Next Article in Journal
Evaluation of Nitrogen and Carbon Stable Isotopes in Filter Feeding Bivalves and Surficial Sediment for Assessing Aquatic Condition in Lakes and Estuaries
Previous Article in Journal
Remote Sensing Retrieval of Total Nitrogen in the Pearl River Delta Based on Landsat8
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Reactor Design for Industrial Wastewater Treatment by Electro-Oxidation Process

1
School of Space & Environment, Beijing Key Lab Bioinspired Energy Materials & Devices, Beihang University, Shahe Campus, Beijing 102206, China
2
China ENFI Engineering Corp., ENFI Equipment Branch, No.12 Fuxing Road, Haidian District, Beijing 100038, China
*
Author to whom correspondence should be addressed.
Water 2022, 14(22), 3711; https://doi.org/10.3390/w14223711
Submission received: 31 October 2022 / Revised: 14 November 2022 / Accepted: 15 November 2022 / Published: 16 November 2022
(This article belongs to the Special Issue Wastewater Treatment and Resource Recovery: Advances and Developments)

Abstract

:
Refractory organic wastewater mainly includes wastewater from papermaking, textile, printing and dyeing, petrochemical, coking, pharmaceutical and other industries, as well as landfill leachate and its membrane-treated concentrate. The traditional biochemical method is difficult to adapt to its harsh conditions such as high toxicity, high organic load and high salinity. Compared to other methods, the electro-oxidation (EO) process owns the attractive characteristics of being clean and eco-friendly, highly efficient and producing no secondary pollution. EO systems mainly include electrodes, a reactor, a power supply and other basic units. The design of reactors with different electrodes was the key link in the application of EO technology. This paper mainly reported the different configurations of electrochemical reactors (ECRs) for refractory organic wastewater treatment, and summarized the advantages and disadvantages of them, including reactor structure, flow mode, operation mode and electrode construction. Compared with traditional reactors, the improved reactors such as 3D-ECR achieve higher mass transfer efficiency by increasing the contact area between the electrode and the fluid. Additionally, it has a higher removal rate of organics and a lower energy consumption. Finally, the future perspectives of the treatment of refractory organic wastewater by ECRs is discussed. This paper is expected to provide a reliable scientific basis for the real application of EO technology in refractory organic wastewater treatment.

1. Introduction

Refractory organic wastewater is complex in composition and highly toxic, which has a significant impact on environmental and human health if it cannot be properly disposed of [1]. Therefore, high-efficiency technologies for the treatment of refractory organic wastewater are in urgent need. In many cases, it is difficult to achieve the ideal treatment effect by using the traditional biological method, and it needs to be performed in tandem with physical and chemical pre- and/or post-treatment [2,3]. Thus, the operation process is complex and the cost is relatively high. Over the past years, many advanced oxidation processes (AOPs) for refractory organic wastewater, including ozonation [4], photocatalytic oxidation [5], Fenton [6], electro-Fenton [7], photo-Fenton [8], persulfate-based oxidation [9] and electro-oxidation (EO) [10], have been deeply studied. The emerging AOPs can degrade or even mineralize the refractory organic compounds through hydroxyl radicals produced directly or indirectly [11]. Compared with other methods, the EO process owns the attractive characteristics of being clean and eco-friendly, highly efficient and without secondary pollution [12]. Therefore, EO has been a research hotpot in refractory organic wastewater treatment in the last several decades.
EO could be divided into direct oxidation and indirect oxidation according to the oxidation mechanism as illustrated in Figure 1. Direct oxidation of organic matter occurs on the anode surface by the anodic generation of OH. Indirect oxidation is performed by using the anodic by-products as oxidants, such as OH, O3 and HClO, to react with organic matter in aqueous solution [13].
The obstacle of EO for industrial application lies in the low current efficiency and high energy consumption [14]. Solving these problems could greatly advance EO technology. So far, the research directions of EO in wastewater treatment mainly focus on the advancement of anodic materials, which is targeting the improvement of the current efficiency for OH generation and the stability or service life of electrode. Another focus is on the reactor configuration of EO. The optimization of reactor design could improve mass transfer coefficient, reaction kinetics and reduce energy consumption.
At present, the main anode materials include a metal electrode (e.g., Fe, Al, Pt) [15], graphite electrode [16], dimensional stable anode (DSA) [17] and boron-doped diamond (BDD) electrode in an EO system for wastewater treatment [18]. To overcome the deficiencies presented by the traditional graphite, platinum, lead-based alloy electrode and lead dioxide electrode, a titanium-based metal oxide electrode, also termed as DSA, is the most widely applied and has been developed for three decades [19]. However, there is a trade-off effect for the performance and service life of these electrodes. For example, Ti/RuO2-IrO2 is very stable and has long service life, but the oxidation efficiency is limited, while Ti/SnO2-Sb2O5 and Ti/PbO2 electrodes show great organic degradation performance but short service life limits their practical application [20]. A BDD electrode is a more robust electrode, because of its considerable chemical and electrochemical stability and the high oxygen evolution potential (OEP) to effectively produce OH [21]. However, the price of the BDD electrode is too expensive for large-scale application.
In addition to the anode, the cathode can also remove the organic matter. The traditional cathodic hydrogen evolution reaction (HER) produces H2, which can separate part of the colloidal organic pollutants in water through gas flotation [22]. The most widely used cathodes are stainless steel [23], nickel foam [24], etc. Recently, researchers also resort to carbon-based gas diffusion cathodes to produce H2O2 through the cathodic oxygen reduction reaction (ORR). H2O2 not only can directly degrade organic matter to some extent, but can also react with the anodic byproduct, such as O3, Cl2 or the added chemical Fe2+, to produce OH through the Fenton or Fenton-like oxidation mechanism [25]. Carbon-based materials with benefits of earth-abundant reserves, flexible structures, excellent electrochemical stability and high efficiency were considered as the optimal catalysts for H2O2 generation [26]. At present, carbon-based materials for a H2O2 generation cathode are graphite [27], carbon felt [28], carbon nanotubes (CNTs) [29], carbon/polytetrafluoroethylene (C/PTFE) [30] and a gas diffusion electrode (GDE) [31]. Moreover, GDEs are more popular because they could enhance the utilization efficiency of O2, and even directly use air to produce H2O2 [32]. So far, over 300 review papers on electrode materials for EO technology in wastewater treatment have been published. Thus, this study will not elaborate more on the electrode materials.
Reactor configuration is crucial of electrocatalytic degradation efficiency of pollutants and minimizing energy consumption for EO technology. The configuration design of electrochemical reactors (ECRs) has a direct impact on the fluid flow, mixing affect, heat transfer, mass transfer, etc. It has an especially significant impact on the mass transfer of reactant in solution, thus affecting the kinetic electrochemical reaction rate and electrocatalytic efficiency [33]. The main task of the reactor design is choosing the form and operation mode of ECRs. Then, it is necessary to calculate the fluid speed, temperature, pressure and the reactor volume according to the characteristics of the organic matter and electrochemical reactions. At the same time, economic benefits and environmental protection requirements should also be considered.
In recent years, the design of ECR configuration has improved greatly. The optimization direction is mainly focused on improving the mass transfer efficiency by increasing the contact area between the electrodes and the fluid. However, most of them still stay in the laboratory or pilot scale, which is not conducive to practical application.
Hence, this mini-review focuses on the configuration development and optimization of ECRs, and discusses advantages and disadvantages of different reactor design as well as the treatment efficiency of wastewater in the EO system. It may also provide theoretical support for the reactor design in fluid simulation, put forward the existing technology problems and the development orientation. It is hoped to provide a scientific basis for broadening EO in real applications for refractory organic wastewater treatment.

2. Electrochemical Reactors

For the design of the ECRs, several principles need to be considered. First, ECR design should meet the requirements of electrocatalysis. Second, the assembly of the ECRs should have high feasibility, indicating the reactor configuration, operation and maintenance should be as simple as possible; the electrode materials need to be easily obtained and inexpensive. Last, the reactor should have good versatility and be environmentally friendly [33]. The ECRs will be classified according to reactor structure, operation mode, flow mode and electrode configuration (Figure 2).

