Performance enhancement strategies for electrooxidation degradation of landfill leachate: A review

Zhenhui Song Xing Wu Tianyu Gao Fubing Yao Xi Tang Qaisar Mahmood Chong-Jian Tang

Citation:  Zhenhui Song, Xing Wu, Tianyu Gao, Fubing Yao, Xi Tang, Qaisar Mahmood, Chong-Jian Tang. Performance enhancement strategies for electrooxidation degradation of landfill leachate: A review[J]. Chinese Chemical Letters, 2025, 36(12): 111008. doi: 10.1016/j.cclet.2025.111008 shu

Performance enhancement strategies for electrooxidation degradation of landfill leachate: A review

English

  • Sanitary landfills are a typical technique for disposing of municipal solid waste [1]. Waste decomposition and residual water in landfills generate significant leachate [2]. Landfill leachate contains high levels of organics and other pollutants, posing a threat to the environment and human health (Table S1 in Supporting information). Biotechnology is a promising and cost-effective wastewater treatment technique [3]. However, such method is limited by the high concentration of bio-refractory pollutants (e.g., humic acid (HA), perfluorooctanesulfonate), heavy metals, ammonium (NH4+), and salt [47]. Advanced oxidation processes (AOPs) such as Fenton, photocatalysis, and ozone catalysis can disrupt the molecular structure of refractory organics in landfill leachate [812]. Although AOPs can generate hydroxyl (OH) or sulfate radicals (SO4•-) to convert organic ammonium into NH4+, NH4+ removal remains a significant challenging [13]. Besides, Fenton or heterogeneous catalytic oxidation processes generate excess sludge, increasing treatment costs, necessitating the development of effective alternative methods.

    Based on landfill age, leachate can be classified into young, intermediate, and aged stages. The chemical oxygen demand (COD) concentration of young and intermediate leachate is typically above 4000 mg/L and can even exceed 10,000 mg/L, and the organics contain higher levels of low molecular weight organic acids and aliphatic organic compounds [14]. The ratio of 5-day biochemical oxygen demand to chemical oxygen demand (BOD5/COD) is commonly above 0.4, indicating that biological method is more suitable [15]. As landfill age increases, low molecular weight organic acids are removed, leaving behind large, refractory compounds such as HA and antibiotics. Although COD drops below 4000 mg/L, the BOD5/COD ratio remains below 0.15 [15], inhibiting the biological activity. However, recalcitrant organic compounds can be effectively degraded by electrochemical oxidation (EO), making this method a promising alternative for treating leachate from aged landfills.

    EO of landfill leachate has gained increasing attention for its effectiveness in degrading recalcitrant organics, removing nitrogen, and producing no secondary pollution. Organics can directly lose electrons on anode surface and simultaneously decomposes into smaller molecule substances [16]. NH4+ can be directly oxidized on noble metal (such as Pt, Pd) anode surface [17]. H2O, Cl-, and SO42- in leachate can be converted into reactive species such as OH, Cl, and SO4•-, which degrade complex humic and fulvic acids into phenolic residues, aniline derivatives, and polycyclic aromatic hydrocarbons, eventually mineralizing them into CO2 [18]. NH4+ can be oxidized to N2 by Cl, which is the main pathway for nitrogen removal in EO [19].

    The direct electron transfer process and indirect oxidation by oxidants primarily depend on the anode for the electrochemical reaction. Recently, various anodes such as boron-doped diamond (BDD), lead dioxide (PbO2), ruthenium dioxide (RuO2), tin dioxide (SnO2) and their composite electrode were studied for EO landfill leachate treatment. For example, 83.7% of COD was removed under 100 mA/cm2 with ruthenium iridium oxide electrode (Ti/RuO2-IrO2) [20]. The use of graphite/PbO2 electrode could remove 35.1% COD and 97.6% NH4+ from leachate after 2 h of electrolysis [21]. However, studies on electrochemical landfill leachate treatment are still largely limited to the laboratory stage due to technical limitations of the electrodes, including poor electron transfer performance, small specific surface area, single reactor operation mode, and insufficient space utilization.

    To overcome these limitations and enhance the electrooxidation efficiency of landfill leachate, extensive research efforts have been undertaken. However, current reports on EO mechanisms are mostly generic, lacking theoretical studies tailored to the characteristics of landfill leachate. Meanwhile, recent review papers on EO have mostly focused on comparing technologies like electrocoagulation, electro-adsorption, electro-Fenton, and photoelectrocatalysis [3,22,23], with scant attention given to enhancing the anodic electrooxidation efficiency. Therefore, this paper specifically summarized the EO mechanisms applicable to treating leachate from aged landfills, proposed feasible strategies to increase the electrode activity, and discussed the influence mechanism of operating parameters. Its goal is to provide guidance for the practical EO engineering of landfill leachate.

    The removal of pollutants in landfill leachate using EO technology can be categorized into direct and indirect pathways (Fig. 1). The direct oxidation refers to the capture of electrons from contaminants by active sites on anode under applied voltage [24]. Such reaction is a straightforward oxidation process. Indirect oxidation is the breakdown or mineralization of pollutants by oxidizing active chemicals (OACs) (OH, SO4•-, active chlorine, etc.) generated from anode [25].

    Figure 1

    Figure 1.  Schematic illustration for the EO process. R denotes an organic pollutant and M is the active site on the electrode surface.

    The direct oxidation process contains two steps. (a) Pollutants diffuse from the electrolyte to active sites on anode surface. (b) The electrode takes electrons from pollutants (Eq. 1) since the active sites have a high capacity for acquiring electrons [26]. The direct oxidation commonly has easy electron transport and quick reactions.

    $ \mathrm{M}+\mathrm{R} \rightarrow \mathrm{M}(^•\mathrm{R})+\mathrm{ne}^{-} \rightarrow \text { Product } $

    (1)

    M is the active site. R is the target pollutant.

    But, for aged leachate, the organic compounds often contain stable molecular structures such as benzene rings, naphthalene rings, heterocycles, which are difficult to degrade through direct oxidation pathways [16]. For example, the free radical quenching experiment confirms that direct oxidation only achieved a removal efficiency of 39.6% for HA [27]. More importantly, pollutants in leachate that cannot be quickly decomposed may accumulate on the electrode surface, leading to passivation. Therefore, degradation of leachate primarily relies on indirect oxidation pathways.

    2.2.1   The mechanism of OACs generation and transformation

    The indirect oxidation is mediated by various OACs including free radicals and reactive substances generated from anode. Generally, water can be decomposed into OH (2.8 V vs. SHE) (Eq. 2) in electrochemical system, which has the strongest oxidizing characteristics [28]. The OH also has high electronegativity or electrophilicity (the electron affinity is 569.3 kJ/mol), tending to attack sites with high electronic density in organics [29]. The OH is produced and adsorbed on the anode surface, where it is employed for organic oxidation (Eq. 3) via electron transfer, dehydrogenation, and addition [30]. The organics can be converted into small molecule organics or CO2 through such reaction. However, OH can also cause oxygen evolution reaction (OER) with H2O (acid condition) or OH- (alkaline condition) and then release electrons (Eqs. 4 and 5) [31]. This side reaction will impair the efficiency of pollutant oxidation and increase energy consumption.

    $ \mathrm{M}+\mathrm{H}_2 \mathrm{O} \rightarrow \mathrm{M}\left({ }{^•\mathrm{OH}}\right)+\mathrm{H}^{+}+\mathrm{e}^{-} $

    (2)

    $ \mathrm{M}({^•\mathrm{OH}})+\mathrm{R} \rightarrow \mathrm{M}+\mathrm{RO}+\mathrm{H}^{+}+\mathrm{e}^{-} $

    (3)

    $ \mathrm{M}({^•\mathrm{OH}})+\mathrm{H}_2 \mathrm{O} \rightarrow \mathrm{M}+\mathrm{O}_2+3 \mathrm{H}^{+}+3 \mathrm{e}^{-} $

    (4)

    $ \mathrm{M}\left({^•\mathrm{OH}}\right)+\mathrm{OH}^{-} \rightarrow \mathrm{M}+\mathrm{O}_2+2 \mathrm{H}^{+}+3 \mathrm{e}^{-} $

    (5)

    In addition, Cl-, a common anion in landfill leachate, can be oxidized into Cl through dimensional stability anode (DSA), such as RuO2-Ir, IrO2-TiO2, and Co3O4 [32]. These DSA can enhance electron transfer and catalytic activity through the cyclic changes in the valence states of their constituent metal elements, enabling direct oxidation of Cl- to Cl at the active sites by electron extraction near the anode. For example, the Co2+/Co3+ redox couple in the Co3O4 anode can act as a catalyst pair to rapidly oxidize Cl- to Cl, enabling efficient removal of NH4+ [33]. The Cl- can also be indirectly oxidized to Cl by OH (Eq. 6) [3]. The Cl will further react to form various active chlorine species (Eqs. 7-9), such as Cl2 (E0 = 1.36 V vs. SHE), HClO (E0 = 1.63 V vs. SHE), and ClO- (E0 = 0.90 V vs. SHE) [34,35]. Notably, active chlorines exist in different forms under different pH, such as Cl2 (pH ≤ 3), HClO (pH 5–7), ClO- (pH ≥ 8) [36]. However, electrooxidation mediated by active chlorine might produce some toxic chlorination byproducts like chloramine, trihalomethanes, haloacetic acids [30,37].

    $ {^•\mathrm{OH}}+\mathrm{Cl}^{-} \rightarrow \mathrm{OH}^{-}+\mathrm{Cl}^• $

    (6)

    $ 2 \mathrm{Cl}^{•} \rightarrow \mathrm{Cl}_2 $

    (7)

    $ \mathrm{Cl}_2+\mathrm{H}_2 \mathrm{O} \rightarrow \mathrm{HOCl}+\mathrm{H}^{+}+\mathrm{Cl}^{-} $

    (8)

    $ \mathrm{HOCl} \rightarrow \mathrm{ClO}^{-}+\mathrm{H}^{+} $

    (9)

    The generation and action mechanisms of SO4•- are shown in Fig. 2. The SO42- and HSO4- in leachate can be directly oxidized to SO4•- on anode surface through single-electron transfer [38]. The potential for the direct conversion of SO42- or HSO4- to SO4•- is typically around 2 V vs. Ag/AgCl [39]. Thus, such direct oxidation pathway is very weakly [40]. The SO4•- is mainly obtained from the indirect oxidation process mediate by OH during electrochemical leachate treatment [41]. Meanwhile, the yield and stability of SO4•- are significantly dependent upon the pH. Low pH (pH ≤ 3), OH can maintain excellent oxidation ability, promoting the generation of SO4•- (Eq. 10) [42]. The SO4•- has good stability and its hydrolysis can be ignored under such condition [43]. In contrary, SO4•- easy hydrolyzes into OH at pH 9.0 (Eqs. 11 and 12). Remarkably, SO4•- will be completely react with OH-/H2O to OH (Eqs. 11 and 12) as pH ≥ 12. The types of active species present in a system can be identified through radical probes, electron paramagnetic resonance (EPR), or quenching experiments (Text S1, Tables S2 and S3 in Supporting information).

    $ \mathrm{SO}_4^{2-}+{^•\mathrm{OH}} \rightarrow \mathrm{SO}_4 ^{•-}+\mathrm{OH}^{-} $

    (10)

    $ \mathrm{SO}_4{ }^{•-}+\mathrm{OH}^{-} \rightarrow \mathrm{SO}_4{ }^{2-}+{^•\mathrm{OH}} $

    (11)

    $ \mathrm{SO}_4{ }^{•-}+\mathrm{H}_2 \mathrm{O} \rightarrow \mathrm{HSO}_4^{-}+ {^•\mathrm{OH}} $

    (12)

    Figure 2

    Figure 2.  Sulfate radical-mediated oxidation and anode-cathode synergy.

    The OACs can also produce from the synergistic performance of oxidation and reduction during electrochemical reaction. For example, the oxygen reduction reaction (ORR) can selectively reduce O2, either generated at the anode or dissolved in water, to H2O2 (Eqs. 13 and 14). Specifically, O2 is first adsorbed onto active sites and then, with the involvement of protons (H+), is ultimately reduced to H2O2 via a two-electron pathway on the cathode surface [44]. Since the reaction requires the protons, it occurs to a greater extent in young and intermediate-aged landfill leachate compared to aged leachate. H2O2 can also be formed through the dimerization of two OH radicals (Eq. 15). Similarly, SO42- is oxidized at the anode surface through a direct oxidation process or an indirect oxidation mediated by OH, generating SO4•-, which can further dimerize into persulfate (S2O82-) (Eq. 16) [45]. The H2O2 and S2O82- can be reactivated at the cathode via a direct electron transfer pathway, producing OH and SO4•- (Eqs. 17 and 18), respectively [45].

    $ 2 \mathrm{H}_2 \mathrm{O} \rightarrow \mathrm{O}_2+4 \mathrm{H}^{+}+4 \mathrm{e}^{-} $

    (13)

    $ \mathrm{O}_2+2 \mathrm{H}^{+}+2 \mathrm{e}^{-} \rightarrow \mathrm{H}_2 \mathrm{O}_2 $

    (14)

    $ 2 {^•\mathrm{OH}} \rightarrow \mathrm{H}_2 \mathrm{O}_2 $

    (15)

    $ 2 \mathrm{SO}_4{ }^{•-} \rightarrow \mathrm{S}_2 \mathrm{O}_8^{2-} $

    (16)

    $ \mathrm{H}_2 \mathrm{O}_2+\mathrm{e}^{-} \rightarrow{ }^{•} \mathrm{OH}+\mathrm{OH}^{-} $

    (17)

    $ \mathrm{S}_2 \mathrm{O}_8{ }^{2-}+\mathrm{e}^{-} \rightarrow \mathrm{SO}_4{ }^{•-}+\mathrm{SO}_4{ }^{2-} $

    (18)

    H2O2 or S2O82- can also be activated by indirect pathways (Fig. 2). The ORR reaction at the cathode can also produce reactive oxygen substances (O2•- and HO2) (Eq. 19). The O2•-, as a highly unstable substance, can combine with H+ to form HO2 (Eqs. 20 and 21). Although O2•- and HO2 with low redox potential have poor oxidation property for pollutants removal, they can effectively oxidize H2O2 and S2O82- to OH and SO4•- by seizing electrons (Eqs. 22-25), respectively [45,46].

    $ \mathrm{O}_2+\mathrm{e}^{-} \rightarrow \mathrm{O}_2^{•-} $

    (19)

    $ \mathrm{O}_2^{•-}+\mathrm{H}^{+} \rightarrow \mathrm{HO}_2^{•} $

    (20)

    $ \mathrm{OH}^{-}+\mathrm{HO}_2^{•} \rightarrow \mathrm{O}_2^{•-}+\mathrm{H}_2 \mathrm{O} $

    (21)

    $ \mathrm{H}_2 \mathrm{O}_2+\mathrm{O}_2^{•-} \rightarrow \mathrm{O}_2+\mathrm{OH}^{-}+{^•\mathrm{OH}} $

    (22)

    $ \mathrm{H}_2 \mathrm{O}_2+\mathrm{HO}_2^{•} \rightarrow \mathrm{O}_2+\mathrm{H}_2 \mathrm{O}+{^•\mathrm{OH}} $

    (23)

    $ \mathrm{S}_2 \mathrm{O}_8^{2-}+\mathrm{HO}_2^{•} \rightarrow \mathrm{O}_2+\mathrm{SO}_4^{2-}+\mathrm{SO}_4^{•-}+\mathrm{H}^{+} $

    (24)

    $ \mathrm{O}_2{ }^{•-}+\mathrm{S}_2 \mathrm{O}_8{ }^{2-} \rightarrow \mathrm{SO}_4{ }^{•-}+\mathrm{SO}_4{ }^{2-}+\mathrm{O}_2 $

    (25)
    2.2.2   Oxidation mechanism of leachate

    Taking the above discussion into account, OH, SO4•- and Cl derivatives may be the primary active agents for leachate treatment. Due to the unique property leachate, its degradation pathway mainly involves the removal of organics and NH4+. The SO42- is a common ion in landfill leachate, which can be oxidized to SO4•- through direct or an OH-mediated indirect oxidation process [41]. The SO4•- can degrade aromatic contaminants in landfill leachate. The UV–vis and three-dimensional fluorescence spectroscopy analysis showed a decrease of about 65% in aromaticity and hydrophobicity, and a 49.03% reduction in humification after SO4•- oxidation of landfill leachate in the Fenton process [47]. However, in EO process, SO4•- can be quenched by OH. For example, SO4•- is not observed in the EPR spectra during the electrooxidation of HA [48,49]. Thus, organics and NH4+in leachate are mainly removed by the OH and Cl derivatives.

    Compared with OH, Cl derivatives show significantly faster degradation rates of pollutants, especially for organics. For example, Cl derivatives exhibited 1.9 times higher efficiency than that of OH in decomposing non-purgeable organic carbon and faster color removal from dark brown landfill leachate [48]. This is because OH accumulates at active sites on anode surface through physical adsorption [30]. To undergo OH mediated oxidation reaction, pollutants must transfer from solution to anode surface. Inversely, Cl derivatives generated on anode are soluble in solution and diffuse throughout the entire reactor, expanding the effective reaction area and enhancing removal efficiency. On the other hand, Cl derivatives can rapidly break down large molecules in leachate into smaller compounds such as phenol, aniline, benzoic acid, and polycyclic aromatic carbon hydride [50,51]. However, their lower oxidation potentials (Cl2 1.36 V vs. SHE, HClO 1.63 V vs. SHE, and ClO- 0.90 V vs. SHE) not only make it difficult to further degrade the stable aromatic ring structures, but also result in the formation of chlorinated aromatic compounds [35]. The OH with higher oxidation potential facilitates dechlorination and hydroxylation of these chlorinated products [52,53]. Research suggests that hydroxylation is a pivotal step in the aromatic ring-opening process, thus OH achieves much higher mineralization efficiency for organic compounds in leachate [28,54]. However, the redox potential and stability of OH is essentially dependent on the solution pH. For example, OH can be quenched by H2O or OH- under neutral or alkaline conditions (Eqs. 4 and 5). Strong alkaline conditions can also decrease the redox potential of OH, weakening the oxidative ability [55]. It is worthy nothing that HA is the most important contaminant in aged leachate, accounting for >60% of total organics. And its existing form is significantly affected by the solution pH. The detailed influence mechanism of pH will be further discussed in the section of 4.3.

    NH4+, especially in aged leachate, can exceed 1500 mg/L [15]. Active chlorine can selectively oxidize NH4+ to N2, which is the primary pathway for NH4+ removal in landfill leachate (Eqs. 26-31) [56]. For example, active chlorine achieved a removal efficiency of 97.85% for NH4+ in landfill leachate after 8 h of EO, while OH only removed 81.18% [57]. Taking the aged leachate (weakly alkaline) into account, active chlorine predominantly exists in the form of ClO- (E0 = 0.90 V vs. SHE). ClO- oxidation of NH4+ can minimize the formation of by-products such as nitrite (NO2-) and nitrate (NO3-) [58].

    $ \mathrm{HOCl}+\mathrm{NH}_4^{+} \rightarrow \mathrm{NH}_2 \mathrm{Cl}+\mathrm{H}_2 \mathrm{O}+\mathrm{H}^{+} $

    (26)

    $ \mathrm{HOCl}+\mathrm{NH}_2 \mathrm{Cl} \rightarrow \mathrm{NHCl}_2+\mathrm{H}_2 \mathrm{O} $

    (27)

    $ \mathrm{H}_2 \mathrm{O}+\mathrm{NHCl}_2 \rightarrow \mathrm{NOH}+2 \mathrm{H}^{+}+2 \mathrm{Cl}^{-} $

    (28)

    $ \mathrm{NOH}+\mathrm{NHCl}_2 \rightarrow \mathrm{~N}_2+\mathrm{HOCl}+\mathrm{H}^{+}+\mathrm{Cl}^{-} $

    (29)

    $ 2 \mathrm{NH}_4^{+}+3 \mathrm{HOCl} \rightarrow \mathrm{~N}_2+3 \mathrm{H}_2 \mathrm{O}+5 \mathrm{H}^{+}+3 \mathrm{Cl}^{-} $

    (30)

    $ \mathrm{NH}_4^{+}+4 \mathrm{HOCl} \rightarrow \mathrm{NO}_2^{-}+\mathrm{H}_2 \mathrm{O}+6 \mathrm{H}^{+}+4 \mathrm{Cl}^{-} $

    (31)

    Overall, Cl derivatives significantly contribute to the breakdown of macromolecules (e.g., HA), NH4+ removal, and decolorization, while OH plays a key role in the complete mineralization of organic pollutants in aged leachate.

    Based on the above discussion, the primary pathway for removing organics and NH4+ from landfill leachate is indirect oxidation mediated by OACs. The generation of OACs is closely linked to the anode. How to design the anode with excellent performance for steady and efficient production of active chemicals is the main focus of the recent of electrochemical landfill leachate treatment. Generally, the performance of anode is related on its physical and electrochemical characteristics including mechanical stability, conductivity, active sites, and mass transfer efficiency. To promote and regulate anode property, several approaches have been proposed (Table S4 in Supporting information), including adding intermediate layer materials, synthesizing porous membrane structures, creating surface microstructures, and filling particle microelectrodes.

    The common anode was constructed with catalyst layer and a conductive base. However, catalyst layer (β-PbO2, RuO2, and IrO2, etc.) has poor adherence to substrate, easy slipping off from electrode during electrochemical reaction, particularly under high current density. Moreover, the catalyst layer coverage on the substrate is often incomplete, leading to passivation of the base. Introducing an active intermediate layer can effectively address these challenges (Fig. 3).

    Figure 3

    Figure 3.  Diagram of the anode with intermediate layer.

    The intermediate layer can enhance material adhesion and reduce substrate exposure [59]. Metal oxide or metal element not only have good adhesion, but also own high conductivity. For instance, Pt, with its excellent adhesion and conductivity, can serve as the intermediate layer to enhance electrode stability and reduce electrochemical impedance [60]. The electrode with SnO2-Sb intermediate layer (Ti/SnO2-Sb/β-PbO2) features a long service life (214 h) and low electrochemical impedance (1.315 Ω/cm2) [61]. Besides, the intermediate layer can regulate the crystal of catalysts structure, improving the performance of electrode. The introduction of siloxides (SiOx), a hydrophilic material, can promote the nucleation of β-PbO2 crystals and further reduce electrochemical impedance [62].

    The intermediate layer material with special morphological characteristics such as two-dimension, porous and nanowire can improve the anode’s performance. The special morphological characteristic provides more sites for catalyst loading and bonding [63]. For instance, the introduction of MXene (Ti3C2Tx) with excellent conductivity, layered structure and hydrophilicity can significantly increase the specific surface area of the anode. The number of active sites and OH output of Ti/Ti3C2Tx/PbO2 were 5.21 and 4.07 times higher than Ti/PbO2, respectively [64]. Besides, carbon nanotubes (CNTs) or other nanomaterial can be regarded as intermediate layer to increasing the specific surface area of anode and improving improve the adhesion between each layer, enhancing the electrochemical performance for landfill leachate treatment [65]. However, although carbon-based materials offer excellent electrochemical performance and morphological versatility, the high potential and acidic environment can lead to carbon oxidation and performance degradation over time [66].

    The EO of leachate is achieved via indirect oxidation with free radical. However, free radical exists only in the boundary layer of micrometer or even nanometer level on electrode surface. The traditional operating mode limits the mass transfer efficiency between pollutants and the anode surface [67]. Flow-through membrane electrodes (MEs) with circular flakes and hollow cylindrical shapes have been developed (Fig. 4). These structures not only increase the electrochemically active sites but also enhance the diffusion of contaminants onto the electrode surface [48,68,69]. Researchers found that flow-through membrane electrode can reduce the thickness of diffusion layer from 100 µm to 1 µm, breaking the shackles of mass transfer by-pass operation [70]. Contaminants can be degraded and even mineralized within 5 s with electrochemical membrane electrode in flow-through mode [71,72]. Titanium dioxide (Ti4O7), SnO2, PbO2, and RuO2, after pressing and calcination, can form good porosity and are commonly used materials for fabricating porous membrane anodes [7376].

    Figure 4

    Figure 4.  Diagram of the porous membrane anode in flow-through mode.

    Doping metal elements (bismuth (Bi), antimony (Sb), niobium (Nb), iridium (Ir), etc.) into porous membrane materials can increase free electron density, reducing electrochemical impedance of electrode. For example, compared with SnO2 porous membrane electrode (182 Ω), Sb-doped SnO2 owns lower electrochemical impedance (62 Ω) [77,78]. Similarly, PbO2-Bi and RuO2-Ir materials have low electrochemical impedance compared to PbO2 and RuO2, respectively [79,80]. The doping metal element can also create oxygen vacancies and increase active sites through the substitution reaction, enhancing the electrocatalytic activity of electrode. For example, doping cerium (Ce) (1–3 at%) into Ti4O7 can significantly facilitate interfacial charge transfer and increase OH production (37%−129%) due to the creating of surficial oxygen vacancies [68]. Anuchai et al. increased the oxygen vacancy concentration in SnO2 by 19.5%, generating more O2- radicals and thereby enhancing the oxidation efficiency of methyl orange from 58% to 98% [81]. Combining carbon-based materials with porous membrane electrodes can also effectively enhance anode activity by forming chemical bond interfaces [82]. Compared with pristine Ti4O7, the introduction of graphene oxide nanoparticles (GONs) significantly decreases charge-transfer resistance (from 73.87 Ω to 8.42 Ω) and generated OH at 2.5–2.8 times higher rate due to the formation of GONs-O-Ti chemical bonds [83]. Li et al. also found that the generation of Fe-O-GO—O-Ti bonds between graphene oxide (GO), iron oxide (Fe2O3), and Ti4O7 can boost the electroactivation of peroxymonosulfate for 1,4-dioxane removal due to the increase of electron transfer rate and reactive sites [84].

    It is worth noting that landfill leachate contains a large amount of suspended organic matter, such as HA, which can lead to membrane fouling and significant deactivation of active sites when treated directly with membrane electrodes. Therefore, even though porous membrane electrodes exhibit high electrooxidation efficiency, they are more suitable for treating low-concentration leachate, such as effluent from biological or coagulation pretreatment processes.

    In practical engineering for EO of leachate, plate electrodes with limited specific surface area and active sites are typically used. Recently, three-dimensional (3D) microstructured anodes (Fig. 5), such as nanotubes, nanowires, and nanorods, have gained attention. These anodes feature large specific surface areas and multiple active sites [85,86]. For instance, NH4+ removal efficiency with cobalt oxide (Co3O4) nanowire was 2.58 times than that of ordinary Co3O4 plate [33].

    Figure 5

    Figure 5.  Diagram of anode with three-dimensional microstructure.

    The catalyst with 3D microstructure exhibits excellent activity for pollutant removal. However, a key challenge is the simple and cost-effective integration of nanostructured catalytic layers with substrates. The common strategy is using polymer as binder to load electrocatalyst onto substrate. But the polymer inevitably increases impedance and affects stability. To solve these problems, researches have been committed to in-situ grow 3D catalyst onto the electrode surface (binder-free electrode). Blue-titanium nanoarrays (BTNAs) can be grown in-situ on Ti substrates through two-step methods involving anodic oxidation and cathodic reduction. Their multiple active sites and good electron transfer ability can significantly increase generation of free radicals [87]. Pierpaoli et al. prepared carbon nanostructured electrode through in-situ grown boron doped vertically aligned graphene walls (BCNWs) onto BDD interfacial layer. It was found that in landfill leachate electrooxidation, COD removal efficiency was 1.66 times than that of BDD with the same geometric area [88].

    Nanoscale modification of 3D catalytic layers is also an effective method to improve anode electrochemical activity. The loading of metal element (such as Pt, Rh, Pd) in the form of single atom or nanocluster can boost the oxidation capacity of the surrounding active sites. For example, the modification with carbon nanotubes supported Pt single atom can minimize oxygen evolution side reactions and boost active chlorine creation [89]. Additionally, introducing more oxygen-containing functional groups (e.g., hydroxyl, carboxyl, and carbonyl) into nanostructured carbon-based materials can alter the electron cloud density of surrounding atoms, expand the potential window, and enhance electrode performance [90].

    The reactive sites and electron transfer rate are the essential parameters for EO of landfill leachate. Generally, they remain constant for a given electrode, limiting the efficiency of leachate degradation. Researchers discovered that adding catalysts, referred to as particle microelectrodes, can improve EO efficiency (Fig. 6). This method not only expands the reaction space but also enhances electron transfer [91]. Carbon-based materials are commonly used as particle microelectrodes to enhance electron transfer and increase reactive sites [92]. Using carbon aerogel as particle microelectrodes to construct a three-dimensional reactor demonstrated enhanced and sustained activity for phenol removal, achieving an efficiency of up to 98% [93].

    Figure 6

    Figure 6.  Diagram of the packed bed reactor with particle microelectrodes.

    Under electric field, particle microelectrodes will polarize to form numerous micro-electrolytic cells, serving as active centers for radical generation [94]. In an electro-Fenton system with Fe particles, OH generation was enhanced, leading to improved COD removal [95]. Chen et al. studied EO of dinitrotoluene (DNT) wastewater assisted by Sn-Sb-Ag-modified ceramic particulates (SCP). The OH content in a three-dimensional electrochemical system filled with SCP particles was 1.85 times higher than that without SCP [96]. Additionally, the use of filler particles provides multiple convergent-divergent channels, preventing blockages and ensuring stable EO of leachate [97]. Therefore, inserting particle microelectrodes into an electrochemical reactor is an effective method to improve landfill leachate treatment efficiency by increasing the electrocatalytic active area, thereby enhancing the removal of COD, NH4+, and chromaticity [98].

    Precisely, enhancing pollutant mass transfer, increasing the specific surface area, and exposing more active sites are critical to improving anodic oxidation performance. Additionally, improving electron transfer efficiency within the anode material to enhance its conductivity can reduce energy consumption and prevent anodic electrocorrosion caused by excessive applied voltage. Therefore, innovative research on anode design and reactor construction should be conducted based on the above key points.

    Besides electrodes, many EO equipment are designed to achieve excellent leachate treatment efficiency. Typically, the equipment was designed with multiple sets of alternating anodes and cathodes, significantly increasing the reaction efficiency and decreasing hydraulic retention time [99]. The electric energy consumption of flow-through mode using porous membrane electrode is 6 times less than that of the traditional flow-by mode due to better mass transfer and more active sites [100]. In addition, the packed-bed composed of many micro-particles can greatly expand the reactive active sites. For example, compared with non-packed system, the packed-bed equipment achieved 1.25 times higher phenol oxidation efficiency at 40 mA/cm2 for 2.72 h [97].

    In practical engineering, EO is often combined with other technologies to fully leverage its efficient degradation advantages and reduce treatment cost. For example, coagulation pretreatment reduced the COD of leachate from 2550 mg/L to 938 mg/L, and EO further degrades it to 88.4 mg/L. The energy consumption of “coagulation + EO” process is 62.26% of EO [101]. The combination of electro-Fenton (pretreatment) and EO for leachate treatment can reduce energy consumption from 218.75 (COD, EO) to 120 kWh/kg [102]. Besides, EO can be combined with cost-effective biological technologies. For instance, EO can decompose the biologically toxic organics in leachate, increasing the BOD5/COD ratio from 0.04 to 0.37 [103]. The cost of EO for leachate treatment is 55.8 €/m3, while it is only 18.7 €/m3 for “EO + Sequencing Batch Biofilter Granular Reactor (SBBGR)” [104]. Therefore, EO should be combined with other technologies to achieve excellent efficiency and low energy cost in practical application.

    EO performance in leachate is influenced by operating parameters such as current density, electrolyte type, pH, and temperature (Table S5 in Supporting information). These factors affect the type, concentration, longevity, and oxidation ability of active species, as well as the state and degradation pathways of pollutants (e.g., HA) [21,56,105113].

    At a given current density, substances in the solution can be activated by the active sites on the anode surface, generating OACs that degrade pollutants in the leachate [114]. Leachate EO is an electron transfer process; the higher the current density, the more intense the redox reactions. Based on the above reaction mechanism discussion, the influence pathway of current density on leachate treatment is involved in the OACs generation. In leachate oxidation, when the current density was increased from 30 mA/cm² to 80 mA/cm², free chlorine production increased by 2.5 times, resulting in an increase in organic nitrogen removal from 17 mg N L-1 h-1 to 66 mg N L-1 h-1 [115]. Kim et al. found that the generation of OH at 30 mA/cm2 is twice that of 10 mA/cm2 using cobalt-promoted lead dioxide anode [116].

    However, excessive current density can trigger halogenation, leading to the formation of chlorinated organic compounds with significant biological toxicity, such as 1,2-dichloroethane, haloacetonitriles, and haloethanes [46]. Moreover, higher current density leads to increased energy consumption due to the enhancement of side reactions, such as oxygen or chlorine evolution (Eqs. 32-34) [117]. Using BDD anode to oxidize leachate, it was found that when the current density increased from 1 mA/cm2 to 30 mA/cm2, the effective utilization rate of electrical energy decreased from 100% to 57.8% [115]. Choosing an adequate current density is critical for encouraging leachate degradation, enhancing economic advantages, and avoiding secondary contamination. In conclusion, the selected range of current density is relatively broad (10–100 mA/cm2) and should be chosen based on the desired effluent standards and control of toxic by-products.

    $ \mathrm{HClO}+\mathrm{Cl}^{-}+\mathrm{H}^{+} \rightarrow \mathrm{Cl}_2+\mathrm{H}_2 \mathrm{O} $

    (32)

    $ 2 \mathrm{Cl}^{-} \rightarrow \mathrm{Cl}_2+2 \mathrm{e}^{-} $

    (33)

    $ 2 \mathrm{H}_2 \mathrm{O} \rightarrow 4 \mathrm{H}^{+}+\mathrm{O}_2+4 \mathrm{e}^{-} $

    (34)

    In EO, the electrolyte not only serves as the medium for electron transfer, but also provides the essential materials for the formation of OACs. Different electrolytes influence the types of active species and the degradation pathways of pollutants.

    The Cl- in leachate is the key content for nitrogen removal. The Cl derivatives generated from Cl- electrooxidation can quickly capture electrons from NH4+ and simultaneously selectively convert it into N2, achieving total nitrogen removal. Cl and its derivatives can remove 98.62% of NH4+ at a concentration of 590 mg/L after 3 h at 125 mA/cm2 [118]. Cabeza et al. found that when Cl-:NH4+ molar ratio exceeds 9:1, NH4+ removal efficiency can reach over 88.51%. Below this ratio, NH4+ removal efficiency significantly decreases [13]. Cifcioglu-Gozuacik et al. reported that when the Cl-:NH4+ ratio is 6:1, NH4+ removal efficiency is only 40% [119]. The NH4+ removal efficiencies are 96.5%, 78.8%, and 61.0% when the Cl-:NH4+ ratios are 9:1, 6:1, and 4.5:1, respectively [120]. The Cl-:NH4+ ratio above 9:1 is recommended to maintain a removal efficiency over 80%. Cl- also plays an important role in leachate decolorization. This is attributed to the rapid reaction between Cl derivatives and the π-π bonds in chromophoric groups of polymeric organic compounds [121].

    The SO42- is another anion in leachate. Theoretically, the SO42- can be oxidized to SO4•- (Eqs. 16 and 35) through electron transfer from anode and OH, promoting organic removal. For example, compared to NO3- electrolyte, the removal rate of dissolved organic carbon (DOC) in SO42- increased from 41% to 51% through electrooxidation due to the generation of SO4•- [122]. However, SO42- is an oxygen-containing ion that will induce oxygen evolution side reactions [113]. The hydrated ionic radius of SO42- (2.40 Å) is larger than that of Cl- (1.81 Å), enabling SO42- to adsorb more easily onto the anode surface, occupying electroactive sites and inhibiting the formation of other active species [123,124].

    $ \mathrm{HSO}_4^{-}+{ }^{•} \mathrm{OH} \rightarrow \mathrm{SO}_4^{•-}+\mathrm{H}_2 \mathrm{O} $

    (35)

    Aged leachate also contains Ca2+ (200 mg/L) and Mg2+ (100 mg/L), which can easy form hydroxide precipitate under alkaline condition (pH ≥ 10.3) [125]. This precipitate will adhere to the anode surface, passivating the active sites. For example, 2,4-dichlorophenol removal efficiency decreased from 67.8% to 33.1% due to the passivation of active sites by Ca2+, Mg2+ [126]. The interface pH at cathode is as higher as 12 even when the solution pH is 7.7–7.9 [127]. To solve the influence of Ca2+ and Mg2+ on electrochemical process, several technologies have been developed. The crystallization fluidized bed effectively reduced the Ca2+ concentration from 537.81 mg/L to 78.43 mg/L [128]. The chemical precipitation controlled Mg2+ concentration below 10 mg/L [129]. Plate and frame filter press and electrodeposition are also efficient for removing Ca2+ and Mg2+.

    Leachate pH varies at different landfill stages. New leachate (0–5 years), rich in volatile fatty acids (VFAs) from anaerobic fermentation, has a pH < 6.5. As VFAs escape or biodegrade, the pH of intermediate leachate (5–10 years) rises to 6.5–7.5. Aged leachate (>10 years), with a pH > 7.5, primarily contains refractory organics like HA [14]. Leachate treatment with EO primarily relies on indirect oxidation mediated by OACs, whose generation is greatly influenced by pH.

    The generation of OH is related to the pH. The oxygen evolution reaction (OER) is the main side reaction for OH production, with its trigger potential decreasing as pH increases (Eqs. 36 and 37) [130]. The cyclic voltammetry (CV) demonstrates that OER redox peak shifts to lower potential at higher pH [131]. The high rate of oxygen evolution on electrode surface will occupy reactive sites for OH generation, reducing the organics removal [132]. The high oxygen evolution potential (1.229 V vs. SHE) in acidic or neutral conditions will facilitate OH generation [130,133], facilitating organics oxidation. In an EO experiment, the removal rate of chloramphenicol at pH 3 is about 2.4 times higher than at pH 11 [134]. Additionally, under acidic conditions, SO4•- is not quenched by hydrolysis (Eq. 31). At pH ≥ 9, SO4•- can react with OH- to form OH (Eq. 30), which then reacts with OH- (Eq. 4) to produce oxygen, quenching OACs.

    $ 4 \mathrm{OH}^{-} \rightarrow 2 \mathrm{H}_2 \mathrm{O}+\mathrm{O}_2+4 \mathrm{e}^{-} $

    (36)

    $ 2 \mathrm{H}_2 \mathrm{O} \rightarrow 4 \mathrm{H}^{+}+\mathrm{O}_2+4 \mathrm{e}^{-} $

    (37)

    Active chlorine is key OACs for NH4+ removal in leachate. At different pH levels, active chlorine exists as Cl2 (pH ≤ 3), HClO (pH 5–7), ClO- (pH ≥ 8) [36]. The NH4+ removal rate is much higher in alkaline than in acidic solutions. Vlyssides et al. reported nearly 100% NH4+ removal through EO at pH 9, compared to only 30% at pH 6 [135]. Importantly, although the active chlorine can directly oxidize NH4+, the products depend on the reaction solution pH [135,136]. In acid condition, NH4+ is easily oxidized to NO3- (by-product) by Cl2, while it is selectively converted to N2 by ClO- under alkaline environment. For example, after 90 min of EO at pH 6, the NO3- concentration reaches 118.2 mg N/L from an initial NH4+ concentration of 1094 mg N/L. In contrast, at pH 9, the NO3- concentration accumulates to only 7.2 mg N/L under the same conditions [135]. The pH of aged leachate typically ranges from 8 to 10, creating an ideal environment for NH4+ removal through EO.

    The presence status of weak acid ion (CO32- and HCO3-) and pollutants in leachate are also affected by the solution pH. The CO32- and HCO3- are quenching agent for OH, Cl and Cl2•-. They will be transferred into CO2 under acidic or neutral conditions (pH < 8.3) especially pH < 4, improving the effective utilization of free radicals OACs. Besides, HA is the main organic pollutant in leachate particularly for aged leachate. It exhibits a flocculent state under acidic conditions (pH < 3.5), covering active sites and hindering pollutant diffusion [137]. For example, for real leachate with an initial COD of 42,000 mg/L, flocculation occurred at pH 6 and the corresponding COD removal efficiency reached 81.9% at pH 4.5 [138]. Besides, H+ induces protonation of carboxyl and phenolic groups in organic, leading to zeta potential increase and hydrophilicity decrease [139]. For example, at pH 2, zeta potential approached 0 mV and 86% of COD in leachate mainly removed by flocculation process [138]. At pH > 6.5, carboxyl and phenolic groups are in an ionic state and form opposite charge electric layers [140]. The eta potential is less than −14.7 mV, improving the flocs resolution and HA dispersion [138]. Therefore, EO for leachate degradation under neutral or weak alkaline conditions is a suitable option.

    High temperatures are conducive to electron transfer, mass transport, ionic strength and conductivity, improving the leachate treatment [56,141]. For example, raising the temperature from 35 ℃ to 55 ℃ improves the removal efficiency of total organic carbon (TOC) and NH4+ from 59.7% and 88.8% to 70.2% and 99.9%, respectively [142]. Adjusting the temperature from 30 ℃ to 60 ℃, the decolorization rate increases from 40% to 70% [113]. Electrode activity is also significantly affected by temperature involving the electroactive area and current density [143]. When the temperature increased from 25 ℃ to 60 ℃, the current density of Pt electrode increased by 42.85% [143]. However, high temperature could also lead to an increase of by-products in electrochemical leachate treatment, especially in the presence of Cl-. For example, the concentration of disinfection by-products (DBP) increased by 50% when the reaction temperature rose from 28 ℃ to 35 ℃ [144]. Perchlorate generation from Cl- oxidation at 20 ℃ was 1.5 times higher than at 6 ℃ [145].

    Electric energy can also be converted into thermal energy, leading to a temperature increase during the EO process of leachate. Sun et al. found that the leachate temperature rose from 23 ℃ to 50 ℃ after electrolysis 3 h [142]. However, in practice, heating large volumes of leachate will increase costs and by-product formation. Therefore, it is recommended to conduct leachate EO at ambient temperatures (15–35 ℃) and utilize the inevitable temperature increase from electrical energy losses to heat the wastewater.

    This review analyzed the main pathways for landfill leachate EO treatment and focused on the mechanisms of OACs generation and transformation for pollutant removal. In addition, recently developed countermeasures for enhancing electrode performance in landfill leachate treatment have been discussed. The introduction of a multifunctional intermediate layer between the substrate and catalyst improves the stability and conductivity of the anode. To accelerate mass and electron transfer, two approaches have been developed: designing porous membrane electrodes and incorporating particle microelectrodes. Regulating the catalyst microstructure can provide sufficient active sites for OACs generation and pollutant removal. Key parameters in the EO of leachate were also discussed to further optimize the application of this technology.

    However, technical barriers still limit the commercialization of each strategy. The following key areas for future research are proposed: (1) Research on EO for leachate degradation should shift from single pollutants (e.g., HA) to complex pollutant systems and from simulated wastewater to actual leachate, thereby advancing the development of high-performance anodes. The cost of electrode preparation and their actual service life need to be further evaluated. Additionally, the impact of Ca2+ and Mg2+ in leachate on electrode passivation, pollution control efficiency, and solution method needs further study. (2) Although EO can efficiently treat landfill leachate, it should be combined with other technologies in practice to increase treatment efficiency and reduce treatment costs. Additionally, the toxic by-products of leachate EO, especially their formation mechanisms and quantitative analysis, is limited. (3) Leachate contains various substance, leading to multiple active species during electrooxidation. Most researches focus on the removal of macroscopic indicators like COD/TOC, color, and NH4+, with limited study on the degradation characteristics and pathways of specific pollutants. 4) Most studies on leachate EO have been conducted at the laboratory stage. More focus is needed on practical applications, particularly the design and development of EO equipment.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Zhenhui Song: Writing – review & editing, Writing – original draft, Visualization, Validation, Methodology, Investigation, Formal analysis, Data curation, Conceptualization. Xing Wu: Validation, Methodology, Investigation, Data curation. Tianyu Gao: Visualization, Methodology, Investigation. Fubing Yao: Writing – review & editing, Visualization, Validation, Investigation, Funding acquisition, Formal analysis, Conceptualization. Xi Tang: Validation, Methodology, Investigation, Conceptualization. Qaisar Mahmood: Writing – review & editing, Visualization, Validation. Chong-Jian Tang: Writing – review & editing, Validation, Supervision, Resources, Project administration, Funding acquisition, Formal analysis, Conceptualization.

    This research was financially supported by the Innovative Research Groups of the National Natural Science Foundation of China (No. 52121004), the project of National Natural Science Foundation (No. U21A20294), the Scientific Foundation of Hunan Province (Nos. 2024RC1012, 2022JJ40622), The Technology Innovation Guidance Project of Jiangxi Province, China (Nos. 20203BDH80W017, 20212BDH81030).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2025.111008.


    1. [1]

      S. Ma, C. Zhou, J. Pan, et al., J. Clean. Prod. 333 (2022) 130234.

    2. [2]

      J. Wiszniowski, D. Robert, J. Surmacz-Gorska, K. Miksch, J.V. Weber, Environ. Chem. Lett. 4 (2006) 51–61. doi: 10.1007/s10311-005-0016-z

    3. [3]

      Z. Guo, Y. Zhang, H. Jia, et al., Sci. Total Environ. 806 (2022) 150529.

    4. [4]

      D. Liu, Y. Yuan, Y. Wei, et al., J. Environ. Sci. 116 (2022) 43–51.

    5. [5]

      N. Aishi, D. Animesh, Mater. Today Proc. 67 (2022) 1290–1297.

    6. [6]

      Y. Deng, C.M. Ezyske, Water Res. 45 (2011) 6189–6194.

    7. [7]

      C. Qu, L. Li, F. Feng, et al., Front. Environ. Sci. Eng. 17 (2023) 115–126.

    8. [8]

      Z. Ma, Y. Yang, Y. Jiang, et al., Chem. Eng. J. 311 (2017) 183–190.

    9. [9]

      M. Umar, H.A. Aziz, M.S. Yusoff, Waste Manag. 30 (2010) 2113–2121.

    10. [10]

      G. Boczkaj, A. Fernandes, Chem. Eng. J. 320 (2017) 608–633.

    11. [11]

      I. Hussain, Y. Zhang, S. Huang, RSC Adv. 4 (2014) 3502–3511.

    12. [12]

      H. Mohammadi, B. Bina, A. Ebrahimi, Process Saf. Environ. Prot. 117 (2018) 200–213.

    13. [13]

      A. Cabeza, A. Urtiaga, M.J. Rivero, I. Ortiz, J. Hazard. Mater. 144 (2007) 715–719.

    14. [14]

      O.T. Can, L. Gazigil, R. Keyikoglu, Environ. Prog. Sustain. Energy 41 (2022) 13722.

    15. [15]

      Y. Zhang, Z. Zheng, H. Liu, et al., J. Chongqing Univ. 46 (2022) 76–88.

    16. [16]

      B.K.D. Pubali Mandal, A.K Gupta, Waste Manag. 69 (2017) 250–273.

    17. [17]

      S.L. He, Q. Huang, Y. Zhang, Y.L. Nie, Water Air Soil Pollut. 226 (2015) 1–7. doi: 10.1111/age.12236

    18. [18]

      M. Fukushima, K. Tatsumi, S. Nagao, Environ. Sci. Technol. 17 (2001) 3683–3690.

    19. [19]

      F. Feng, Z. Liu, X. Tang, et al., Water Res. 229 (2023) 119393.

    20. [20]

      J. Li, Z. Yang, H. Xu, et al., RSC Adv. 6 (2016) 47509–47519.

    21. [21]

      P. Mandal, A.K. Gupta, B.K. Dubey, J. Water Process Eng. 33 (2020) 101119.

    22. [22]

      B. Biswas, S. Goel, Chemosphere 302 (2022) 134709.

    23. [23]

      Y.G. Asfaha, A.K. Tekile, F. Zewge, Clean. Eng. Technol. 4 (2021) 100261.

    24. [24]

      S.B.M.S.S. Movahed, Environ. Sci. Pollut. Res. 28 (2021) 59594–59607.

    25. [25]

      M.J. Park, T. Lee, J. Korean Soc. Water Sci. Technol. 28 (2020) 31–38. doi: 10.17640/kswst.2020.28.3.31

    26. [26]

      R. Fu, P.-S. Zhang, Y.X. Jiang, L. Sun, X.H. Sun, Chemosphere 311 (2022) 136993.

    27. [27]

      C. Tang, W. Yan, C. Zheng, Front. Environ. Sci. Eng. 8 (2014) 337–344. doi: 10.1007/s11783-013-0545-9

    28. [28]

      F. Abdelmalek, R.A. Torres, E. Combet, et al., Sep. Purif. Technol. 63 (2008) 30–37.

    29. [29]

      L. Weihua, M. Ji, X. Zhang, J. Yang, Environ. Chem. 26 (2007) 58–61.

    30. [30]

      N. Sergienko, E.C. Lumbaque, N. Duinslaeger, J. Radjenovic, Appl. Catal. B: Environ. 334 (2023) 122831.

    31. [31]

      L. Trotochaud, S.W. Boettcher, Scripta Mater. 74 (2014) 25–32.

    32. [32]

      R. Zhang, J. Cao, T. Peng, K. Wu, Y. Shu, Electrochim. Acta 487 (2024) 144166.

    33. [33]

      Y. Zhang, W. Tang, J. Bai, et al., J. Hazard. Mater. 424 (2022) 127662.

    34. [34]

      Y. Deng, J.D. Englehardt, Waste Manag. 27 (2007) 380–388.

    35. [35]

      A. Thiam, E. Brillas, F. Centellas, P.L. Cabot, I. Sirés, Electrochim. Acta 173 (2015) 523–533.

    36. [36]

      B. Xu, S.M. Iskander, Z. He, Environ. Res. 182 (2020) 109006.

    37. [37]

      L.C. Chiang, J.E. Chang, T.C. Wen, Water Res. 29 (1995) 671–678.

    38. [38]

      J. Radjenovic, M. Petrovic, Water Res. 94 (2016) 128–135.

    39. [39]

      A.N. Arenhart Heberle, M. Garcia-Gabaldon, E.M. Ortega, A.M. Bernardes, V. Perez-Herranz, Chemosphere 236 (2019) 124318.

    40. [40]

      K.C. de Freitas Araujo, D.R. da Silva, E.V. dos Santos, H. Varela, C.A. Martinez-Huitle, J. Electroanal. Chem. 860 (2020) 133927.

    41. [41]

      J. Cai, T. Niu, P. Shi, G. Zhao, Small 15 (2019) 1900153.

    42. [42]

      S.O. Ganiyu, C.A. Martínez-Huitle, M.A. Oturan, Curr. Opin. Electrochem. 27 (2021) 100678.

    43. [43]

      L. Chen, C. Lei, Z. Li, et al., Chemosphere 210 (2018) 516–523.

    44. [44]

      D.D. Kiper, M. Qui, G. Passard, C. Costentin, D. Nocera, Abstr. Pap. Am. Chem. Soc. 258 (2019) 8671–8679.

    45. [45]

      J. Radjenovic, D.L. Sedlak, Environ. Sci. Technol. 49 (2015) 11292–11302. doi: 10.1021/acs.est.5b02414

    46. [46]

      P. Devi, U. Das, A.K. Dalai, Sci. Total Environ. 571 (2016) 643–657.

    47. [47]

      Z. Gu, W. Chen, Q. Li, A. Zhang, Process Saf. Environ. Prot. 133 (2020) 32–40.

    48. [48]

      L. Cui, Y. Zhang, K. He, M. Sun, Z. Zhang, Sep. Purif. Technol. 293 (2022) 121112.

    49. [49]

      T. Maqbool, Q.V. Ly, K. He, et al., J. Membr. Sci. 651 (2022) 120460.

    50. [50]

      L. Pinhedo, R. Pelegrini, R. Bertazzoli, A. Motheo, Appl. Catal. B: Environ. 57 (2005) 75–81.

    51. [51]

      H.-R. Schulten, Fresenius J. Anal. Chem. 351 (1995) 62–73.

    52. [52]

      M. Zhou, Z. Wu, Y. Cong, Q. Ye, D. Wang, China Environ. Sci. 23 (2003) 225–229.

    53. [53]

      Q. Qiao, S. Singh, S.L. Lo, et al., Chemosphere 275 (2021) 129848.

    54. [54]

      T. Xu, X. Xiao, H. Liu, China Environ. Sci. 24 (2004) 547–551.

    55. [55]

      D. Pang, Y. Liu, H. Song, et al., Chem. Eng. J. 405 (2021) 126982.

    56. [56]

      M. Panizza, M. Delucchi, I. Sirés, J. Appl. Electrochem. 40 (2010) 1721–1727. doi: 10.1007/s10800-010-0109-7

    57. [57]

      L. Tran Le, Environ. Technol. Innov. 20 (2020) 101099.

    58. [58]

      Z. Zhang, J. Liu, H. Ai, et al., J. Environ. Chem. Eng. 11 (2023) 109834.

    59. [59]

      S. Liu, R. Liu, Y. Zhang, et al., Chemosphere 237 (2019) 124471.

    60. [60]

      M.J. Nunes, N. Monteiro, M.J. Pacheco, A. Lopes, L. Ciríaco, J. Environ. Sci. Health Part A 51 (2016) 839–846. doi: 10.1080/10934529.2016.1181455

    61. [61]

      Y. Wang, M. Chen, C. Wang, et al., Chem. Eng. J. 374 (2019) 626–636.

    62. [62]

      I. Elaissaoui, H. Akrout, S. Grassini, D. Fulginiti, L. Bousselmi, Mater. Des. 110 (2016) 633–643.

    63. [63]

      C. Tang, Z. Liu, D. Cui, et al., Electrochim. Acta 399 (2021) 139398.

    64. [64]

      S. Man, D. Luo, Q. Sun, et al., J. Hazard. Mater. 430 (2022) 128440.

    65. [65]

      G. Zhang, L. Zhao, X. Hu, X. Zhu, F. Yang, Appl. Catal. B: Environ. 313 (2022) 121453.

    66. [66]

      M.X. Qiao, Y. Zhang, L.F. Zhai, M. Sun, Chem. Eng. J. 344 (2018) 410–418.

    67. [67]

      C.A. Martinez-Huitle, M.A. Rodrigo, I. Sires, O. Scialdone, Chem. Rev. 115 (2015) 13362–13407. doi: 10.1021/acs.chemrev.5b00361

    68. [68]

      H. Lin, R. Xiao, R. Xie, et al., Environ. Sci. Technol. 55 (2021) 2597–2607. doi: 10.1021/acs.est.0c06881

    69. [69]

      Z. Pan, C. Song, L. Li, et al., Chem. Eng. J. 376 (2019) 120909.

    70. [70]

      B.P. Chaplin, Environ. Sci. Process. Impacts 16 (2014) 1182–1203.

    71. [71]

      P. Gayen, J. Spataro, S. Avasarala, et al., Environ. Sci. Technol. 52 (2018) 9370–9379. doi: 10.1021/acs.est.8b03038

    72. [72]

      L. Guo, Y. Jing, B.P. Chaplin, Environ. Sci. Technol. 50 (2016) 1428–1436. doi: 10.1021/acs.est.5b04366

    73. [73]

      L. Wang, J. Lu, L. Li, Y. Wang, Q. Huang, Water Res. 170 (2020) 115254.

    74. [74]

      O. Ganzenko, P. Sistat, C. Trellu, et al., Chem. Eng. J. 419 (2021) 129467.

    75. [75]

      S. Liu, Y. Wang, X. Zhou, et al., Electrochim. Acta 253 (2017) 357–367.

    76. [76]

      H. Lin, H. Peng, X. Feng, et al., Water Res. 190 (2021) 116790.

    77. [77]

      M. Martí-Calatayud, E. Dionís, S. Mestre, V. Pérez-Herranz, J. Clean. Prod. 363 (2022) 132342.

    78. [78]

      A. Chen, S. Xia, H. Pan, et al., J. Electroanal. Chem. 824 (2018) 169–174.

    79. [79]

      L. Chang, Y. Zhou, X. Duan, W. Liu, D. Xu, J. Taiwan Inst. Chem. Eng. 45 (2014) 1338–1346.

    80. [80]

      Y. Qin, T. Yu, S. Deng, et al., Nat. Commun. 13 (2022) 1–8.

    81. [81]

      S. Anuchai, S. Phanichphant, D. Tantraviwat, et al., J. Colloid Interface Sci. 512 (2018) 105–114.

    82. [82]

      J. Xie, J. Ma, C. Zhang, et al., Environ. Sci. Technol. 54 (2020) 5227–5236. doi: 10.1021/acs.est.9b07398

    83. [83]

      W. Li, R. Xiao, J. Xu, et al., Water Res. 216 (2022) 118287.

    84. [84]

      W. Li, R. Xiao, H. Lin, et al., J. Hazard. Mater. 424 (2022) 127342.

    85. [85]

      S.S.P. Rahardjo, Y.J. Shih, Chem. Eng. J. 452 (2023) 139370.

    86. [86]

      H. Xie, H. Zhang, X. Wang, et al., Front. Environ. Sci. Eng. 18 (2024) 3–15.

    87. [87]

      J. Cai, M. Zhou, Y. Pan, X. Du, X. Lu, Appl. Catal. B: Environ. 257 (2019) 117902.

    88. [88]

      M. Pierpaoli, P. Jakobczyk, M. Sawczak, et al., J. Hazard. Mater. 401 (2021) 123407.

    89. [89]

      T. Lim, G.Y. Jung, J.H. Kim, et al., Nat. Commun. 11 (2020) 1–11.

    90. [90]

      P. Jakóbczyk, G. Skowierzak, I. Kaczmarzyk, et al., Chemosphere 304 (2022) 135381.

    91. [91]

      M. Zhou, W. Fu, H. Gu, L. Lei, Electrochim. Acta 52 (2007) 6052–6059.

    92. [92]

      C. Zhang, Y. Jiang, Y. Li, et al., Chem. Eng. J. 228 (2013) 455–467.

    93. [93]

      G. Lv, D. Wu, R. Fu, J. Hazard. Mater. 165 (2009) 961–966.

    94. [94]

      W. Can, H. Yao-Kun, Z. Qing, J. Min, Chem. Eng. J. 243 (2014) 1–6.

    95. [95]

      L. Yan, H. Ma, B. Wang, Y. Wang, Y. Chen, Desalination 276 (2011) 397–402.

    96. [96]

      Y. Chen, W. Shi, H. Xue, et al., Electrochim. Acta 58 (2011) 383–388.

    97. [97]

      P. Li, Y. Zhao, L. Wang, et al., Electrochemistry 82 (2014) 1056–1060. doi: 10.5796/electrochemistry.82.1056

    98. [98]

      Y. Deng, X. Zhu, N. Chen, et al., Sci. Total Environ. 745 (2020) 140768.

    99. [99]

      S. Veli, A. Arslan, M. Isgoren, D. Bingol, D. Demiral, Environ. Chall. 5 (2021) 100369.

    100. [100]

      J.F. Perez, J. Llanos, C. Saez, et al., Chem. Eng. J. 351 (2018) 766–772.

    101. [101]

      L.C. Chiang, J.E. Chang, C.T. Chung, Environ. Eng. Sci. 18 (2001) 369–379. doi: 10.1089/109287501753359609

    102. [102]

      A. Urtiaga, A. Rueda, A. Anglada, I. Ortiz, J. Hazard. Mater. 166 (2009) 1530–1534.

    103. [103]

      Y.Y. Chu, M.H. Zhu, C. Liu, Environ. Eng. Sci. 32 (2015) 445–450. doi: 10.1089/ees.2014.0433

    104. [104]

      G. Del Moro, L. Prieto-Rodriguez, M. De Sanctis, et al., Chem. Eng. J. 288 (2016) 87–98.

    105. [105]

      P.B. Moraes, R. Bertazzoli, Chemosphere 58 (2005) 41–46.

    106. [106]

      T.L. Luu, Environ. Technol. Innov. 20 (2020) 101099.

    107. [107]

      D. Yu, Y. Pei, J. Environ. Manage. 321 (2022) 115890.

    108. [108]

      X.-M. Li, China Water Wastewater 17 (2001) 14–17.

    109. [109]

      M. Panizza, C.A. Martinez-Huitle, Chemosphere 90 (2013) 1455–1460.

    110. [110]

      T. Cui, Y. Zhang, W. Han, et al., Chem. Eng. J. 315 (2017) 335–344.

    111. [111]

      S.Y. Guvenc, Y. Daniser, E. Can-Güven, G. Varank, A. Demir, Environ. Eng. Res. 28 (2023) 210419.

    112. [112]

      K. Pan, M. Tian, Z.H. Jiang, B. Kjartanson, A. Chen, Electrochim. Acta 60 (2012) 147–153.

    113. [113]

      E. Turro, A. Giannis, R. Cossu, et al., J. Hazard. Mater. 190 (2011) 460–465.

    114. [114]

      C. Zhang, M. Zhou, G. Ren, et al., Water Res. 70 (2015) 414–424.

    115. [115]

      H. Feng, Z. Chen, X. Wang, S. Chen, J. Crittenden, Chem. Eng. J. 413 (2021) 921–928. doi: 10.1111/poms.13285

    116. [116]

      J. Kim, G.V. Korshin, Ozone Sci. Eng. 30 (2008) 113–119.

    117. [117]

      Y. Sun, P. Li, H. Zheng, et al., Chem. Eng. J. 308 (2017) 1233–1242.

    118. [118]

      T. Hasnine, E.C. Lumbaque, Q. Yuan, Environ. Technol. 39 (12) (2023) 2687–2703.

    119. [119]

      B. Cifcioglu-Gozuacik, S.M. Ergenekon, B. Ozbey-Unal, et al., Water Sci. Technol. 84 (2021) 752–762. doi: 10.2166/wst.2021.261

    120. [120]

      P. Mandal, M.K. Yadav, A.K. Gupta, B.K. Dubey, Sep. Purif. Technol. 247 (2020) 116910.

    121. [121]

      H.J. Fan, H.Y. Shu, H.S. Yang, W.C. Chen, Sci. Total Environ. 361 (2006) 25–37.

    122. [122]

      A.Y. Bagastyo, D.J. Batstone, K. Rabaey, J. Radjenovic, Water Res. 47 (2013) 242–250.

    123. [123]

      Q. Lv, X. Zang, X. Li, G. Li, Fluid Phase Equilib. 458 (2018) 272–277.

    124. [124]

      A.Y. Bagastyo, D.J. Batstone, K. Rabaey, J. Radjenovic, Water Res. 47 (2013) 242–250.

    125. [125]

      R. Zhao, J. Chen, J. Liu, M. Li, H. Yang, Water Sci. Technol. 87 (2023) 366–380. doi: 10.2166/wst.2023.002

    126. [126]

      L. Hu, L. Shi, F. Shen, et al., Water Res. 225 (2022) 119210.

    127. [127]

      R. Mao, X. Zhao, H. Lan, H. Liu, J. Qu, Water Res. 77 (2015) 1–12.

    128. [128]

      L. Rui, L. Guoyuan, Z. Shunming, et al., Ind. Water Treat. 14 (2024) 12–25. doi: 10.21656/1000-0887.440186

    129. [129]

      Z. Wang, Z. Cui, D. Zhao, et al., China Energy Environ. Prot. 46 (2024) 155–161.

    130. [130]

      J.W. Lee, B. Kim, J.Y. Seo, et al., Appl. Surf. Sci. 610 (2023) 1–6.

    131. [131]

      K.A. Stoerzinger, R.R. Rao, X.R. Wang, et al., Chem 2 (2017) 668–675.

    132. [132]

      A.S. Koparal, E. Onder, U.B. Ogutveren, Desalination 197 (2006) 262–272.

    133. [133]

      X. Duan, C. Su, L. Zhou, et al., Appl. Catal. B: Environ. 194 (2016) 7–15.

    134. [134]

      J. Chen, Y. Xia, Q. Dai, Electrochim. Acta 165 (2015) 277–287.

    135. [135]

      A. Vlyssides, P. Karlis, M. Loizidou, A. Zorpas, D. Arapoglou, Environ. Technol. 22 (2001) 1467–1476.

    136. [136]

      2 B. Zhou, Z. Yu, Q. Wei, et al., Appl. Surf. Sci. (2016) 406–415.

    137. [137]

      R.A. Alvarez-Puebla, J.J. Garrido, Chemosphere 59 (2005) 659–667.

    138. [138]

      M. Laura Pinedo, B.D. Riascos, X.E. Quintero, C. Costa, Waste Manag. 144 (2022) 163–172.

    139. [139]

      T.J. Manning, T. Bennett, D. Milton, Sci. Total Environ. 257 (2000) 171–176.

    140. [140]

      F. de Souza, S.R. Braganca, Mater. Res. 21 (2018) 0759–0766.

    141. [141]

      A. Anglada, A.M. Urtiaga, I. Ortiz, J. Hazard. Mater. 181 (2010) 729–735.

    142. [142]

      D. Sun, X. Hong, Z. Cui, et al., J. Hazard. Mater. 388 (2020) 121768.

    143. [143]

      H.A. Gasteiger, N. Marković, P.N. Ross, E.J. Cairns, J. Electrochem. Soc. 141 (1994) 1795–1803. doi: 10.1149/1.2055007

    144. [144]

      X. Liu, Z. He, Sci. Total Environ. 730 (2020) 139171.

    145. [145]

      N. Ambauen, C. Weber, J. Muff, C. Halle, T. Meyn, J. Appl. Electrochem. 50 (2020) 1175–1188. doi: 10.1007/s10800-020-01476-3

  • Figure 1  Schematic illustration for the EO process. R denotes an organic pollutant and M is the active site on the electrode surface.

    Figure 2  Sulfate radical-mediated oxidation and anode-cathode synergy.

    Figure 3  Diagram of the anode with intermediate layer.

    Figure 4  Diagram of the porous membrane anode in flow-through mode.

    Figure 5  Diagram of anode with three-dimensional microstructure.

    Figure 6  Diagram of the packed bed reactor with particle microelectrodes.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  22
  • HTML全文浏览量:  8
文章相关
  • 发布日期:  2025-12-15
  • 收稿日期:  2024-10-31
  • 接受日期:  2025-02-26
  • 修回日期:  2025-01-13
  • 网络出版日期:  2025-02-26
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章