduction: A critical review. Chem. Eng. J. 2021, 410, 128368. [Google Scholar] [CrossRef]
  93. Li, D.; Zheng, T.; Liu, Y.L.; Hou, D.; Yao, K.Y.K.; Zhang, W.; Song, H.R.; He, H.Y.; Shi, W.; Wang, L.; et al. A novel Electro-Fenton process characterized by aeration from inside a graphite felt electrode with enhanced electrogeneration of H2O2 and cycle of Fe3+/Fe2+. J. Hazard. Mater. 2020, 396, 122591. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, Q.Z.; Zhou, M.H.; Ren, G.B.; Li, Y.W.; Li, Y.C.; Du, X.D. Highly efficient electrosynthesis of hydrogen peroxide on a superhydrophobic three-phase interface by natural air diffusion. Nat. Commun. 2020, 11, 1731. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, W.; Li, Y.C.; Li, Y.W.; Zhou, M.H.; Arotiba, O.A. Electro-Fenton and photoelectro-Fenton degradation of sulfamethazine using an active gas diffusion electrode without aeration. Chemosphere 2020, 250, 126177. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, G.; Yao, Y.C.; Tang, K.; Wang, G.C.; Zhang, W.J.; Zhang, Y.F.; Andersen, H.R. Cost-efficient microbial electrosynthesis of hydrogen peroxide on a facile-prepared floating electrode by entrapping oxygen. Bioresour. Technol. 2021, 342, 125995. [Google Scholar] [CrossRef] [PubMed]
  97. Poza-Nogueiras, V.; Moratalla, A.; Pazos, M.; Sanromán, A.; Sáez, C.; Rodrigo, M.A. Exploring the pressurized heterogeneous electro-Fenton process and modelling the system. Chem. Eng. J. 2022, 431, 133280. [Google Scholar] [CrossRef]
  98. Wang, H.; Wang, Y.; Zhang, J.; Liu, X.; Tao, S. Electronic structure engineering through Fe-doping CoP enables hydrogen evolution coupled with electro-Fenton. Nano Energy 2021, 84, 105943. [Google Scholar] [CrossRef]
  99. Yu, Y.; Chen, Z.; Guo, Z.; Liao, Z.; Yang, L.; Wang, J.; Chen, Z. Removal of refractory contaminants in municipal landfill leachate by hydrogen, oxygen and palladium: A novel approach of hydroxyl radical production. J. Hazard. Mater. 2015, 287, 349–355. [Google Scholar] [CrossRef]
  100. Zhang, Y.; Fan, J.; Yang, B.; Huang, W.; Ma, L. Copper–catalyzed activation of molecular oxygen for oxidative destruction of acetaminophen: The mechanism and superoxide-mediated cycling of copper species. Chemosphere 2017, 166, 89–95. [Google Scholar] [CrossRef]
  101. Yang, Z.; Zhang, X.; Pu, S.; Ni, R.; Lin, Y.; Liu, Y. Novel Fenton-like system (Mg/Fe-O2) for degradation of 4-chlorophenol. Environ. Pollut. 2019, 250, 906–913. [Google Scholar] [CrossRef]
  102. Pi, L.; Cai, J.; Xiong, L.; Cui, J.; Hua, H.; Tang, D.; Mao, X. Generation of H2O2 by on-site activation of molecular dioxygen for environmental remediation applications: A review. Chem. Eng. J. 2020, 389, 123420. [Google Scholar] [CrossRef]
  103. Meijide, J.; Pazos, M.; Sanromán, M.A. Heterogeneous electro-Fenton catalyst for 1-butylpyridinium chloride degradation. Environ. Sci. Pollut. Res. 2019, 26, 3145–3156. [Google Scholar] [CrossRef]
  104. Choe, Y.J.; Kim, J.; Byun, J.Y.; Kim, S.H. An electro-Fenton system with magnetite coated stainless steel mesh as cathode. Catal. Today 2021, 359, 16–22. [Google Scholar] [CrossRef]
  105. Shen, W.; Lin, F.; Jiang, X.; Li, H.; Ai, Z.; Zhang, L. Efficient removal of bromate with core-shell Fe@Fe2O3 nanowires. Chem. Eng. J. 2017, 308, 880–888. [Google Scholar] [CrossRef]
  106. Mu, Y.; Ai, Z.; Zhang, L. Phosphate Shifted Oxygen Reduction Pathway on Fe@Fe2O3 Core–Shell Nanowires for Enhanced Reactive Oxygen Species Generation and Aerobic 4-Chlorophenol Degradation. Environ. Sci. Technol. 2017, 51, 8101–8109. [Google Scholar] [CrossRef] [PubMed]