3. ECRs with Different Cell Architecture

The box reactor is the simplest structure of ECRs, applied with a plate electrode into a rectangular tank. It is widely applied in industrial electrolysis, organic synthesis, electroplating and other fields. A rounded beaker with nonuniform stirring pattern under the open system is usually utilized for wastewater treatment in lab, as this type of reactor shares the same principle as the box reactor [34]. However, it is difficult to apply a box reactor or rounded beaker on a large-scale, due to the low volume and time efficiency, the flow regime and mass transport effects that cannot be easily determined [35].
To improve the reactor mass transfer and spatiotemporal efficiency, reactors with special structures are developed, such as the plate-frame ECR (filter-press reactor) and tubular ECR. The plate-frame ECR is becoming popular due to its high performance in industrial applications, such as wastewater treatment, electrodialysis, energy storage process and electrosynthesis [36]. Compared with other reactors, the plate-frame ECR has the benefits of low pressure loss, a high space utilization rate and low cell voltage. The plate-frame reactor reduces energy consumption owing to the closer electrode space, so it shows great potential to realize EO in practical wastewater treatment. The tubular ECR uses tubular electrodes in the tank. Compared with the box reactor, it has more uniform current distribution and relatively large water polarity. Moreover, the efficiency of mass transfer and organic degradation efficiency could also be improved in tubular ECR. Therefore, the tubular reactor could further expand the research and application of EO technology.

3.1. Plate-Frame ECR

The plate-frame ECR, also called filter-press ECR as shown in Figure 3, is a commonly and extensively studied ECR [37]. The system usually consists of electrodes fitted in a parallel plate assembly and frames to hold the electrodes. The electrode formation varies, including non/porous or mesh plate, etc. The fluid flow may be continuous at a single or multiple inlets and outlets [38].
Except for the above mentioned merits, the plate-frame ECR also has relatively uniform potential and current distributions, which is good for reactor scaling up. The cell architecture is flexible, and is easy to combine with other technologies. The fluid flow and mass transfer are well-defined in a rectangular flow channel [39]. The plate-frame ECR has shown good performance in EO for wastewater treatment, and part of study results are shown in Table 1. Barbosa et al. (2018) studied an electrolytic flow process associated with the wastewater treatment using a filter-press reactor. The results showed that COD removal was 90% in 90 min [40]. Qu et al. (2019) studied a simultaneous EO and in situ electro-peroxone (ORR-EO) process for the degradation of leachate concentrate in a plate-frame ECR, obtaining 80% of TOC removal (from 250 mg/L to 50 mg/L) [41]. However, the dead zone problem in plate-frame ECR is relatively serious, in which it cannot reach the ideal mixing state, and adversely affects the mass transfer of reactant in electrolyte.
At present, the researchers mainly improve it by rotating electrodes and application of turbulence promoters [33]. Quan et al. [42] designed a new plate-frame ECR for the treatment of landfill leachate. To keep the flow field uniform in the reactor, two perforated liquid distributors were mounted in the frame. The results showed that COD in the leachate could be reduced below 100 mg/L from 5000 mg/L after 60 min, reaching 98% of COD removal [42]. Nippatla and Philip [43] designed a rotating bipolar disc electrolytic plate-frame ECR for textile wastewater treatment. The turbulence was created, and mass transfer was enhanced, by electrode rotation. Thus, the COD removal was 4.8 times higher in a rotating electrode system compared to static electrodes [43]. Wachter et al. (2021) used a turbulence promoter in an electrochemical filter-press reactor to enhance the mass transfer, and the bisphenol S (BPS) degradation rate was increased approximately threefold to attain 98% of BPS removal [44].
With the continuous improvement in the design of the ECR, the application of the plate-frame reactor is still limited by the low mass transfer efficiency and high energy consumption. Moreover, the reactor optimization process makes the structure of reactor more complex and increases the construction cost. Therefore, the balance between improving efficiency and elevating cost needs to be considered to scale up plate-frame ECR in EO applications.

3.2. Tubular ECR

The tubular reactor, also known as a plug-flow reactor, was developed in the oil smelting industry in the 1940s. It belongs to a continuous operating reactor. The traditional tubular ECR is composed of an anode rod nested inside the center of the outer cathode tube, called a concentric tubular reactor as shown in Figure 4.
The tubular ECR has many advantages. Compared with a box reactor, tubular ECR electrodes have larger specific surface areas. It is easier to transfer the heat generated due to the ohmic loss in the electrolysis process to the external environment to avoid the accumulation of heat inside the reactor and provide long-term stable operation of the reactor. Tubular ECRs could reduce the reaction dead zone. The flow rate of fluid in the plug-flow reactor is faster, thus, it could decrease the volume of the reactor and improve the wastewater treatment efficiency. Additionally, the tubular reactor is convenient for combination and connection. The tubular ECR has been extensively and successfully applied in organic wastewater treatment, especially for high strength organic wastewater, such as dyeing, papermaking, food manufacturing, pharmaceutical wastewater, etc. Part of the research progress was listed in Table 2.
Körbahti et al. [48] investigated the treatment of dyeing wastewater in a tubular ECR, and obtained the respective COD, color and turbidity removals of 44.3%, 86.2% and 87.1% in 6 h. Moreover, the reactor maintains a high mass transfer coefficient of 3.62 × 10−6 m/s [48]. Ibrahim et al. (2014) used a tubular ECR with a Ti/RuO2 anode for the degradation of petroleum refinery wastewater, and complete removal of total petroleum hydrocarbon was achieved in 24 min with 85% of COD removal; they also obtained good reuse of petroleum refinery wastewater [49].
Although the performance of tubular reactors was satisfactory, it still has drawbacks. For example, the current density was not evenly distributed in a tubular reactor as the surface area of cylindrical anode and cathode are not equal, which needs optimization. The optimization of the reactor is mainly conducted from two aspects. One is to increase the contact area of the electrode with the fluid by changing the configuration of the electrode, such as mesh plate electrodes, porous electrodes and membrane electrodes. Another one is to reduce the internal dead zone of the reactor by changing the flow rate or pattern of fluid, according to the simulation results of the flow regime by computational fluid dynamics (CFD).
To deal with the problems of uneven current distribution, Li et al. [50] developed a tubular ECR with mesh plate electrodes and the solution flowed horizontally through the meshed plate electrodes. Moreover, CFD and particle image velocity (PIV) technology were used to simulate the internal flow characteristics. The results showed the mesh plate electrode could significantly promote the flow homogenization and reduce the back-reflux area in the reaction region by 50% [50]. The conclusion was coincident with Ibrahim’s study results [51]. Wang et al. [52] presented a tubular ECR based on a vertical-flow reactor with mesh plate electrodes. The results showed that the turbulent intensity increased by 200% around the electrode surface. The mass transfer coefficient of the novel tubular reactor was more than twice of that of the traditional tubular reactor [52]. The same type of reactor was reported on in other research, which demonstrated that the increase of the anode effective area was one of the main factors to enhance degradation efficiency of wastewater [53].
For improving the turbulent intensity of reactors, Guo and You [54] presented a tubular ECR based on spiral flow caused by a turbulence promoter for degradation of methylene blue in wastewater. The results showed that the mean flow rate and turbulent intensity of the new tubular reactor were increased by 500–700% and nearly 200%, respectively, compared to the traditional one [54].
At present, the research of tubular ECRs for refractory organic wastewater treatment remains in laboratory. The influencing factors on electrochemical reactions, and the process mechanism have not been fully understood, which needs further study to provide a more reliable theoretical basis for practical applications. At the same time, it is necessary to carry out large-scale reactor design. Researchers could analyze the flow conditions and the influence of chemical reactions on the flow by CFD and optimize the operating parameters accordingly.

4. Operation Mode of ECR

ECR can be operated in batch and continuous modes. The wastewater treatment in batch ECR is generally operated in a cycling pass in a fixed period of time, or until the organics are degraded enough to meet the discharging requirement. Batch treatment is normally suitable for small-scale or intermittent supplies of wastewater, and has the merit of a high impetus for the degradation kinetic rate, as the initial concentration of wastewater is high. However, batch-type reactors have the problem of limited treatment capacity. A continuous ECR is generally referred to as a “plug flow” reactor that operates in a single pass mode, with the reactant (organics) in influent contact with the electrodes in parallel or in series. From the engineering point of view, continuous mode is more preferred, as continuous ECRs provide major advantages such as faster reactions, cleaner products and safer and easier to scaling up, which outcompetes the limitations of batch reactors for large-scale wastewater treatment.

5. The Flow Mode and Electrode Configuration of ECRs

The ECR can be operated in flow-through and flow-by mode, according to the flow mode and the electrode configuration, as shown in Figure 5.