  107. Ding, Y.; Ruan, Y.; Zhu, L.; Tang, H. Efficient oxidative degradation of chlorophenols by using magnetic surface carboxylated Cu0/Fe3O4 nanocomposites in a wide pH range. J. Environ. Chem. Eng. 2017, 5, 2681–2690. [Google Scholar] [CrossRef]
  108. Cerro-Lopez, M.; Castro-Pastrana, L.I.; Campos-Delgado, J.; Rubio-Rosas, E.; Bustos, E.; Martínez-Huitle, C.A. Mesostructured lead dioxide grown on titania nanotubes for diclofenac water removal through electrocatalytic and photoelectrocatalytic processes. Environ. Res. 2023, 231, 116094. [Google Scholar] [CrossRef]
  109. Umukoro, E.H.; Kumar, N.; Ngila, J.C.; Arotiba, O.A. Expanded graphite supported p-n MoS2-SnO2 heterojunction nanocomposite electrode for enhanced photo-electrocatalytic degradation of a pharmaceutical pollutant. J. Electroanal. Chem. 2018, 827, 193–203. [Google Scholar] [CrossRef]
  110. Hu, X.; Qin, J.Z.; Wang, Y.B.; Wang, J.J.; Yang, A.J.; Tsang, Y.F.; Liu, B.J. Synergic degradation Chloramphenicol in photo-electrocatalytic microbial fuel cell over Ni/MXene photocathode. J. Colloid Interface Sci. 2022, 628, 327–337. [Google Scholar] [CrossRef]
  111. Yap, K.L.; Ho, L.N.; Ong, S.A.; Guo, K.; Oon, Y.S.; Ong, Y.P.; Thor, S.H. Crucial roles of aeration and catalyst on caffeine removal and bioelectricity generation in a double chambered microbial fuel cell integrated electrocatalytic process. J. Environ. Chem. Eng. 2021, 9, 104636. [Google Scholar] [CrossRef]
  112. Guo, W.X.; Chen, M.Q.; Liu, X.Q.; Cheng, F.L.; Lu, X.H. Mo2C/Reduced Graphene Oxide Composites with Enhanced Electrocatalytic Activity and Biocompatibility for Microbial Fuel Cells. Chem.-Eur. J. 2021, 27, 4291–4296. [Google Scholar] [CrossRef] [PubMed]
  113. Kaur, R.; Kushwaha, J.P.; Singh, N. Electro-oxidation of Ofloxacin antibiotic by dimensionally stable Ti/RuO2 anode: Evaluation and mechanistic approach. Chemosphere 2018, 193, 685–694. [Google Scholar] [CrossRef]
  114. Chauhan, R.; Srivastava, V.C. Electrochemical denitrification of highly contaminated actual nitrate wastewater by Ti/RuO2 anode and iron cathode. Chem. Eng. J. 2020, 386, 122065. [Google Scholar] [CrossRef]
  115. Fenti, A.; Jin, Y.; Rhoades, A.J.H.; Dooley, G.P.; Iovino, P.; Salvestrini, S.; Musmarra, D.; Mahendra, S.; Peaslee, G.F.; Blotevogel, J. Performance testing of mesh anodes for in situ electrochemical oxidation of PFAS. Chem. Eng. J. Adv. 2022, 9, 100205. [Google Scholar] [CrossRef]
  116. Goyal, A.; Srivastava, V.C. Treatment of highly acidic wastewater containing high energetic compounds using dimensionally stable anode. Chem. Eng. J. 2017, 325, 289–299. [Google Scholar] [CrossRef]
  117. Sahu, O.P.; Gupta, V.; Chaudhari, P.K.; Srivastava, V.C. Electrochemical treatment of actual sugar industry wastewater using aluminum electrode. Int. J. Environ. Sci. Technol. 2015, 12, 3519–3530. [Google Scholar] [CrossRef]
Figure 1. The scheme of electrocatalytic processes for wastewater treatment: (a) DEO, (b) IEO, (c) ER, (d) cathodic EFP with oxygen reduction, (e) sacrificial anode EFP, and (f) Fe2⁺ recycling EFP with cathodic oxygen reduction.
Figure 1. The scheme of electrocatalytic processes for wastewater treatment: (a) DEO, (b) IEO, (c) ER, (d) cathodic EFP with oxygen reduction, (e) sacrificial anode EFP, and (f) Fe2⁺ recycling EFP with cathodic oxygen reduction.