5.1. Flow-By Mode

The flow-by mode of the ECR refers to when the flow direction of fluid is parallel to the electrodes, so that the aqueous solution cannot cross the electrode. Therefore, the electrodes morphology selected are mostly plate electrodes. As known, chemical and mass transfer processes could limit contaminant degradation rates. Due to the high reactivity of OH, the reaction with pollutants occurs in a narrow zone (<1 μm) on the anode surface, but the diffusion boundary layer of the anode surface is over 100 μm in the flow-by mode [59]. Thus, the limited mass transfer hindered by the boundary layer results in low treatment efficiency and high energy consumption.
Changing the fluid flow rate could improve the mass transfer. The flow rate of the liquor in the reactor determines whether the flow pattern is laminar or turbulent. This distinction determines the residence time of reactants in the reactor and on the surface of the electrodes. Rafael Machado Reis et al. (2013) reported on the degradation of dipyrone in a flow-by ECR with a laminar (50 L h−1) or turbulent (300 L h−1) regime, and a 92.5% TOC removal rate was achieved at an electrolyte flow rate of 300 L h−1 [60]. However, recently, more research has resorted to high active surface area electrodes, which are flow-through mode electrodes, to improve mass transfer.

5.2. Flow-Through Mode

Flow-through mode refers to when the fluid direction is perpendicular to the electrode. Thus, flow-through ECR requires porous or mesh electrodes and the solution passes through pores 0.1–1.0 mm wide [61]. The flow-through mode could enhance the mass transfer of reactant by suppressing the boundary layer, thereby improving the degradation efficiency of different pollutants [59]. It is testified that the reaction rate constants of flow-through and flow-by modes were 0.0238 min−1 and 0.0115 min−1 according to the experimental data. The result means that flow-through mode ECR could drastically reduce the thickness of the diffusion boundary layer [62]. Moreover, the porous structure could improve the current efficiency by providing more reaction sites and reducing applied current. Therefore, it is beneficial to improve the utilization efficiency of electrons and increase the electrode lifetime [63].
Recently, the reported flow-through electrodes include porous or mesh anodes made of CNTs, Ti4O7, Ti/PbO2, Ti/RuO2 and Ti/SnO2-Sb. The pore size of the electrode is one of the key parameters for electrochemical reactions in an EO system. Chen et al. (2020) studied the pseudo-first-order rate constants (k) of MP-Ti ENTA/SnO2-Sb2O3, which were estimated to be 0.38, 1.63 and 1.24 min−1 with the nominal pore sizes of 10, 20 and 50 μm, respectively, in a flow-through ECR system [64].
To obtain an efficient and low-energy ECR for wastewater treatment, Pérez et al. [65] developed a microfluid flow-through ECR with a narrow internal electrode space and flow-through electrodes. Compared with a commercial flow-by ECR, the microfluid flow-through ECR required between 4- and 10-fold less current and from 6 to 15 times less energy consumption [65]. Yang et al. [63] constructed an electro-peroxone membrane filtration (EPMF) system based on OH production, and enhanced mass transfer with a CNT membrane cathode operated in flow-through mode. The results showed that the kinetic rate constant (k) was 1735.30 × 10–3 s−1 for ibuprofen (IBU) removal. In contrast, the value of k was only 16.4 × 10–3 s−1 in traditional batch reactors [63]. Zeng et al. [66] studied the efficient electro-oxidation of sulfamethoxazole (SMX) by a TiO2 nanotube arrays-based electrocatalytic flow-through membrane, and the SMX removal efficiency of the flow-through mode was increased by 71%. It is verified, the superiority of a porous flow-through anode [66]. Zhang et al. [67] developed a flow-through ECR with a Ti4O7-based membrane electrode. It proved that this flow-through mode could enhance the mass transfer with a high tetracycline degradation efficiency (97.24%) and low energy consumption (0.18 kWh/g DOC) [67]. Maldonado et al. (2021) developed a boron-doped diamond (BDD) anode flow-through ECR for the electro-oxidation of perfluoroalkyl acids (PFAAs) in landfill leachates, and obtained a 73.5% of the total PFAAs removal after a 2 h reaction [68]. Ormeno-Cano and Radjenovic (2022) studied electrochemical degradation of antibiotics using flow-through graphene sponge electrodes. The results showed the removal of antibiotics was above 80% observed [69]. Chen et al. [62] developed a high-efficiency EO system in flow-through mode based on macroporous enhanced TiO2 nanotube array/SnO2-Sb (MP-Ti-ENTA/SnO2-Sb) anode for treatment of reverse osmosis concentrates. The COD removal efficiency of the flow-through ECR was higher than 70% after 60 min, while the flow-by ECR required 120 min and used more energy [62].
In summary, compared with flow-by ECRs, flow-through ECRs show better performance in the electro-oxidation of organics and exhibit broader application prospects for wastewater treatment.

6. ECR Classification according to the Electrode Configuration

The ECR electrode configuration is divided into two-dimensional electrode (2D) and three-dimensional electrode (3D) configuration, according to the presence or the absence of particle electrodes in the reactor. A 3D-ECR consists of granular electrode materials packed in the middle of the traditional 2D flat plate electrode, and operated under a flow-by or flow-through mode. The particle electrodes can be polarized by an external electric field to form a large amount of bipolar micro electrodes [70]. The electrochemical reaction can occur on the surface of each particle. Thus, these particle electrodes extend surface area and enhance the mass transfer of the reactants in 3D-ECR [71].
The mechanism of organic degradation in 3D-ECR depends on the type of organic matter, electrode materials and particles as shown in Figure 6. The anodic and cathodic reactions are consistent with 2D-ECRs and the main redox reaction still occurs on the surface of the electrodes. The particle electrodes, including granular activated carbon (GAC), metal particles, carbon aerogel (CA) and modified kaolin, have good electrical conductivity and large surface areas. In the reaction, the particle electrodes are polarized to generate OH to degrade pollutants. Moreover, the particle electrodes would also remove some pollutants through adsorption [72]. It could improve the degradation efficiency by improving the concentration of pollutants at the particle electrode interface [73]. It is proven that the degradation rate of pollutants is mainly related to the adsorption performance of the particle electrode in a low voltage system [74]. Therefore, the degradation rate of organic matter in the whole reactor is greatly improved.
The configurations of 3D-ECR are varied according to whether the particles are filled in fixed or fluidized mode. The fixed bed ECR and fluidized bed ECR were as shown in Figure 7.

6.1. Fixed Bed 3D-ECR

The fixed bed electrode has high spatiotemporal efficiency, area–volume ratio and current efficiency as well as easy controlling. The particles electrode in a fixed bed 3D-ECR hardly gain evenly distributed current and potential. To promote the mass transfer, magnetic stirring or air flow stirring is required. However, it is easy to form a short-circuit current and cake the particle electrode due to fixed bed mode, and thus, directly affect the current and energy efficiency of ECRs. An approach for reducing the short circuit problem in fixed bed ECRs is using particle electrodes with different densities, which are distributed evenly in the reactor [75].