Processes 12 01615 g001
Figure 2. The schematic mechanism for pollution treatment by IEO: (a) MOX oxidation on the OV-enriched SnOx/La-Sb anode, (b) pollution degradation by Fe/Co-N/P-9 carbon nanofiber, (c) the degradation NPYR in 3DER with ZnAl-LDH/AC system, (d) leachate pretreatment with Fe-DAS dual anode configuration.
Figure 2. The schematic mechanism for pollution treatment by IEO: (a) MOX oxidation on the OV-enriched SnOx/La-Sb anode, (b) pollution degradation by Fe/Co-N/P-9 carbon nanofiber, (c) the degradation NPYR in 3DER with ZnAl-LDH/AC system, (d) leachate pretreatment with Fe-DAS dual anode configuration.
Processes 12 01615 g002
Figure 3. The mechanism of ER for pollutant treatment: (a) the high-efficiency removal of heavy metals from soil leachate, (b) the electrocatalytic nitrate reduction to ammonia.
Figure 3. The mechanism of ER for pollutant treatment: (a) the high-efficiency removal of heavy metals from soil leachate, (b) the electrocatalytic nitrate reduction to ammonia.
Processes 12 01615 g003
Figure 4. (a) The mechanism of degradation of 4-chlorophenol through EFP with the activation of ZVMs. (b) A possible mechanism of bromate removal through EFP activated by core–shell Fe@Fe2O3 nanowires. (c) Schematic illustration for enhanced H2O2 generation with Fe@Fe2O3 Core–Shell nanowires in the presence of phosphate. (d) Possible mechanism of EFP with the activation of O2 on Cu0/Fe3O4 composites.
Figure 4. (a) The mechanism of degradation of 4-chlorophenol through EFP with the activation of ZVMs. (b) A possible mechanism of bromate removal through EFP activated by core–shell Fe@Fe2O3 nanowires. (c) Schematic illustration for enhanced H2O2 generation with Fe@Fe2O3 Core–Shell nanowires in the presence of phosphate. (d) Possible mechanism of EFP with the activation of O2 on Cu0/Fe3O4 composites.
Processes 12 01615 g004
Figure 5. (a) The photo-electrocatalytic mechanism of degradation Chloramphenicol in photo-electrocatalytic microbial fuel cell over Ni/MXene photocathode. (b) The mechanism of caffeine removal and bioelectricity generation in CuO/C cathode-based microbial fuel cell.
Figure 5. (a) The photo-electrocatalytic mechanism of degradation Chloramphenicol in photo-electrocatalytic microbial fuel cell over Ni/MXene photocathode. (b) The mechanism of caffeine removal and bioelectricity generation in CuO/C cathode-based microbial fuel cell.
Processes 12 01615 g005
Table 1. Refractory pollutants and their adverse effects.
Table 1. Refractory pollutants and their adverse effects.
Type of PollutantsSourcesAdverse Effects
AntibioticsMedical wastewaterAllergic reactions, immune system damage, impact on children’s development
Personal care productsMunicipal wastewaterEndocrine disruption, neurological damage, liver and kidney damage, respiratory system effects
DyesPrinting and dyeing textiles, rubber, leather, papermaking industriesCarcinogenicity, mutagenicity, and teratogenicity; skin and mucous membrane irritation; neurological damage
Endocrine disruptorsPesticides, plastics, food additivesEndocrine system disruption, neurological damage, increased cancer risk, developmental effects
Heavy metalsIndustrial wastewaterNeurological damage, reproductive system damage, increased risk of cancer, cardiovascular diseases
Table 2. The advantages and disadvantages of methods for treating refractory wastewater.
Table 2. The advantages and disadvantages of methods for treating refractory wastewater.
Treatment MethodsAdvantagesDisadvantages
Physical methodsHigh treatment efficiency, simple process, and low treatment costsUnable to completely degrade pollutants
Chemical methodsEasy to operate, strong adaptability, broad treatment rangeHigh costs, potential for secondary pollution
Biological methodsEnvironmentally friendly, low cost, capable of degrading a variety of pollutantsLong treatment time, sensitive to the environment, difficult to treat high-concentration pollutants
Electrocatalytic processesLow secondary pollution, mild reaction conditions, high efficiencyHigh energy consumption, electrode materials may need to be replaced regularly
Table 3. The advantages and disadvantages of electrocatalytic technologies.
Table 3. The advantages and disadvantages of electrocatalytic technologies.
Treatment MethodsAdvantagesDisadvantages
DEOEasy to operate, minimal secondary pollutionLow mass transfer efficiency.
IEOHigh treatment efficiency, wide applicabilityHigh energy consumption.
ERHigh treatment efficiency, environmentally friendlyRequires highly selective catalysts to avoid side reactions
EFPHigh treatment efficiencyRequires maintaining an acidic environment
Table 4. Recent progress of DEO.
Table 4. Recent progress of DEO.
Electrode CatalystPollutantsOperating ParametersRemoval Rate
Cu/NF [12]ammonia1.1 V cell potential, 0.01 M ammonia, 100 mM Na2SO4, pH 1161%
Ni2P-s/NF [13]ammoniacurrent density 17.8 mA cm−2, 160 mM ammonia70%
carboxylated multi-walled carbon nanotube-reduced graphene oxide [14]enrofloxacin//
Table 5. Recent progress of IEO for refractory pollutants.
Table 5. Recent progress of IEO for refractory pollutants.
Electrode CatalystPollutantsOperating ParametersActive SpeciesRemoval Rate
Ti3C2Tx [16]metformin0.525 C g−1, 0.005 mA cm−2, pH 6 in absence of NaCl; or 26.25 C g−1, 0.5 mA cm−2 in the presence of 2.5 w/v% NaCl•OH99%
Polypyrrole (PPy) coated carbon (PPy@CCM) [17]phenol2.0 V potential, 2.50 g L−1 Na2SO4, 1.5 mL min−1 flow rate, 50 mg L−1 phenol, 0.5 kWh kgCOD−1 energy consumption•OHPhenol 99.51%, COD 89.90%
GAC@Ni/Fe [18]sulfamethylthiadiazole (SMT)5 V cell voltage, 3 g particle electrode dosage, 2 cm electrode plate spacing, 1 mg L−1 SMT initial concentration•OH90.89% in 30 min
NiO–TiO2 [19]ammonia200 mM NH4OH, 100 mM NaNO3, pH 9 with a current density of 30 mA cm−2•OH96.4% in 9 h
spinel CuxCo1−xMn2O4 [20]tetracyclinecurrent density 20 mA cm−2, pH = 3 and 20 mg L−1 tetracycline•OH91.3% in 120 min
GAC-Co-Mn [21]sulfanilamidevoltage 15 V, weak acid and weak base, aeration velocity 16 L h−1, sulfanilamide 50 mg L−1•OHSulfanilamide 92.881% and TOC 82.335% in 90 min
Ti–Ag/γ-Al2O3 [22]pesticide wastewatercurrent density of 30 mA cm−2, initial pH of 2.0, electrode distance of 3.0 cm, air flow of 3.0 L min−1/COD 82.50%
SnO2-Sb NPA [23]Bisphenol Acurrent density 20 mA cm−2, neutral pH, 25 mg L−1 BPA and 0.1 M Na2SO4•OH and •O2Bisphenol A 100%, TOC 42.8% in 120 min
Ti/SnO2–RuO2 [24]cefotaxime sodiumcurrent density 25 mA cm−2, NaCl 5 mM, pH 7, cefotaxime 100 mg L−1•OH and ClO86.33% in 45 min
SrxLa1-xMnyCo1-yO3-δ [25]ammonium nitrogen and methyl orangeT = 35 °C, current density = 20 mA cm−2, pH = 9, CCl = 2000 mg L−1•OH, •O2 and active chlorineammonium nitrogen 98.23%, methyl orange 99.61% and COD 88.16% in 60 min
SnOx/La-Sb [26]moxifloxacin10 mg L−1 moxifloxacin, current density 2.5 mA cm−2 in 0.05 M Na2SO4 at neutral pH•OHnearly 100% and TOC 41% in 30 min
CuFe2O4/ACF cathode and RuO2/Ti anodes [27]phenol1.00 mM PS, current density of 50 A m−2, and initial pH of 7.0•OH and SO4•−97% in 60 min
boron-doped diamond anode [28]2,4-Dichlorophenoxyacetic acid2 M sulfuric acid, current efficiency 5 mA cm−2SO4•−96%
Fe [29]landfill leachatepH 6.8, applied current of 200 mA, oxidant dosage of 6 mM•OH and SO4•−TOC 73% in 80 min
DSA [30]landfill leachatecurrent density of 12.5 mA cm−2 and NaOH dosage of 10 g/LRCS and •OHCOD 50.94% NH4+-N 38.11%
platinum-coated titanium [31]Humic acids10 mM NaCl, current efficiency 153 A m−2RCS and •OH/
iron anode reactor and Ti/RuO2 anode reactor (TAR) [32]landfill leachateIAR: current density 70 mA/cm2 and initial pH value 6.5
TAR:current density 70 mA/cm2 and initial pH value 9.0
RCS and •OHTN 98.3%, COD 94.6%
Table 6. Commonly used RCS quenching agents, their types, and their corresponding rate constants.