6.2. Fluidized Bed 3D-ECR

The fluidized bed 3D-ECR provides a larger electrode activation area and high mass transfer rate, because the strong disturbance of the particles effectively reduces the concentration polarization of the electrolyte on the particle surface. However, uneven distribution of the current, and potential generated because the particle electrodes are not in close contact with one another can lead to reduced treatment efficiency. Kitaev and Rubanovich [76] developed that a reactor equipped with a mixer, consisting of an electric rotating shaft with a plurality of turning vanes arranged along the longitudinal axis, which has the advantages of saving energy. The power consumption per cubic meter of wastewater was approximately 10~12 kWh with organic treatment efficiency of 95–98% [76].
The 3D-ECR system has been widely used in the treatment of various wastewater, such as phenolic compounds, coking, oils, etc., as summarized in Table 3. Ni et al. [77] compared the TOC and COD degradation in Rhodamine B wastewater using a cycling-flow 2D-ECR and a 3D-ECR system. The results showed that the removal efficiencies of TOC and COD in a 3D-ECR were both approximately twice as those in 2D-ECR [77]. Wei et al. [78] investigated a 3D-ECR with GAC and a porous ceramsite particle as the electrode for pretreatment of heavy oil refinery wastewater. Results showed that the removal efficiencies of COD, TOC and toxicity units were 45.5%, 43.3% and 67.2%, respectively, and the ratio of BOD/COD was increased from 0.10 to 0.29, which is beneficial for further biological treatment [78].
The conventional particle electrodes, such as GAC, showed a modest degradation ability for pollutants and relatively high chemical stability during the EO process for wastewater treatment. However, these particle electrodes do not have enough catalytic activity to completely degrade the contaminants adsorbed on their surface, and the pores of the particle electrodes are easily clogged by contaminants during the reaction, so it will influence long term operation performance of wastewater treatment [79].
In recent years, it became a hot spot to enhance the EO performance of 3D-ECRs focusing on the surface modification of a particle electrode by loading catalyst. Li et al. [80] prepared GAC doped with TiO2-SiO2 oxide (TiO2-SiO2/GAC) as particle electrodes in a 3D-ECR for dyeing wastewater treatment. The results showed the TiO2-SiO2/GAC particles had a lower specific surface area and pore volume than the original GAC; this means that TiO2-SiO2 covered part of the outer surface and the internal pores of GAC. It could provide more probability in the degradation of the pollutants adsorbed into the pores [80]. Chen et al. [81] studied a 3D-ECR using the Sn/Sb-Mn-GAC particle electrodes to effectively treat 4-chlorophenol in wastewater. The oxygen evolution potential (OEP) of the Sn/Sb-Mn-GAC particle electrode was much higher than that of GAC, which can produce more oxidizing radicals and a higher removal rate; a rate of 96% in 60 min was obtained [81].
So far, refractory organic wastewater treatment with EO has achieved good results in the laboratory, but the cases of large-scale EO treatment are still lacking. The reasons are as follows: (1) The kinetic and thermodynamic processes of adsorption, desorption and electrochemical oxidation that occurred on the 3D electrode surface are poorly defined. (2) The electrocatalysis mechanism of organics in 3D-ECR is more complex than in 2D-ECR, and the wastewater treatment efficiency is influenced by many parameters, such as the cell structure, bed expansion rate, voltage, current density, reaction time, electrode material and reactant concentration, etc.
Increasing the voltage or current density to a certain extent is beneficial to pollutant degradation, but it also increases energy consumption. It is necessary to balance the processing effect and energy consumption [87]. Furthermore, increasing the amount of particle electrodes would result in more particle electrodes, but highly packed particle electrodes would occupy much space in the 3D-ECR, leading to reduce the amount of wastewater treatment [88].
The 3D-ECR system deserves comprehensive development due to its unique performance. In addition to the research on electrode materials for refractory organic wastewater treatment [89], the main research trend in the direction of ECR lies in optimizing and expanding the design of 3D-ECR. During the continuous operating process, the system may lose the ability to degrade the pollutants due to the accumulation of the pollutants in the reaction system. Optimizing the operating parameters and maintaining the stable operation of the 3D-ECR system has great commercial value, which is greatly significant for future development.

7. Conclusions and Future Perspectives

The design and optimization of ECR configuration plays a crucial role in the development of EO technology. The designs of ECRs are diverse and depend on different objectives, as well as the shape and characteristics of the electrodes in wastewater treatment. In this review, the key points were discussed as follow:
(1)
Compared with the plate-frame ECR, the tubular ECR has less dead zone and a higher mass transfer efficiency due to the larger specific surface areas of electrodes. The mass transfer efficiency is the one of most important factors for degradation efficiency of wastewater in ECRs. The optimization reactors enhance degradation efficiency of wastewater mainly by changing the shape of electrodes or increasing fluid flow rate.
(2)
The limited mass transfer hindered by boundary layer in flow-by ECRs. Additionally, flow-through ECRs could enhance the mass transfer of reactant by suppressing the boundary layer.
(3)
From the engineering point of view, continuous mode operation is more preferred because of faster reactions, cleaner products and safer and easier scaling up than batch mode for large-scale wastewater treatment in EO system.
(4)
Particle electrodes in the 3D-ECR could improve the degradation efficiency by improving the concentration of pollutants at the particle electrode interface. The 3D-ECR system deserves comprehensive development due to the particle electrodes. However, the mechanism of reactions is more complicated in a 3D-ECR. Optimizing the operating parameters and enhancing electrode materials performance of adsorption and catalysis to maintain the stable operation are significant for future development of 3D-ECRs.
At present, CFD has become an important model tool for reactor design. It could predict fluid flow, heat and mass transfer, mixing effect and electrochemical reaction mechanism in ECRs by different models. Combining simulation and practice can evaluate the flow pattern of fluid in the reactor more accurately, get more scientific and effective electrochemical performance and provide a theoretical basis for further optimization and improvement for the realization of refractory organic wastewater removal. Additionally, the removal efficiency of pollutants and energy consumption for high concentrations of refractory organic wastewater limit to the application of single EO technology. Therefore, it can also be combined with other technologies, such as membrane filtration, photochemical [90] and electrochemical technology (electroflotation [91], electrocoagulation, electrodialysis, etc.) to achieve more efficient and energy-saving removal of refractory organic wastewater. For example, the effectiveness of combined electrocoagulation (EC) with electro-oxidation (EO) for refractory organic wastewater was proven. EC as a pre-treatment could degrade various pollutants such as organic matter, suspended solids, colloidal, etc. After a certain degree of treatment, the organic matter of wastewater could be removed in an EO reactor. An amount of works showed that the EC-EO technology could not only improve the removal rate of organic matter, but also greatly reduce the energy consumption. Therefore, it may be a popular field to design and optimize combined electrochemical reactors.

Author Contributions

J.L.: Writing- Original draft preparation; N.R.: Literature review; C.Q.: Draw designs; S.L.: Files Preparation; Y.X.: Conceptualization; D.L.: Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by grants from the National Natural Science Foundation of China [No. 52070008].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Luan, M.; Jing, G.; Piao, Y.; Liu, D.; Jin, L. Treatment of refractory organic pollutants in industrial wastewater by wet air oxidation. Arab. J. Chem. 2017, 10, S769–S776. [Google Scholar] [CrossRef] [Green Version]
  2. Deng, M.; Wu, K.; Yang, T.; Yu, D.; Liu, G.; Gong, S.; Sun, D.; Petru, M. Construction of Novel Electro-Fenton Systems by Magnetically Decorating Zero-Valent Iron onto RuO2-IrO2/Ti Electrode for Highly Efficient Pharmaceutical Wastewater Treatment. Water 2022, 14, 1044. [Google Scholar] [CrossRef]
  3. Zwane, B.N.; Orimolade, B.O.; Koiki, B.A.; Mabuba, N.; Gomri, C.; Petit, E.; Bonniol, V.; Lesage, G.; Rivallin, M.; Cretin, M.; et al. Combined Electro-Fenton and Anodic Oxidation Processes at a Sub-Stoichiometric Titanium Oxide (Ti4O7) Ceramic Electrode for the Degradation of Tetracycline in Water. Water 2021, 13, 2772. [Google Scholar] [CrossRef]
  4. Wu, D.; Li, Y.; Lu, G.; Lin, Q.; Wei, L.; Zhang, P. Removal of Aqueous Para-Aminobenzoic Acid Using a Compartmental Electro-Peroxone Process. Water 2021, 13, 2961. [Google Scholar] [CrossRef]
  5. Zangeneh, H.; Zinatizadeh, A.A.L.; Habibi, M.; Akia, M.; Hasnain Isa, M. Photocatalytic oxidation of organic dyes and pollutants in wastewater using different modified titanium dioxides: A comparative review. J. Ind. Eng. Chem. 2015, 26, 1–36. [Google Scholar] [CrossRef]
  6. Lin, R.; Li, Y.; Yong, T.; Cao, W.; Wu, J.; Shen, Y. Synergistic effects of oxidation, coagulation and adsorption in the integrated fenton-based process for wastewater treatment: A review. J. Environ. Manag. 2022, 306, 114460. [Google Scholar] [CrossRef]
  7. Anil, G.; Scaria, J.; Nidheesh, P.V. Removal of Synthetic Dye from Aqueous Solution Using MnFe2O4-GO Catalyzed Heterogeneous Electro-Fenton Process. Water 2022, 14, 3350. [Google Scholar] [CrossRef]
  8. Du, C.; Zhang, Y.; Zhang, Z.; Zhou, L.; Yu, G.; Wen, X.; Chi, T.; Wang, G.; Su, Y.; Deng, F.; et al. Fe-based metal organic frameworks (Fe-MOFs) for organic pollutants removal via photo-Fenton: A review. Chem. Eng. J. 2022, 431, 133932. [Google Scholar] [CrossRef]
  9. Tian, K.; Hu, L.; Li, L.; Zheng, Q.; Xin, Y.; Zhang, G. Recent advances in persulfate-based advanced oxidation processes for organic wastewater treatment. Chinese Chem. Lett. 2022, 33, 4461–4477. [Google Scholar] [CrossRef]
  10. Soomro, G.S.; Qu, C.; Ren, N.; Meng, S.; Li, X.; Liang, D.; Zhang, S.; Li, Y. Efficient removal of refractory organics in landfill leachate concentrates by electrocoagulation in tandem with simultaneous electro-oxidation and in-situ peroxone. Environ. Res. 2020, 183, 109249. [Google Scholar] [CrossRef]
  11. Lauzurique, Y.; Espinoza, L.C.; Huiliñir, C.; García, V.; Salazar, R. Anodic Oxidation of Industrial Winery Wastewater Using Different Anodes. Water 2022, 14, 95. [Google Scholar] [CrossRef]
  12. Qu, C.; Li, Y.-g.; Meng, S.-j.; Li, X.-h.; Zhang, S.-j.; Liang, D.-w. Enhanced refractory organics removal by OH and 1O2 generated in an electro-oxidation system with cathodic Fenton-like reaction. J. Hazard. Mater. 2022, 434, 128923. [Google Scholar] [CrossRef] [PubMed]
  13. Qu, C.; Soomro, G.S.; Ren, N.; Liang, D.-w.; Lu, S.-f.; Xiang, Y.; Zhang, S.-j. Enhanced electro-oxidation/peroxone (in situ) process with a Ti-based nickel-antimony doped tin oxide anode for phenol degradation. J. Hazard. Mater. 2020, 384, 121398. [Google Scholar] [CrossRef] [PubMed]
  14. Pulkka, S.; Martikainen, M.; Bhatnagar, A.; Sillanpää, M. Electrochemical methods for the removal of anionic contaminants from water—A review. Sepa. Purif. Technol. 2014, 132, 252–271. [Google Scholar] [CrossRef]
  15. Łuba, M.; Mikołajczyk, T.; Pierożyński, B.; Smoczyński, L.; Wojtacha, P.; Kuczyński, M. Electrochemical Degradation of Industrial Dyes in Wastewater through the Dissolution of Aluminum Sacrificial Anode of Cu/Al Macro-Corrosion Galvanic Cell. Molecules 2020, 25, 4108. [Google Scholar] [CrossRef] [PubMed]
  16. Piotrowska, G.; Pierozynski, B. Electrooxidation of phenol on carbon fibre-based anodes through continuous electrolysis of synthetic wastewater. Pol. J. Chem. Technol. 2018, 20, 96–102. [Google Scholar] [CrossRef] [Green Version]
  17. Qu, C.; Ren, N.; Zhang, S.-J.; Li, Y.-G.; Meng, S.-J.; Li, X.-H.; Wang, S.-Q.; Liang, D.-W.; Li, A.-R. Degradation of triclosan by anodic oxidation/in-situ peroxone process: Kinetics, pathway and reaction mechanism. Chemosphere 2021, 272, 129453. [Google Scholar] [CrossRef]
  18. Zhou, B.; Yu, Z.; Wei, Q.; Long, H.; Xie, Y.; Wang, Y. Electrochemical oxidation of biological pretreated and membrane separated landfill leachate concentrates on boron doped diamond anode. Appl. Surf. Sci. 2016, 377, 406–415. [Google Scholar] [CrossRef]
  19. Subba Rao, A.N.; Venkatarangaiah, V.T. Metal oxide-coated anodes in wastewater treatment. Environ. Sci. Pollut. R. 2014, 21, 3197–3217. [Google Scholar] [CrossRef]
  20. Coster, J.D.; Vanherck, W.; Appels, L.; Dewil, R. Selective electrochemical degradation of 4-chlorophenol at a Ti/RuO2-IrO2 anode in chloride rich wastewater. J. Environ. Manag. 2017, 190, 61–71. [Google Scholar] [CrossRef]
  21. Ornelas Dávila, O.; Lacalle Bergeron, L.; Dávila Jiménez, M.M.; Sirés, I.; Brillas, E.; Navarro, A.F.R.; Arandes, J.B.; Sancho Llopis, J.V. Study of the electrochemical oxidation of 4,6-dimethyldibenzothiophene on a BDD electrode employing different techniques. J. Electroanal. Chem. 2021, 894, 115364. [Google Scholar] [CrossRef]
  22. Alfonso-Muniozguren, P.; Cotillas, S.; Boaventura, R.A.R.; Moreira, F.C.; Lee, J.; Vilar, V.J.P. Single and combined electrochemical oxidation driven processes for the treatment of slaughterhouse wastewater. J. Clean. Prod. 2020, 270, 121858. [Google Scholar] [CrossRef]
  23. Lei, Y.; Zhan, Z.; Saakes, M.; van der Weijden, R.D.; Buisman, C.J.N. Electrochemical recovery of phosphorus from wastewater using tubular stainless-steel cathode for a scalable long-term operation. Water Res. 2021, 199, 117199. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, W.; Ai, Z.; Zhang, L. Design of a neutral three-dimensional electro-Fenton system with foam nickel as particle electrodes for wastewater treatment. J. Hazard. Mater. 2012, 243, 257–264. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, K.; Dai, Q.; Hu, C.; Tong, Y.; Wang, Y.; Song, S.; Dai, L. Earth-abundant metal-free carbon-based electrocatalysts for Zn-air batteries to power electrochemical generation of H2O2 for in-situ wastewater treatment. Chem. Eng. J. 2021, 416, 128338. [Google Scholar] [CrossRef]
  26. Qu, C.; Liang, D.-w. Novel electrochemical advanced oxidation processes with H2O2 generation cathode for water treatment: A review. J. Environ. Chem. Eng. 2022, 10, 107896. [Google Scholar] [CrossRef]
  27. Zhang, G.; Yang, F.; Gao, M.; Fang, X.; Liu, L. Electro-Fenton degradation of azo dye using polypyrrole/anthraquinonedisulphonate composite film modified graphite cathode in acidic aqueous solutions. Electrochimi. Acta 2008, 53, 5155–5161. [Google Scholar] [CrossRef]
  28. Zhou, L.; Zhou, M.; Zhang, C.; Jiang, Y.; Bi, Z.; Yang, J. Electro-Fenton degradation of p-nitrophenol using the anodized graphite felts. Chem. Eng. J. 2013, 233, 185–192. [Google Scholar] [CrossRef]
  29. Khataee, A.R.; Safarpour, M.; Zarei, M.; Aber, S. Electrochemical generation of H2O2 using immobilized carbon nanotubes on graphite electrode fed with air: Investigation of operational parameters. J. Electroanal. Chem. 2011, 659, 63–68. [Google Scholar] [CrossRef]
  30. Hou, J.; Xu, Z.; Ji, J.; Zhao, Y.; Gao, M.; Jin, C. Enhanced in-situ electro-generation of H2O2 using PTFE and NH4HCO3 modified C/PTFE electrode for treatment of landfill leachate. J. Environ. Manage. 2021, 295, 112933. [Google Scholar] [CrossRef]
  31. Garcia-Rodriguez, O.; Lee, Y.Y.; Olvera-Vargas, H.; Deng, F.; Wang, Z.; Lefebvre, O. Mineralization of electronic wastewater by electro-Fenton with an enhanced graphene-based gas diffusion cathode. Electrochim. Acta 2018, 276, 12–20. [Google Scholar] [CrossRef]
  32. Geraldino, H.C.L.; Freitas, T.K.F.S.; Manholer, D.D.; França, F.; Oliveira, J.H.; Volnistem, E.A.; Lima, A.R.F.; Bertotti, M.; Girotto, E.M.; Garcia, J.C. Electrochemical generation of H2O2 using gas diffusion electrode improved with rGO intensified with the Fe3O4/GO catalyst for degradation of textile wastewater. J. Water Process Eng. 2020, 36, 101377. [Google Scholar] [CrossRef]
  33. Perry, S.C.; Ponce de León, C.; Walsh, F.C. Review—The Design, Performance and Continuing Development of Electrochemical Reactors for Clean Electrosynthesis. J. Electrochem. Soc. 2020, 167, 155525. [Google Scholar] [CrossRef]
  34. Szpyrkowicz, L.; Kaul, S.N.; Neti, R.N. Tannery wastewater treatment by electro-oxidation coupled with a biological process. J. Appl. Electrochem. 2005, 35, 381–390. [Google Scholar] [CrossRef]
  35. Monteil, H.; Pechaud, Y.; Oturan, N.; Trellu, C.; Oturan, M.A. Pilot scale continuous reactor for water treatment by electrochemical advanced oxidation processes: Development of a new hydrodynamic/reactive combined model. Chem. Eng. J. 2021, 404, 127048. [Google Scholar] [CrossRef]
  36. Regalado-Méndez, A.; Mentado-Morales, J.; Vázquez, C.E.; Martínez-Villa, G.; Cordero, M.E.; Zárate, L.G.; Skogestad, S.; Peralta-Reyes, E. Modeling and Hydraulic Characterization of a Filter-Press-Type Electrochemical Reactor by Using Residence Time Distribution Analysis and Hydraulic Indices. Int. J. Chem. React. Eng. 2018, 16, 1–15. [Google Scholar] [CrossRef]
  37. Frías-Ferrer, Á.; Tudela, I.; Louisnard, O.; Sáez, V.; Esclapez, M.D.; Díez-García, M.I.; Bonete, P.; González-García, J. Optimized design of an electrochemical filter-press reactor using CFD methods. Chem. Eng. J. 2011, 169, 270–281. [Google Scholar] [CrossRef]
  38. Arenas, L.F.; Ponce de León, C.; Walsh, F.C. Critical Review—The Versatile Plane Parallel Electrode Geometry: An Illustrated Review. J. Electrochemi. Soc. 2020, 167, 023504. [Google Scholar] [CrossRef]
  39. Pinto, C.; Fernandes, A.; Lopes, A.; Nunes, M.J.; Baía, A.; Ciríaco, L.; Pacheco, M.J. Reuse of Textile Dyeing Wastewater Treated by Electrooxidation. Water 2022, 14, 1084. [Google Scholar] [CrossRef]
  40. Barbosa, A.D.; da Silva, L.F.; de Paula, H.M.; Romualdo, L.L.; Sadoyama, G.; Andrade, L.S. Combined use of coagulation (M. oleifera) and electrochemical techniques in the treatment of industrial paint wastewater for reuse and/or disposal. Water Res. 2018, 145, 153–161. [Google Scholar] [CrossRef]
  41. Qu, C.; Lu, S.; Liang, D.; Chen, S.; Xiang, Y.; Zhang, S. Simultaneous electro-oxidation and in situ electro-peroxone process for the degradation of refractory organics in wastewater. J. Hazard. Mater. 2019, 364, 468–474. [Google Scholar] [CrossRef] [PubMed]
  42. Quan, X.; Cheng, Z.; Chen, B.; Zhu, X. Electrochemical oxidation of recalcitrant organic compounds in biologically treated municipal solid waste leachate in a flow reactor. J. Environ. Sci. 2013, 25, 2023–2030. [Google Scholar] [CrossRef]
  43. Nippatla, N.; Philip, L. Performance evaluation of a novel electrolytic reactor with rotating and non rotating bipolar disc electrodes for synthetic textile wastewater treatment. J. Environ. Chem. Eng. 2020, 8, 103462. [Google Scholar] [CrossRef]
  44. Wachter, N.; Bocchi, N.; Rocha-Filho, R.C. Use of a turbulence promoter in an electrochemical filter-press reactor: Consolidated evidence of significant enhancement of organics mass transport and degradation rates. Sep. Purif. Technol. 2021, 276, 119292. [Google Scholar] [CrossRef]
  45. Palma-Goyes, R.E.; Vazquez-Arenas, J.; Torres-Palma, R.A.; Ostos, C.; Ferraro, F.; González, I. The abatement of indigo carmine using active chlorine electrogenerated on ternary Sb2O5-doped Ti/RuO2-ZrO2 anodes in a filter-press FM01-LC reactor. Electrochim. Acta 2015, 174, 735–744. [Google Scholar] [CrossRef]
  46. Brocenschi, R.F.; Rocha-Filho, R.C.; Bocchi, N.; Biaggio, S.R. Electrochemical degradation of estrone using a boron-doped diamond anode in a filter-press reactor. Electrochim. Acta 2016, 197, 186–193. [Google Scholar] [CrossRef]
  47. Dória, A.R.; Pupo, M.; Santos, G.d.O.S.; Vilar, D.d.S.; Torres, N.H.; Romanholo Ferreira, L.F.; Cavalcanti, E.B.; Eguiluz, K.I.B.; Salazar-Banda, G.R. Electrochemical oxidation of indanthrene blue dye in a filter-press flow reactor and toxicity analyses with Raphidocelis subcapitata and Lactuca sativa. Ecotox. Environ. Safe. 2020, 198, 110659. [Google Scholar] [CrossRef]
  48. Körbahti, B.K.; Tanyolaç, A. Electrochemical treatment of simulated industrial paint wastewater in a continuous tubular reactor. Chem. Eng. J. 2009, 148, 444–451. [Google Scholar] [CrossRef]
  49. Skban Ibrahim, D.; Seethala Devi, P.; Veerababhu, C.; Balasubramanian, N. Treatment of Petroleum Effluent Using a Tubular Electrochemical Reactor. Petrol. Sci. Technol. 2014, 32, 1932–1939. [Google Scholar] [CrossRef]
  50. Li, T.; Li, X.; Chen, F.; Wang, J. Three-Dimensional CFD Simulation of the Tubular Electrochemical Reactor with Meshed Plate Electrodes. J. Electrochem. Soc. 2014, 161, E81–E86. [Google Scholar] [CrossRef]
  51. Ibrahim, D.S.; Veerabahu, C.; Palani, R.; Devi, S.; Balasubramanian, N. Flow dynamics and mass transfer studies in a tubular electrochemical reactor with a mesh electrode. Comput. Fluids 2013, 73, 97–103. [Google Scholar] [CrossRef]
  52. Wang, J.; Li, T.; Zhou, M.; Li, X.; Yu, J. Characterization of hydrodynamics and mass transfer in two types of tubular electrochemical reactors. Electrochim. Acta 2015, 173, 698–704. [Google Scholar] [CrossRef]
  53. Fan, T.-X.; Cai, Y.; Chu, G.-W.; Luo, Y.; Zhang, L.-L.; Chen, J.-F. A Novel Rotating Multielectrodes Reactor for Electrochemical Oxidation Process Intensification. Ind. Eng. Chem. Res. 2019, 58, 2396–2404. [Google Scholar] [CrossRef]
  54. Guo, X.; You, S. Characterization of hydrodynamics and electrochemical treatment of dye wastewater in two types of tubular electrochemical reactors. IOP Conf. Ser. Earth Environ. Sci. 2017, 81, 012008. [Google Scholar] [CrossRef] [Green Version]
  55. Mythilishri, R.; Kamalakannan, V.P.; Saravanathamizhan, R.; Balasubramanian, N. Kinetic and residence time distribution modeling of tubular electrochemical reactor: Analysis of results using Taguchi method. Water Pract. Technol. 2020, 16, 108–116. [Google Scholar] [CrossRef]
  56. Vijayakumar, V.; Saravanathamizhan, R.; Balasubramanian, N. Electro oxidation of dye effluent in a tubular electrochemical reactor using TiO2/RuO2 anode. J. Water Process Eng. 2016, 9, 155–160. [Google Scholar] [CrossRef]
  57. Li, D.; Tang, J.; Zhou, X.; Li, J.; Sun, X.; Shen, J.; Wang, L.; Han, W. Electrochemical degradation of pyridine by Ti/SnO2-Sb tubular porous electrode. Chemosphere 2016, 149, 49–56. [Google Scholar] [CrossRef]
  58. Xu, A.; Han, W.; Li, J.; Sun, X.; Shen, J.; Wang, L. Electrogeneration of hydrogen peroxide using Ti/IrO2-Ta2O5 anode in dual tubular membranes Electro-Fenton reactor for the degradation of tricyclazole without aeration. Chem. Eng. J. 2016, 295, 152–159. [Google Scholar] [CrossRef]
  59. Chaplin, B.P. Critical review of electrochemical advanced oxidation processes for water treatment applications. Environ. Sci.-Proc. Imp. 2014, 16, 1182–1203. [Google Scholar] [CrossRef]
  60. Reis, R.M.; Baio, J.A.F.; Migliorini, F.L.; Rocha, R.d.S.; Baldan, M.R.; Ferreira, N.G.; Lanza, M.R.d.V. Degradation of dipyrone in an electrochemical flow-by reactor using anodes of boron-doped diamond (BDD) supported on titanium. J. Electroanal. Chem. 2013, 690, 89–95. [Google Scholar] [CrossRef]
  61. Bard, A.J.; Faulkner, L.R. Student Solutions Manual to Accompany Electrochemical Methods: Fundamentals and Applications, 2nd ed; Wiley: New York, NY, USA, 2002. [Google Scholar]
  62. Chen, M.; Zhao, X.; Wang, C.; Pan, S.; Zhang, C.; Wang, Y. Electrochemical oxidation of reverse osmosis concentrates using macroporous Ti-ENTA/SnO2-Sb flow-through anode: Degradation performance, energy efficiency and toxicity assessment. J. Hazard. Mater. 2021, 401, 123295. [Google Scholar] [CrossRef]
  63. Yang, Q.; Li, M.; Wang, J.; Huang, H. Synergistic electro-catalytic oxidation of ibuprofen in electro-peroxone system with flow-through carbon nanotube membrane cathode. Chem. Eng. J. 2022, 435, 135180. [Google Scholar] [CrossRef]
  64. Chen, M.; Wang, C.; Zhao, X.; Wang, Y.; Zhang, W.; Chen, Z.; Meng, X.; Luo, J.; Crittenden, J. Development of a highly efficient electrochemical flow-through anode based on inner in-site enhanced TiO2-nanotubes array. Environ. Int. 2020, 140, 105813. [Google Scholar] [CrossRef] [PubMed]
  65. Pérez, J.F.; Llanos, J.; Sáez, C.; López, C.; Cañizares, P.; Rodrigo, M.A. Development of an innovative approach for low-impact wastewater treatment: A microfluidic flow-through electrochemical reactor. Chem. Eng. J. 2018, 351, 766–772. [Google Scholar] [CrossRef]
  66. Zeng, W.; Liang, H.; Zhang, H.; Luo, X.; Lin, D.; Li, G. Efficient electrochemical oxidation of sulfamethoxazole by a novel reduced TiO2 nanotube arrays-based flow-through electrocatalytic membrane. Sep. Purif. Technol. 2022, 289, 120720. [Google Scholar] [CrossRef]
  67. Zhang, Y.; Ding, J.; Gao, Q.; Jiang, B.; Li, C.; Zhao, Q. Synthesis of low-cost Ti4O7 membrane electrode for electrooxidation of tetracycline under flow-through conditions: Performance, kinetics and mechanism. Process Saf. Environ. 2022, 159, 931–943. [Google Scholar] [CrossRef]
  68. Maldonado, V.Y.; Landis, G.M.; Ensch, M.; Becker, M.F.; Witt, S.E.; Rusinek, C.A. A flow-through cell for the electrochemical oxidation of perfluoroalkyl substances in landfill leachates. J. Water Process Eng. 2021, 43, 102210. [Google Scholar] [CrossRef]
  69. Ormeno-Cano, N.; Radjenovic, J. Electrochemical degradation of antibiotics using flow-through graphene sponge electrodes. J. Hazard. Mater. 2022, 431, 128462. [Google Scholar] [CrossRef]
  70. Kothari, M.S.; Aly Hassan, A.; Shah, K.A. Three-Dimensional Electrochemical Oxidation of Recalcitrant Dye Using Green Iron Microparticles. Water 2021, 13, 1925. [Google Scholar] [CrossRef]
  71. Ghanbarlou, H.; Pedersen, N.L.; Simonsen, M.E.; Muff, J. Nitrogen-Doped Graphene Iron-Based Particle Electrode Outperforms Activated Carbon in Three-Dimensional Electrochemical Water Treatment Systems. Water 2020, 12, 3121. [Google Scholar] [CrossRef]
  72. Yu, D.; Pei, Y.; Ji, Z.; He, X.; Yao, Z. A review on the landfill leachate treatment technologies and application prospects of three-dimensional electrode technology. Chemosphere 2022, 291, 132895. [Google Scholar] [CrossRef] [PubMed]
  73. Ma, J.; Gao, M.; Shi, H.; Ni, J.; Xu, Y.; Wang, Q. Progress in research and development of particle electrodes for three-dimensional electrochemical treatment of wastewater: A review. Environ. Sci. Pollut. Res. 2021, 28, 47800–47824. [Google Scholar] [CrossRef]
  74. Yu, D.; Cui, J.; Li, X.; Zhang, H.; Pei, Y. Electrochemical treatment of organic pollutants in landfill leachate using a three-dimensional electrode system. Chemosphere 2020, 243, 125438. [Google Scholar] [CrossRef] [PubMed]
  75. Li, H.; Yang, H.; Cheng, J.; Hu, C.; Yang, Z.; Wu, C. Three-dimensional particle electrode system treatment of organic wastewater: A general review based on patents. J. Clean. Prod. 2021, 308, 127324. [Google Scholar] [CrossRef]
  76. Kitaev, A.; Rubanovich, M. Apparatus and Method for Treatment of Wastewater. U.S. Patent 12/958,098 WO2009147666, 10 December 2009. [Google Scholar]
  77. Ni, J.; Shi, H.; Xu, Y.; Wang, Q. A Comparison of the Mechanism of TOC and COD Degradation in Rhodamine B Wastewater by a Recycling-Flow Two- and Three-dimensional Electro-Reactor System. Water 2020, 12, 1853. [Google Scholar] [CrossRef]
  78. Wei, L.; Guo, S.; Yan, G.; Chen, C.; Jiang, X. Electrochemical pretreatment of heavy oil refinery wastewater using a three-dimensional electrode reactor. Electrochim. Acta 2010, 55, 8615–8620. [Google Scholar] [CrossRef]
  79. Zhan, J.; Li, Z.; Yu, G.; Pan, X.; Wang, J.; Zhu, W.; Han, X.; Wang, Y. Enhanced treatment of pharmaceutical wastewater by combining three-dimensional electrochemical process with ozonation to in situ regenerate granular activated carbon particle electrodes. Sep. Purif. Technol. 2019, 208, 12–18. [Google Scholar] [CrossRef]
  80. Li, X.-Y.; Xu, J.; Cheng, J.-P.; Feng, L.; Shi, Y.-F.; Ji, J. TiO2-SiO2/GAC particles for enhanced electrocatalytic removal of acid orange 7 (AO7) dyeing wastewater in a three-dimensional electrochemical reactor. Sep. Purif. Technol. 2017, 187, 303–310. [Google Scholar] [CrossRef]
  81. Chen, Y.; Qin, Q.; Xie, S.; Nong, C.; Hu, J.; Li, M. Optimization of the Electrochemical Treatment of 4-chlorophenol Wastewater Using Response Surface Method. Electrochemistry 2020, 88, 282–289. [Google Scholar] [CrossRef]
  82. Xiong, Y.; He, C.; Karlsson, H.T.; Zhu, X. Performance of three-phase three-dimensional electrode reactor for the reduction of COD in simulated wastewater-containing phenol. Chemosphere 2003, 50, 131–136. [Google Scholar] [CrossRef]
  83. Wu, X.; Yang, X.; Wu, D.; Fu, R. Feasibility study of using carbon aerogel as particle electrodes for decoloration of RBRX dye solution in a three-dimensional electrode reactor. Chem. Eng. J. 2008, 138, 47–54. [Google Scholar] [CrossRef]
  84. Liu, Y.; Wu, Z.-y.; Peng, P.; Xie, H.-b.; Li, X.-y.; Xu, J.; Li, W.-h. A pilot-scale three-dimensional electrochemical reactor combined with anaerobic-anoxic-oxic system for advanced treatment of coking wastewater. J. Environ. Manag. 2020, 258, 110021. [Google Scholar] [CrossRef] [PubMed]
  85. Yan, L.; Ma, H.; Wang, B.; Wang, Y.; Chen, Y. Electrochemical treatment of petroleum refinery wastewater with three-dimensional multi-phase electrode. Desalination 2011, 276, 397–402. [Google Scholar] [CrossRef]
  86. Liu, Y.; Yu, Z.; Hou, Y.; Peng, Z.; Wang, L.; Gong, Z.; Zhu, J.; Su, D. Highly efficient Pd-Fe/Ni foam as heterogeneous Fenton catalysts for the three-dimensional electrode system. Catal. Commun. 2016, 86, 63–66. [Google Scholar] [CrossRef] [Green Version]
  87. Wang, C.-T.; Chou, W.-L.; Kuo, Y.-M.; Chang, F.-L. Paired removal of color and COD from textile dyeing wastewater by simultaneous anodic and indirect cathodic oxidation. J. Hazard. Mater. 2009, 169, 16–22. [Google Scholar] [CrossRef]
  88. Mohammadi, H.; Bina, B.; Ebrahimi, A. A novel three-dimensional electro-Fenton system and its application for degradation of anti-inflammatory pharmaceuticals: Modeling and degradation pathways. Process Saf. Environ. 2018, 117, 200–213. [Google Scholar] [CrossRef]
  89. Kumar, P.; Kumar, V.; Singh, J.; Kumar, P. Electrokinetic assisted anaerobic digestion of spent mushroom substrate supplemented with sugar mill wastewater for enhanced biogas production. Renew. Energy 2021, 179, 418–426. [Google Scholar] [CrossRef]
  90. Otter, P.; Mette, K.; Wesch, R.; Gerhardt, T.; Krüger, F.-M.; Goldmaier, A.; Benz, F.; Malakar, P.; Grischek, T. Oxidation of Selected Trace Organic Compounds through the Combination of Inline Electro-Chlorination with UV Radiation (UV/ECl2) as Alternative AOP for Decentralized Drinking Water Treatment. Water 2020, 12, 3275. [Google Scholar] [CrossRef]
  91. Nunes, R.R.; Ribeiro, R.; Morão, G.M.; Rezende, M.O.O.; Moreira-Santos, M. Treatment of Wastewaters Containing Sulfonylurea Herbicides by Electroflotation: Chemical and Ecotoxicological Efficacy. Water 2022, 14, 2723. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the reaction mechanism of an EO reactor.
Figure 1. Schematic diagram of the reaction mechanism of an EO reactor.
Water 14 03711 g001
Figure 2. Reactor configurations of ECRs.
Figure 2. Reactor configurations of ECRs.
Water 14 03711 g002
Figure 3. Schematic diagram of traditional plate-frame ECRs.
Figure 3. Schematic diagram of traditional plate-frame ECRs.
Water 14 03711 g003
Figure 4. Schematic diagram of traditional tubular ECRs.
Figure 4. Schematic diagram of traditional tubular ECRs.
Water 14 03711 g004
Figure 5. Schematic diagram of flow-through (a) and flow-by (b) mode of ECR.
Figure 5. Schematic diagram of flow-through (a) and flow-by (b) mode of ECR.
Water 14 03711 g005
Figure 6. The structure of 3D-ECR.
Figure 6. The structure of 3D-ECR.
Water 14 03711 g006
Figure 7. 3D fixed bed mode ECR (a) and 3D fluidized bed mode ECR (b).
Figure 7. 3D fixed bed mode ECR (a) and 3D fluidized bed mode ECR (b).
Water 14 03711 g007
Table 1. The treatment of different organic wastewater using plate-frame ECRs.
Table 1. The treatment of different organic wastewater using plate-frame ECRs.
Reactor TypeAnodeCathodeWastewaterRemoval RateRef.
filter-press reactorSb2O5-doped
Ti/RuO2-ZrO2
stainless
steel
indigo
carmine
90% COD in 120 min[45]
filter-press
reactor
Ti–Pt/β-PbO2
BDD
stainless steelEstrone 35% COD in 60 min
98% COD in 30 min
[46]
filter-press reactorBDDstainless steelindanthrene blue dye91% color in 180 min[47]
plate-frame reactorTi/SnO2-Sb2O5GDElandfill leachate80% TOC in 240 min[41]
single-compartment filter-press reactorBDDstainless steelpaint wastewater90% COD in 90 min[40]
plate-frame reactor with perforated liquid distributors Ti/RuO2-IrO2stainless steelsolid waste
leachate
98% COD in 60 min[42]
rotating bipolar disc plate-frame reactorFeFetextile wastewater90% COD in 12 min[43]
filter-press reactor with turbulence promoterBDDstainless-steelBPS84% TOC in 265 min[39]
Table 2. Treatment of different organic wastewater by tubular ECR.
Table 2. Treatment of different organic wastewater by tubular ECR.
Reactor TypeAnodeCathodeWastewaterRemoval RateRef.
traditional
tubular reactor
Pt/Ticarbon fibre dyeing wastewater44.3% COD
in 6 h
[48]
traditional
tubular reactor
Ti/RuO2stainless steelpetroleum refinery wastewater 85% COD
in 24 min
[49]
tubular reactor in batch with recirculation modeTi/Ti0.7Ru0.3O2stainless steeldye waste water75% of Acid red 87[55]
tubular reactor with mesh plate
electrodes
TiO2/RuO2TiEvans blue dye80% COD
in 100 min
[56]
tubular reactor with porous electrodesTi/SnO2-Sbstainless steelpyridine
wastewater
86% TOC
in 180 min
[57]
dual tubular membranes reactorTi/IrO2–Ta2O5CB-PTFE modified graphite tricyclazole79% COD
in 20 min
[58]
tubular reactor based on spiral flowTi/Ti4O7stainless steelmethylene blue simulation
wastewater
50% TOC
in 105 min
[54]
Table 3. Treatment of different organic wastewater by 3D-ECR.
Table 3. Treatment of different organic wastewater by 3D-ECR.
Reactor TypeElectrode MaterialsWastewaterTreatment ResultsRef.
3D fluidized bed ECR in batchAnode: stainless steel
Cathode: stainless steel
Particle electrode:GCA
PhenolCOD removal of 60% at an airflow of 5 min−1 and voltage of 30 V[82]
3D fixed bed ECRAnode: graphite
Cathode: stainless steel
Particle electrode: CAs
dye wastewaterdecolorization ratio was 98% (pH = 2.8, voltage = 20 V, treatment time = 30 min, air flow rate = 0.4 L min−1)[83]
recycling-
flow 3D-ECR
Anode: Ti/RuO2/TiO2
Cathode: Ti/RuO2/TiO2
Particle electrode: Columnar
GAC
Rhodamine B wastewaterCOD removal of 86.9% at HRT = 20 min, initial concentration = 100 min, volume = 500 mL, pH = 7, NaSO4 = 2 g/L and voltage = 5 V.[77]
pilot-scale plate-frame
fluidized bed ER
Anode: Ti/RuO2-IrO2
Cathode: Ti/RuO2-IrO2
Particle electrode: GAC
coking wastewaterremoval 94.4% of COD and
76.2% of TN at a low EC of
0.22 kWh/kg COD and 4.69
kWh/kg TN.
[84]
3D multi-phase ECRAnode: graphite
Cathode: graphite
Particle electrode: Fe
petroleum refinery wastewaterCOD removal of 92.8% at pH = 6.5, cell voltage = 12 V and fine Fe particle = 4 g and air flow rate = 1.5 L/min[85]
3D-ECRAnode: DSA
Cathode: Ti
Particle electrode: Sn/Sb-Mn-GAC
4-chlorophenol4-chlorophenol removal of 96.13% in 60 min at Na2SO4 concentration of 2 g·L−1, electrode distance of 2 cm, current intensity of 2 A, and particle dosage of 14 g.[81]
3D-ECRAnode: Ti/RuO2-IrO2
Cathode: stainless steel
Particle electrode: TiO2-SiO2/GAC
dye wastewaterdecolorization rate and COD removal efficiency are 83.20% and 48.95% at electric field intensity = 1 V/cm, TiO2-SiO2/GAC = 200 g/L, Na2SO4 = 0.07 mol/L, pH = 3 in 120 min.[80]
3D-ECRAnode: Pt
Cathode: Pt
Particle electrode:Pd-Fe/Ni
DimetridazoleThe dimetridazole removal of 96.5 at current density = 31 mA/cm2, C0 = 50 mg/L,
pH = 3, air flow rate = 1.0 L/min
[86]
3D batch type ECRAnode: Ti/Sb0.1Sn0.9O2
Cathode: stainless steel
Particle electrode: GAC and ceramsite particle (PCP)
oil refinery wastewaterremoval of COD, TOC and toxicity units were 45.5%, 43.3% and 67.2% (GAC percentage = 75%, current
density = 30 mA/cm2, pH not adjusted and treatment time = 100 min)
[78]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, J.; Ren, N.; Qu, C.; Lu, S.; Xiang, Y.; Liang, D. Recent Advances in the Reactor Design for Industrial Wastewater Treatment by Electro-Oxidation Process. Water 2022, 14, 3711. https://doi.org/10.3390/w14223711

AMA Style

Liu J, Ren N, Qu C, Lu S, Xiang Y, Liang D. Recent Advances in the Reactor Design for Industrial Wastewater Treatment by Electro-Oxidation Process. Water. 2022; 14(22):3711. https://doi.org/10.3390/w14223711

Chicago/Turabian Style

Liu, Jun, Na Ren, Chao Qu, Shanfu Lu, Yan Xiang, and Dawei Liang. 2022. "Recent Advances in the Reactor Design for Industrial Wastewater Treatment by Electro-Oxidation Process" Water 14, no. 22: 3711. https://doi.org/10.3390/w14223711

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

Article Metrics

Back to TopTop