Table 6. Commonly used RCS quenching agents, their types, and their corresponding rate constants.
Type of QuencherQuenching of Free RadicalsReaction Rate Constant
Benzoicacid (BA)•OH, •ClK*OH = 5.9 × 109 M−1 s−1, KCl* = 1.8 × 1010 M−1 s−1
1-4Dioxomethylbenzene (DMOB)•OH, •Cl, •ClOK*OH = 7.0 × 109 M−1 s−1, KCl* = 1.8 × 1010 M−1 s−1, KClO* = 2.1 × 109 M−1 s−1
Cassimapine (CBZ)•OH, •Cl, •ClO, •Cl2K*OH = 8.8 × 109 M−1 s−1, KCl* = 3.3 × 1010 M−1 s−1, KClO* = 1.97 × 108 M−1 s−1, KClO2*- = 4.3 × 107 M−1 s−1
Table 7. The recent research on electrocatalytic reduction for degradation of pollutants.
Table 7. The recent research on electrocatalytic reduction for degradation of pollutants.
Electrode CatalystPollutantsOperating ParametersRemoval Rate
conductive carbon fibers cloth cathode and Platinum-coated titanium panel anode [57]Nickel Sulfate and Copper Sulfate (0.06 M)low voltage supply energy of 10 V, and pH value of 6.897% in 20 h
IrO2–Ta2O5/Ti anode and graphite cathode [58]mixed heavy metal ions extracted from soils by citratecurrent density 26 mA cm−299.4% in 12 h
multi-hole stainless steel anode and multi-hole titanium cathode [59]Cr and Cu ionscurrent density 0.40 mA cm−2COD 87.23%, TOC 80.42%, Cr 91.25% Cu 95.97%
PdNi/PPy-rGO/Ni [60]diclofenac20 mg L−1 diclofenac, current 7 mA100% in 140 min
multi-walled carbon nanotubes MWCNT-Ti4O7 [61]dibromoacetic acid (DBAA)1 mg L−1 DBAA in 10 mM KH2PO4/K2HPO4 at −1.5 V SHE96%
Pd7Au3 ANs [62]4-chlorophenol50 mM potassium sulfate solution mixed with 50 mg L−1 4-chlorophenol, −1.1 V (vs. SCE)98.35% in 4 h
Cu3P/CF [63]nitrates−1.2 V (vs. Ag/AgCl), 50 mM Na2SO497.7%
Co/CoO NSAs [64]nitrates−1.6 V (vs. SCE), 0.1M Na2SO491.2%
Pd-TiO2 [65]nitrates−0.55 V (vs. RHE), 0.25 M NO399.6%
Table 8. Total operating cost for different types of pollutants.
Table 8. Total operating cost for different types of pollutants.
Electrode CatalystElectrolytePollutantsElectrical Energy ConsumedCost of Electrical EnergyCost of ElectrodesTotal Operating Cost
BDD,TinO2n−1,Ti/IrO2-Ta2O5poly (methyl methacrylate)perfluorooctane sulfonate [115]4.0 kWh m−30.45 EUR m−3/0.45 EUR m−3
Ti/RuO2NaClpharmaceuticals wastewater0.542 kWh m−30.033 EUR m−31.99 EUR m−32.02 EUR m−3
Ti/RuO2NH4Clnitrate [114]108.04 kWh m−37.89 USD m−3286.25 USD m−3294 USD m−3
Ti/RuO2/highly acidic wastewater [116]735.72 kWh m−336.7 USD m−3400 USD m−3436 USD m−3
Al/sugar industry wastewater [117]58 kWh m−33.57 USD m−32.65 USD m−36.22 USD m−3
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, X.; Yang, J.; Guo, J.; Xiong, W.; Leung, M.K.H. Advances and Prospects in Electrocatalytic Processes for Wastewater Treatment. Processes 2024, 12, 1615. https://doi.org/10.3390/pr12081615

AMA Style

Zhou X, Yang J, Guo J, Xiong W, Leung MKH. Advances and Prospects in Electrocatalytic Processes for Wastewater Treatment. Processes. 2024; 12(8):1615. https://doi.org/10.3390/pr12081615

Chicago/Turabian Style

Zhou, Xince, Jiajie Yang, Jiahuan Guo, Wei Xiong, and Michael K. H. Leung. 2024. "Advances and Prospects in Electrocatalytic Processes for Wastewater Treatment" Processes 12, no. 8: 1615. https://doi.org/10.3390/pr12081615

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop