Next Article in Journal
Multidimensional Ecosystem Mapping: Towards a More Comprehensive Spatial Assessment of Nature’s Contributions to People in France
Next Article in Special Issue
The Application of Chitosan-Based Adsorbents for the Removal of Hazardous Pollutants from Aqueous Solutions—A Review
Previous Article in Journal
Research on the Bearing Capacity and Sustainable Construction of a Vacuum Drainage Pipe Pile
Previous Article in Special Issue
Study on the Adsorption Characteristics of Calcareous Sand for Pb(II), Cu(II) and Cd(II) in Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Performance Stability and Regeneration Property of Catalytic Membranes Coupled with Advanced Oxidation Process: A Comprehensive Review

1
College of Environmental Science and Engineering, Dalian Maritime University, Dalian 116026, China
2
Tianjin Key Laboratory of Brine Chemical Engineering and Resource Eco-Utilization, College of Chemical Engineering and Materials Science, Tianjin University of Science & Technology, Tianjin 300457, China
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(9), 7556; https://doi.org/10.3390/su15097556
Submission received: 28 March 2023 / Revised: 30 April 2023 / Accepted: 3 May 2023 / Published: 4 May 2023
(This article belongs to the Special Issue Water Treatment by Adsorption and Oxidation)

Abstract

:
The performance stability and regeneration property of catalytic membranes should be considered from a practical viewpoint. In this review, recent works concerning this issue have been comprehensively summarized. The performance stability of catalytic membranes was first summarized in terms of the two commonly employed evaluation ways, i.e., multi-cycle experiments and long-time filtration. The main deactivation reasons included the adsorption of pollutants and its oxidation intermediates, the change in chemical properties of the membranes, as well as leaching of metal species. After that, the regeneration property of the catalytic membranes was discussed according to the different regeneration methods including solvent washing, heat treatment, advanced oxidation, and others. This review further highlights the current achievements and hurdles in the performance stability and regeneration property of catalytic membranes and proposes recommendations for future works, including more attention on changes in the kinetic constant, trans-membrane pressure/flux and removals of total organic carbon/chemical oxygen demand, the selection of suitable cycle numbers and operation time, the development of new regeneration methods, as well as more studies on the pilot and larger scales.

1. Introduction

Clean water is essential to the survival and development of human society, but the various organic pollutants in water, including dyes, antibiotics, pesticides, and endocrine disruptors, pose potential threats to the environment and human health. Among the various physical, chemical, and biological remediation technologies, the advanced oxidation process (AOP) is known as one of the most promising approaches considering its wide applicability to treat different pollutants and the ability to degrade or even mineralize them. Based on the oxidant or energy input used to produce active species, AOPs can be classified as persulfate oxidation, H2O2 oxidation (Fenton oxidation), ozone oxidation, photocatalytic oxidation, electrochemical oxidation, and others [1,2]. In the latter two processes, light (UV or visible light) and electric energy as external energy sources are introduced, respectively. In these AOPs, catalysts, especially the heterogenous ones, are generally utilized to boost the removal efficiency of those organic pollutants. However, after the catalytic oxidation processes, the suspended catalysts need to be separated from the treated water by filtration, centrifugation, or other methods, which is rather tedious and time-consuming.
The membrane technology is another widely investigated remediation method for water treatment. However, membrane separation is a physical process with which the organic pollutants can only be concentrated rather than degraded. Recently, by coupling AOP and membrane technology, catalytic membranes fabricated by loading catalysts on membranes have been developed and employed for the treatment of organic pollutants in water. The pristine membranes possess limited ability for pollutant degradation. When flowing through the catalytic membranes, the organic pollutants can be degraded under the catalytic effects of the loaded catalysts. After being treated with the membranes, water and catalysts are separated, and clean water as the filtrate is collected, avoiding the routine catalytic separation step. Meanwhile, during the filtration process, accelerated mass transfer of organic pollutants and oxidants from the bulk solution to the catalyst surface is achieved. Compared to conventional batch systems, the employment of catalytic membranes can lead to a higher production rate of oxidative species. Enhanced mass transfer can also promote the interaction between pollutants and the produced radicals, leading to increased reaction rate and enhanced removal efficiency. For instance, CoAl-layered double hydroxide (LDH) was loaded on polyvinylidene fluoride (PVDF) to fabricate the catalytic membrane of LDHm, which was then employed for the degradation of ranitidine (RA) [3]. The LDHm/peroxymonosulfate (PMS) system resulted in around 85% removal of RA in 10 min, while the removals with the pristine PVDF membrane and the bulk LDH/PMS system were only <10% and 37–44% [3]. The LDHm can be facilely separated after use, while the powdered LDH catalysts need to be separated by filtration or centrifugation. One disadvantage of LDHm compared to bulk LDH may be its increased fabrication cost, but the loading of LDH on the membrane can minimize the catalyst loss during repeated use. As another example, Co2+/Mxene loaded on PVDF showed 97% removal of tetracycline (TC) in the presence of PMS, while the removal with the PVDF/PMS system was only around 40% [4]. For comparison, the catalytic membrane was immersed in a solution containing both TC and PMS to conduct AOP in the batch mode, showing a poor TC removal of around 60%, which demonstrated the indispensable role of membrane filtration in the catalytic membrane-based degradation process [4].
Catalytic membranes can be fabricated in many approaches. The most widely used one is the filtration method, where catalytic particles were dispersed in solvents and then loaded on a substrate membrane by filtration. Other methods such as phase inversion, hydrothermal treatment, impregnation, electro-deposition, wet-spinning, atomic layer deposition, etc. were also widely used. To conduct AOP with catalytic membranes, four different experiment modes have been used based on the employed filtration mode (dead-end or cross-flow) and the operation mode (single pass or recirculation), which are shown in Figure 1a–d. The membranes can also be used in the batch mode where no filtration is performed (Figure 1e). This method is widely employed in the literature, but in this case, the membranes acted as structured catalysts rather than membranes, which is outside the scope of this current review and not discussed below.
Ideally, catalytic membranes should be repeatedly used with stable activity. However, although organic pollutants can be degraded with them, complete mineralization of the pollutants is hard to achieve. As a result, oxidation intermediates were produced and accumulated on the membrane, leading to membrane fouling and decreased catalytic activity. Considering this, in order to evaluate the lifespan of a certain catalytic membrane, its performance stability and regeneration property should be investigated. When evaluating the performance stability, the catalytic membrane was used for multi-cycles or for a long operation time without any treatment. In this process, the substrate membrane and the catalysts loaded on them were used as a whole, and there was no need to separate them. The regeneration property is investigated in a different way, where the used membrane was regenerated by a certain method (solvent washing, heat treatment, or other methods) and then used again. Note that in some cases, the catalysts on the substrate membrane were separated, regenerated, and loaded again, and performance of the re-fabricated catalytic membrane was compared with the fresh one to illustrate the regeneration efficiency. This method is only applicable to the catalytic membranes formed by filtration, where the catalysts generally weakly interact with the substrate membrane and can be easily separated by back washing or other physical methods.
In the past decades, catalytic membranes have been extensively investigated for water treatment, and several excellent reviews have been published faced with this fast-developing field. For example, in our previous work [5], the coupled process integrating membrane filtration and electrochemical oxidation was reviewed, where the strategies for process design, the membrane materials, and the specific features of the coupled process were discussed. Electroactive membranes were reviewed in another work [6], which included an energy analysis of the overall process. In a more recent review [7], the application of catalytic membranes coupled with electrochemical oxidation for the treatment of antibiotic residues was summarized, focusing on the selection of various membrane electrode materials and their modification methods. The coupling of membrane filtration with peroxide and persulfate was reviewed in several other works [8,9,10], where the different process configurations, membrane materials, fabrication methods, and reaction mechanisms were summarized and discussed. The catalytic membranes coupled with various AOPs were discussed in a recent work in terms of fabrication methods and reaction mechanisms [11]. In another work [12], the removal of organic pollutants from wastewater by AOP was reviewed, where the coupling of AOP with membrane filtration was discussed based on the process figurations.
In these previous review works, although the performance stability and regeneration property of several catalytic membranes were briefly mentioned, the contents were mainly concerned with material design and mechanism analysis. To the best of our knowledge, no recent review paper focusing on the performance stability and regeneration property of catalytic membranes for the removal of organic pollutants from water is available. In this work, we intend to present a comprehensive review focusing on this issue based on the literature in the recent five years (2017–2022). The performance stability of catalytic membranes was first summarized in terms of the two commonly employed evaluation ways, i.e., multi-cycle experiments and long-time filtration, where the different deactivation reasons were discussed. After that, the regeneration property of the catalytic membranes was summarized and discussed according to the different regeneration methods including solvent washing, heat treatment, and others. This review further highlights the current achievements and hurdles in the performance stability and regeneration property of catalytic membranes.

2. Reaction Mechanisms and Impacting Factors of Catalytic Membranes Coupled with Advanced Oxidation Process

When flowing through the catalytic membranes, the organic pollutants and the oxidants would react with each other thanks to the catalytic effects of the composite membranes, leading to the degradation of these organic pollutants. Note that adsorption, which is generally involved in heterogeneous reactions, may also play a role in this process. However, if adsorption is the main mechanism responsible for the removal of pollutants when treated with a membrane, this membrane should be considered as an adsorption membrane rather than a catalytic membrane. Saturated adsorption membranes cannot be reused again without desorption treatment, which is out of the scope of the current review.
In the catalytic filtration process, the direct oxidation of the pollutants by the oxidants may make some contributions, but in most cases, the reaction is not this straightforward. In this section, the reaction mechanisms for the catalytic membrane-based AOP will be briefly summarized based on previous reports [5,11,12].
As mentioned above, AOPs can be classified as persulfate oxidation, H2O2 oxidation (Fenton oxidation), ozone oxidation, photocatalytic oxidation, electrochemical oxidation, and others such as catalytic wet oxidation and sonolysis. In all of these processes, the input of oxidants and/or energy is required, which pave the oxidation of pollutants. The input oxidants include persulfate, H2O2, O3, and O2. The input energy is UV/visible light and electric energy for photocatalytic and electrochemical oxidation, while in other cases, heat should be considered as the input energy, although the reactions are usually conducted at room temperature. The oxidation pathways can be generally classified into the following two series (Figure 2):
(a)
the radical pathway, where various radicals are produced and then react with the pollutants;
(b)
the non-radical pathway, where the oxidation is achieved by direct electron transfer or with the assistance of other reactive species other than radicals.
Generally, various radical species including the hydroxyl radical (·OH), sulfate radical (SO4·−), superoxide radical (O2·−), and others may be involved in the radical pathways. These radical species are produced in different procedures among different AOPs. For photocatalysis, the absorption of the light source led to the excitation of the semiconductive catalyst, producing valence band holes (h+) and conduction band electrons (e), and h+ further reacted with water to generate ·OH. For electrochemical oxidation, ·OH is produced through water oxidation on the anodes. For persulfate oxidation, Fenton oxidation, and ozone oxidation, ·OH is produced by the reactions between these oxidants (persulfate, H2O2, and O3) and the catalytic species. SO4·− is a unique radical produced from decomposition of persulfate, but it may also become involved in electrochemical oxidation when sulfates are used as the supporting electrolyte. Several methods have been employed for the determination of oxidative species in AOP systems. Conducting quenching experiments is an indirect method, where they can be identified based on the varied reaction rates between scavengers and them. Electron spin resonance (ESR) is another widely used approach, where the radicals are trapped by 5,5-dimethyl-1-pyrroline N-oxide (DMPO) and other reagents to produce spin adducts, providing qualitative and quantitative information with their hyperfine coupling constants and signal intensities. The concentrations of the oxidative species can also be determined by the high-performance liquid chromatography method through the reaction of benzoic acid and other probes with them. The non-radical pathways are also complex, where various mechanisms including singlet oxygen (1O2), high-valent metal-oxo species, as well as direct electron transfer may all make some contributions.
Even in a specific AOP, the involved reaction pathways may vary from case to case, which depends on the type of catalyst, the structure of the target pollutant, the reaction conditions, and the co-existent species in the water matrix. More in-depth discussions on the reaction mechanisms in the AOPs are out of the scope of the current review. No matter what mechanisms are involved, all of these pathways will lead to the consumptions of oxidants and energy, resulting in the oxidation of the organic pollutants in water. To evaluate the performance of this process, the removal efficiency, which is calculated by measuring the change in concentration of the target pollutant before and after treatment, is the generally utilized index. For catalytic membrane-based AOP, other indexes, such as the reaction time and apparent kinetic constant in the recirculation mode as well as the trans-membrane pressure (TMP) and flux in the single pass mode, should also be paid attention to. In addition, remember that complete mineralization of the pollutants is hard to achieve, and the produced oxidation intermediates may pose a potential threat. To evaluate the mineralization degree, the removals of total organic carbon (TOC) and chemical oxygen demand (COD) can be also measured.
The performance of a certain catalytic membrane to removal a certain target pollutant may be impacted by many factors. Operation conditions, including initial concentration of the pollutant, the dosage of oxidant, the amount of energy input (temperature, intensity of light, current density, etc.), and flux, will certainly play some roles. Environmental factors, including pH, inorganic ions, and natural organic matter, may also influence the removal efficiencies. The pH will influence the charge properties of both the catalytic membrane and the pollutant molecules as well as the stability of the oxidants, altering their interactions and changing the reaction results. The inorganic ions and natural organic matter can take part in the radical reactions, playing either positive or negative roles. These factors may also influence the performance stability and regeneration property of catalytic membranes by changing the level of metal leaching or the concentration and distribution of oxidation intermediates.
In the following sections, the performance stability and regeneration property of catalytic membranes are reviewed with the results in previous works summarized in tables. The removal efficiencies as the main index evaluating the performances of catalytic membranes are listed in the tables for comparison, together with some important factors including the target pollutants, the initial concentrations, and the fluxes. The performance stability of catalytic membranes is first discussed in Section 3, after which the regeneration property of the catalytic membranes is discussed in Section 4.

3. Performance Stability of Catalytic Membranes Coupled with Advanced Oxidation Process

Generally, the performance stability of catalytic membranes can be evaluated in two ways. In the first method, after the fresh membrane was used for the first time, it was directly used again in the catalytic process without any regeneration treatment. The catalytic run can be performed in several cycles with the same membrane. The same operational factors were used, and the removal efficiencies of the target pollutant in all these catalytic runs were recorded and compared. In the second method, the catalytic run was conducted continually for a long period of time, and samples of the filtrate were withdrawn periodically to record the change in removal efficiency of the pollutant. In this section, recent works regarding these two evaluation methods will be discussed in Section 3.1 and Section 3.2, with the previously reported experimental results summarized in Table 1 and Table 2, respectively.

3.1. Performance Stability of Catalytic Membranes Tested by Multi-Cycle Experiments

In this section, the performance stability of catalytic membranes evaluated by multi-cycle experiments was summarized and analyzed (Table 1), where the related reports were classified based on the type of advanced oxidation process. The employed substrate membrane, catalyst, fabrication method, membrane type (flat-sheet, tubular, or hollow fiber), filtration type (dead-end or cross-flow), operation mode (single pass or recirculation), target pollutant, initial concentration of pollutant, flux, removal efficiency of pollutant, and number of cycles are provided in the table.
For comparison of the performance stability results in these works, a dimensionless factor (η) was determined as follows:
η = Rn/R1 × 100%,
where n stands for the number of catalytic runs conducted for the same membrane. Rn represents the removal efficiency of the target pollutant obtained in the last catalytic run, while R1 stands for the value obtained with the fresh membrane. Thus, η should be in the range of 0–100%, and a larger η indicates better performance stability within the same number of catalytic cycles.
As summarized in Table 1, for the performance stability tests of these catalytic membranes tested by multi-cycle experiments, the employed number of cycles were in the range of 3–30, obtaining η values in the range of 26–100%. Obviously, a larger number of cycles and a higher η value should indicate better performance stability. From this point of view, three catalytic membranes, including the CoAl-LDH/PVDF membrane, the Mn2O3/ceremic membrane, and the free standing reduced graphene oxide (rGO)@TiO2 membrane showed outstanding performances, all of which could be repeatedly used for 10 cycles with almost no deterioration in the removal efficiencies towards their target pollutants. Another two membranes, i.e., the CuMn2O4/ceramic membrane and the N-rGO/ceramic membrane, also performed very well, showing η values over 90% after repeatedly used for more than 10 cycles (30 and 18 cycles), respectively.
In the following sub-sections, the summarized results in Table 1 are further discussed based on the type of advanced oxidation process.

3.1.1. Multi-Cycle Performance of Catalytic Membranes Coupled with Persulfate Oxidation

In persulfate-based oxidation processes, persulfates including PMS and peroxydisulfate (PDS) are utilized as oxidants. In catalytic persulfate oxidation systems, various reactive oxidative species (ROSs) can be generated, such as ·OH, SO4·−, O2·−, 1O2, and high-valent iron-oxo, which then react with the organic pollutants. In addition, the pollutants can also be oxidized through non-radical electron transfer, where the catalyst acts as a medium and bridges the electron transfer between persulfate and the pollutants. Thus, both ROSs and the non-radical electron transfer pathway can lead to oxidation, decomposition, and mineralization of the various organic pollutants in water. In powder-based heterogenous systems, metal-based and carbon-based catalysts have been widely investigated and used for the activation of persulfates. When coupled with the membrane process, metal or carbon catalysts need be introduced to the substrate membranes to fabricate catalytic membranes.
The substrate membranes can be either organic or inorganic. The PVDF membrane is the most frequently used organic substrate membrane. For example, the powder catalyst of CoAl-layered double hydroxide (LDH) was loaded on a PVDF substrate membrane by vacuum filtration. The catalytic membrane was then used for the activation of PMS to degrade RA, which was a histamine H-receptor antagonist for the treatment of gastrointestinal disease, achieving a removal efficiency of 94% under optimum conditions [3]. The removal efficiency remained stable in 10 cycles without obvious changes, indicating the high stability of the LDH catalytic membrane [3]. In another report [4], Mxene modified by Co2+ was also loaded on a flat PVDF membrane by vacuum filtration. The resulting Co2+/Mxene catalytic membrane could efficiently degrade 98.2% TC, which was an antibiotic, in the presence of PMS. When reusing the membrane, the removal efficiency dropped to 90.2% and 82.5% in the second and third run, respectively [4]. When FeOCl/MoS2 was used as the catalyst in another work [13], a similar vacuum filtration method was utilized to immobilize the catalyst to a PVDF membrane, except that the substrate membrane was modified first by coating with polydopamine (PDA). The FeOCl/MoS2 membrane was employed to remove Rhodamine B (RhB) in five cycles, showing acceptable performance stability with a η value of 87% [13]. In some reports, composite catalysts constructed from both metal and carbon materials are investigated. For instance, MnO2/carbon nanotube (CNT) [14] and FeCoS@N-rGO [15] were combined with PVDF, and the obtained catalytic membranes were used to degrade RhB and sulfamethoxazole (SMX), respectively. Both membranes could maintain high removal efficiencies in multicycle tests, showing η values of 100% [14] and 91% [15] after 8 and 5 cycles, respectively.
Besides the vacuum filtration method, the phase inversion method is another commonly employed approach to fabricate PVDF-based catalytic membranes. For instance, cobalt nanoparticles were encapsulated into nitrogen-doped carbon shells, and the resulting composite material (Co@N-C) was introduced within the pores of PVDF membrane by the phase inversion method [16]. Using this membrane, the degradation of TC was operated in the cross-flow mode, achieving 99.3% removal in 60 min. After being used for 5 cycles, the removal efficiency was still above 90%. The high stability of the PVDF/Co@N-C membrane was not only attributed to the stability of the Co@N-C catalyst but also to the strong adhesion of the catalysts to the membrane channels [16].
In addition to PVDF, other types of organic substrate membranes are also investigated, including polytetrafluoroethylene (PTFE), polyacrylonitrile (PAN), polysulfone (PSF), nylon, and others. Similar to the case of PVDF, these organic membranes could not activate persulfate, acting as supports only. To introduce catalytic species, various metal-based and carbon-based catalysts as well as their composite were loaded, including Co3O4@NCNTs/g-CN [17], Fe2O3@Co0.08Fe1.92@nitrogen-doped rGO@CNT (FeCo@GCTs/GO) [18], N-doped GO [19], Fe-doped LaCoO3 [20], SrCuxCo1−xO3−λ-deposited MCM-41 (SCCM) [21], Fe2O3@CNT [22], and nitrogen-doped graphene (NG)/rGO/CNT [23]. These as-synthesized catalytic membranes possessed high removal efficiencies of over 90% for the target pollutants, but the efficiencies declined after being used cyclically for 3–5 times. For the membrane of SCCM/PSF, although a η value of ~100% was achieved in the fifth cycle for the degradation of Rhodamine B, the authors noted that a minor decline in the removal efficiency could be observed and attributed this to the adsorbed organic molecules and their degradation products [21]. This is a generally mentioned reason when discussing the declined performance in multi-cycle degradation tests. The adsorbed organic species could block the active sites, hindering the adsorption and activation of continuously entering persulfates and consuming them as well. Consequently, the performance of the catalytic membrane gradually deteriorates in multi-cycle tests due to the accumulation of these organic species.
Moreover, other reasons may also lead to decreased activities in multi-cycle tests. For metal-based catalysts, leaching of metal species is always observed. Especially, for persulfate-based oxidation processes, the addition of persulfates would shift the solution pH to more acidic values and enhance metal leaching. These leached ions may promote the degradation of organic pollutants through homogeneous reactions. Thus, the loss of metal ions through leaching may be a possible reason for the weakened catalytic performance of the membrane. Considering this, the concentration of leached metal ions is generally measured. For instance, with the FeCo@GCTs/GO membrane, the leached concentrations of total Fe and Co ions were measured to be 0.24 and 0.31 mg/L [18]. Furthermore, when conducting SMX degradation with the dissolved ions, low removal efficiencies less than 7% were observed for both Fe2+ and Co2+ [18]. This result suggested that homogeneous activation under this low metal leaching level only made a minor contribution, and the degradation of the pollutants was mainly achieved through heterogenous reactions. However, in the membrane filtration process, metal species continuously leached out, and the content of metal catalysts in the membrane became lower and lower. Thus, even a low leaching level may lead to distinct decrease in the heterogenous catalytic activity due to the accumulation of metal loss in multi-cycle catalytic filtration runs. Another reason for the decreased activity of metal-based catalysts is the change in valence states. Although metal species of both low and high-valence states could activate persulfates through cyclic conversion, the transformation from high-valence species (Co(III), Fe(III), etc.) to low-valence species (Co(II), Fe(II)) is much slower. As a result, when the fraction of high-valence species in the catalysts is increased, lower catalytic performance will be expected. As an example, after being used for the degradation of TC, both the ratios of Co(II)/Co(III) and Fe(II)/Fe(III) in the Fe-doped LaCoO3/PAN membrane decreased, which may result in the slightly declined removal efficiency after five repeated runs [20].
When it comes to metal-free carbon-based catalytic membranes, other reasons may also make a contribution. As an example, the nitrogen-doped graphene (NG)/rGO/CNTs/nylon membrane was utilized to active PDS for the degradation of SMX [23]. The fresh membrane possessed an ID/IG ratio of 1.82 as characterized by Raman spectroscopy, indicating the defects-rich structure of the carbon catalysts and exhibiting strong adsorption and catalytic activity. After 5 filtration cycles, the ID/IG ratio was decreased to 1.50, indicating the consumption of these surface defects and the production of oxygen-containing functional groups, which was responsible for the reduction in removal efficiency [23]. In another report, the decreased efficiency of the N-doped GO/PTFE membrane after PDS activation was partially attributed to the attack of radical species on GO, which led to the formation of GO fragments and broke the structure of the membrane [19].
Compared to organic polymeric membranes, inorganic membranes are more costly due to their energy and time-consuming production procedures, but they also possess several merits including superior thermal, mechanical, and chemical stabilities as well as long service life. Thus, inorganic membranes are also widely investigated as substrate membranes to fabricate catalytic membranes. Inorganic membranes can be roughly classified as ceramic membranes, metal membranes, and carbon membranes, among which ceramic membranes are employed in a higher frequency. Similar to the case with polymeric membranes, the pristine ceramic membranes, which are often synthesized using Al2O3 as the main matrix material, show limited activity in persulfate-based oxidation processes. Thus, active metal species should be added to fabricate catalytic membranes. These additional species can be added in two ways, i.e., during the preparation of the ceramic membranes, or through post-treatment modification of synthesized ones. For instance, a CuO-doped ceramic membrane was constructed by the combined phase-inversion and liquid-phase sintering method starting from Al2O3, TiO2, and CuO powders [24]. In several other reports [25,26], CuO was also employed as the active metal species, but it was introduced by post-treatment modification through an impregnation and calcination method using aqueous solutions containing Cu salts. Filtration can also be used to modify the pristine ceramic membrane, such as in the case of loading Co3O4@CNT [27]. When Mn2O3 was used to modify the ceramic membrane in another work [28], the pristine membrane was sprayed with Mn2O3-Al2O3 slurry and then calcinated, obtaining the catalytic membrane.
In a recent report [29], the authors intended to employ Co-coated granular-activated carbon (Co@GAC) as the catalytic species. The above methods are not applicable to introduce Co@GAC to the ceramic membrane because the GAC could not withstand the high-temperature calcination process. Thus, the authors designed a novel method by filling the channels of the ceramic membrane with Co@GAC (Figure 3). The Co@GAC-membrane module also contained a PP punching mesh, which could prevent Co@GAC from flowing out of the channels during filtration process [29].
The above-mentioned catalytic ceramic membranes were employed to degrade various organic pollutants including Rhodamine B, sulfadiazine (SDZ), BPA, and acetaminophen with the assistance of PMS, all possessing high performance stabilities with η values larger than 90% after being used for 5–10 cycles. Similar to the organic substrate membrane-based ones, the slightly decreased performance could be attributed to the adsorption of pollutants/oxidation products, the slight metal leaching, and/or the change in their surface properties such as the valence states of the active metal species.

3.1.2. Multi-Cycle Performance of Catalytic Membranes Coupled with H2O2 Oxidation

In the Fenton oxidation process, H2O2 was used as the oxidant. The reaction between H2O2 and catalysts, such as Fe2+ and Fe3+ ions, led to the production of ·OH, which then paved the oxidation of organic pollutants. Compared to homogenous ions, heterogeneous iron-based catalysts are more environmentally friendly and can be loaded on substrate membranes for Fenton oxidation applications. For example, a composite membrane was synthesized by combining poly(tetrafuoroethylene-co-hexafuoropropylene) (FEP) and Fe2O3 in a thermoforming method, where FEP resin, Fe2O3, and dioctyl phthalate (DOP) as a plasticizer were mixed and then treated with a thermo-compressor at 270 °C [30]. The membrane with 10 wt.% Fe2O3 could remove 99.69% methylene blue (MB) when operated in the cross-flow mode, and the removal efficiency declined to 95.15% and 72.85% in the second and third degradation runs. The main reason for the decreased performance was attributed to the blocking of pores in the membranes. By decreasing the content of Fe2O3 to 5 wt.%, the performance stability was increased, achieving ~91%, 90.85%, and 82.35% in three consecutive runs [30]. In another report, goethite and maleate ferroxane as green and inexpensive iron oxy-hydroxides (FeOOH) were combined with PAN by a phase inversion method, and the resulting catalytic membranes could efficiently remove amoxicillin in the dead-end filtration mode [31]. Close removal efficiencies of 86.3% and 92.3% were obtained with goethite and maleate ferroxane, respectively, and the slight reductions in multi-cycle runs were ascribed to the accumulation of organic molecules on the membrane surface and the concentration polarization phenomenon [31].
In the traditional Fenton oxidation process, the reduction of Fe(III) to Fe(II) is generally sluggish due to the inherently slow kinetics, and H2O2 should be added externally. Recently, electrical or light energy has been proved to speed up the regeneration of Fe(II) species. Meanwhile, H2O2 can sometimes be produced in situ in these processes, avoiding the safety issues related to its storage and transportation. These emerging Fenton oxidation processes, known as electro-Fenton and photo-Fenton oxidation processes, are discussed later in Section 3.1.4 and Section 3.1.5 below.

3.1.3. Multi-Cycle Performance of Catalytic Membranes Coupled with Ozone Oxidation

Ozone is another widely employed oxidant for pollutant degradation. Although ozone can directly oxidize some pollutants, low removal efficiency and limited mineralization are achieved with sole ozone. In contrast, catalysts can trigger the decomposition of ozone to produce ·OH with a higher redox potential, favoring the oxidation and mineralization of organic pollutants.
The multi-cycle performance of catalytic membranes in the ozonation process has been reported in several reports. CuO was loaded on a tubular ceramic membrane, and the fabricated catalytic membrane could remove 1,4-dioxane stably in 4 cycles in the presence of ozone [32]. The performance stability of the composite membrane was attributed to the low leaching of Cu, which was measured to be 0.58 ppm in each cycle [32]. In another work, MnMe oxides (Me = Fe, Co, Ce) were embedded on ceramic membranes through filtration, in situ reaction with KMnO4, and calcination. The one loaded with MnCe oxide showed the best catalytic ozonation performance among them three, and the η value was found to be 94% after 5 degradation runs, which was attributed to the slight membrane fouling as well as leaching of Mn and Ce [33]. For the degradation of benzophenone-3 (BP-3), another bi-metal oxide CuMn2O4 was loaded on a ceramic membrane through impregnation and calcination [23]. The prepared catalytic membrane showed excellent performance stability, showing a high η value of 91% after being used for 30 cycles. The main reason for the slightly decreased removal efficiency was inferred to be leaching of Cu and Mn, which were around 0.032 and 0.230 mg/L, while the contribution from the accumulated oxidation intermediates was considered to be negligible [34]. In contrast, the declined removal efficiency for the nitrogen-doped reduced graphene oxide (N-rGO) on the ceramic membrane, which was a metal-free catalyst, was mainly attributed to the adsorption of oxidation products on its active sites [35].

3.1.4. Multi-Cycle Performance of Catalytic Membranes Coupled with Photocatalysis

In the advanced oxidation process, photo irradiation can be introduced as an external energy source, promoting the production of holes, ·OH, and other oxidative species. In some cases, additional oxidants such as H2O2 and persulfates are added, which are known as photo-Fenton and photo-driven persulfate activation processes. Traditionally, ultraviolet (UV) radiation is usually used as the light source in the photocatalysis process. However, UV only takes a small part in the natural sunlight (<5%), and the utilization of visible light is receiving increasing attention recently.
By loading photocatalysts on various substrate membranes, a series of composite membranes have been fabricated and successfully utilized for the degradation of different organic pollutants. In a recent report, TiO2, the most widely investigated photocatalyst, was loaded on three different polymer membranes including PSF, PVDF, and PFTE [36]. The three catalytic membranes were then used for the photocatalytic degradation of diclofenac and ethinylestradiol. Interestingly, although the TiO2/PSF membrane possessed the highest activity, it was damaged after the first run due to the low resistance of PSF to UV exposure. The other two membranes were tested for three consecutive runs, and the declined removal efficiencies were attributed to the adsorption of pollutants on the membrane as well as leaching of TiO2 [36]. In another work, TiO2 was loaded on a ceramic membrane, which could stably degrade MB in four cycles [37]. The composite of graphene oxide (GO) and TiO2 was loaded on another ceramic membrane, and the degradation of three organic pollutants was conducted [38]. Under both dead-end and cross-flow filtrations modes, the multi-cycle performances were investigated by calculating the rate constants, and different results were revealed. No activity loss but a slight enhancement was observed for the dead-end filtration mode owing to the compaction effect during pressure-driven filtration. For the cross-flow filtration mode, some loss in activity was found, which was ascribed to the shear force under high flow rate [38]. The GO-TiO2 composite was loaded on a cellulose acetate substrate membrane in another work [39], and the catalytic membrane showed consistent dye removal ability for Congo red (CR) in three cycles. To improve the photocatalytic efficiency, TiO2 was modified by nitrogen doping to obtain N-TiO2 and then loaded on a PSF membrane together with GO [40]. The prepared catalytic membrane could remove MB under the irradiation of sunlight, but the photocatalytic ability decreased along with repeated runs, which was caused by the loss of catalysts. Different from the filtration fabrication method used in these works, the catalyst of reduced GO (rGO)@TiO2 was loaded on a stainless-steel mesh by electro-spinning and calcination to fabricate the rGO@TiO2 catalytic membrane [41]. The membrane possessed a highly stable degradation performance for propranolol with no obvious deterioration over the ten-cycle tests, and this was probably owing to the unique preparation method, which allows the catalysts to form a free-standing membrane [41]. Besides TiO2-based catalysts, other types of catalysts are also investigated. For instance, ZnIn2S4, which has a narrow band gap and shows high activity under the irradiation of visible light, was deposited on PVDF and then utilized for the degradation of fluvastatin. After six catalytic filtration cycles, a high η value of 94% was revealed, and the decreased removal efficiency was caused by the occupation of the active sites and catalyst loss due to the weak interaction between the PVDF substrate membrane and the ZnIn2S4 catalyst [42].
When H2O2 is added as an external oxidant in the photocatalytic process, the degradation of organic pollutants may be enhanced. For instance, the ceramic membrane with TiO2 loaded on it could more efficiently remove MB under visible light in the presence of additional H2O2, compared to the case without it [37]. The authors noted that the reactive species resulted from the transformation between Ti3+ and Ti4+ in the TiO2 catalyst, operating in a Fenton-like mechanism similar to the couple of Fe2+ and Fe3+ [37]. More commonly, the photo-Fenton process is operated with iron-based catalysts. For example, α-Fe2O3 as a widely used heterogenous Fenton catalyst was loaded on ceramic membrane, which was then employed for the photo-Fenton degradation of TC, showing a η value of 88% after 5 catalytic runs [43]. Another widely investigated Fenton catalyst, β-FeOOH, was loaded on PAN membrane [44] and air-laid paper [45], respectively. The photo-Fenton activities of the two fabricated membranes were then investigated, and both declined slightly after five catalytic cycles. In another work, a novel ferrocene-based membrane was fabricated by layer-by-layer interfacial polymerization and ion-exchange using PSF membrane as the substrate, which could remarkably remove bisphenol S (BPS) through the photo-Fenton process [46]. The reusability of the membrane was tested by three consecutive runs. The removal efficiencies of BPS in the second and third cycles were not provided, but the degradation rate in the three cycles calculated by the pseudo-first-order kinetic model was reported, which declined from 0.092 to 0.068 min−1 [46]. In these works, the reasons for the weakened performances of the used membranes were not discussed. Based on the low metal leaching and stable properties of the membranes before and after use in these works, the declined activities were probably attributed to the adsorbed degradation intermediates, which were generally detected and analyzed in these reports [43,44,45,46].
When persulfate instead of H2O2 is added, the photo-driven persulfate activation process is developed. In this process, classical metal catalysts based on Co or Fe are generally employed. For example, ZIF-67, a Co-containing metal organic framework (MOF), was loaded on a PP membrane by in situ synthesis to remove MB and methyl orange (MO) in the co-presence of visible light and PMS [47]. In another work, α-Fe2O3 and bacterial cellulose were together loaded on PTFE membrane, where bacterial cellulose helped to disperse α-Fe2O3 particles and thus enhanced the visible light absorption ability of the catalytic membrane [48]. For the treatment of berberine, which was an antibiotic, CoFe2O4 and carbon nanofiber were loaded on PVDF membrane. Interestingly, with the assistance of Ag/ZnO nanorods array/Ni foam as the photo-anode, the CoFe2O4/carbon nanofiber/PVDF membrane cathode also possessed electricity production, providing a promising strategy towards water treatment with energy recovery [49]. High η values after several cycles were observed for these three membranes, and the slight reduction was attributed to the incomplete degradation of contaminants during the former cycles [47].
In the above three reports, persulfate as the additional oxidant was added along with the feed solution. Interestingly, in another work, the persulfate of PDS was not added in the feed but loaded on the membrane. In that work, PVDF was used as the substrate membrane, and it was modified by N,N-dimethylaminoethyl methacrylate to introduce positively charged sites for the electrostatic attraction of negatively charged PDS anions [50]. The membrane with prior loaded PDS could degrade ofloxacin under UV radiation thanks to the photo-driven activation of PDS. However, due to the gradual consumption of PDS, the membrane could only be used for several times. The author noted that by combining with cathode electrolysis or other technologies, the UV activation process of PDS may be revisable, extending the service life of the membrane [50].

3.1.5. Multi-Cycle Performance of Catalytic Membranes Coupled with Electrocatalysis

Similar to the case in photocatalysis, electron energy can also be introduced into AOP as an external energy source. When conducting electrochemical advanced oxidation, two electrodes are required, i.e., an anode and a cathode. In some cases, the oxidation of organic pollutants is achieved on the anode through either direct oxidation or indirect oxidation, and the cathode only acts as a counter electrode. In some other cases, the cathode also plays an important role. For example, in the electro-Fenton process, the cathode promoted the in situ generation of H2O2 with the supply of O2. In the electron-activated persulfate oxidation process, the cathode contributed to the activation of persulfate. In these two processes, the regeneration of low-valent metal species was also achieved on the cathode, which provided electrons for the reduction of their high-valent counterparts. When coupled with the membrane process, the catalytic membranes can act as either or both, and the compositions of both anodes and electrodes are listed in Table 1 for clarity.
Conductive porous supports made of metal such as stainless steel and titanium can be directly modified to fabricate electrochemical catalytic membranes. For example, Pd was supported on porous titanium plate by spray coating and calcination to fabricate a membrane anode. Using titanium mesh as a cathode, the electrochemical membrane system could remove SMX stably in 10 cycles [51]. In another work, the same RuO2-plated titanic mesh was used as both anode and cathode, except that the anode was further modified through electrospinning and carbonization to load a layer of graphene/SnO2/carbon nanofibers. The system also possessed high stability for SMX degradation with a η value of 94% after 6 cycles [52]. A porous Ti plate was modified to introduce blue TiO2 on its surface, which could efficiently and stably remove triclosan by sequential adsorption and electrochemical degradation [53].
Non-conductive supports such as PTFE and PVDF can also be modified to construct electrochemical catalytic membranes, but an additional conductive support is generally required. For instance, a CNT/PTFE membrane was combined with a stainless-steel network to act as the anode, and a pristine stainless-steel network served as the cathode. The system could degrade several organic pollutants including SMX, ciprofloxacin, and amoxicillin, showing different η values of 72%, 79%, and 92% after 4 cycles. The weaker performance stability of the former two pollutants was probably attributed to the high molar weight oxidation products formed during their degradation, which accumulated on the membrane and led to higher mass transfer resistance [54]. Single cobalt atom/nitrogen atom co-doped graphene [55] and nano porous carbon/CuFeO2 [56] were employed for the modification of PTFE and PVDF membranes in another two reports, which were then used as anodes for the electrochemical oxidation of two dye pollutants (MB and MO), respectively. In the latter case, the deceased removal efficiency upon repeated use of the membrane was attributed to the reduction of active sites caused by membrane fouling [56].
In the above reports, the modified catalytic membrane is generally employed as the anode, and an additional cathode is required. Interestingly, a novel CNT-based membrane with a sandwich-like structure was designed and fabricated, which consisted of an outer CNT layer, a middle PVDF layer, and an inner CNT layer [57]. The composite membrane served as a complete electrochemical system where the two CNT layers acted as two electrodes and the PVDF layer as an insulating separator. With this system, microcystin-LR (MC-LR) can be stably removed through adsorption and electrochemical oxidation cycles [57].
In a recent work, an electrocatalytic ceramic membrane filtration system was developed for the degradation of 2,4-dichlorophenoxyacetic acid (2,4-D). H2O2 played an indispensable role in the reaction system, which was generated from continuously supplied oxygen [58]. This was different from the photo-Fenton process discussed above, where H2O2 was added rather than generated in situ. Reusability tests showed that the reaction system could remove 2,4-dichlorophenoxyacetic acid stably in 8 cycles, demonstrating the excellent stability of the membrane [58].
A MnFe2O4-rGO-PVDF composite membrane was fabricated by vacuum filtration and then employed as the cathode in the electron-activated persulfate activation system. The removal efficiency of oxytetracycline (OTC) gradually decreased after multiple use, which was ascribed to inevitable material loss caused by mechanical wear during assembly and disassembly of the filtration apparatus [59].
As mentioned above, an anode and a cathode are generally required to construct an electrochemical reaction system. Interestingly, in a recent work, a Janus electrified ceramic membrane was fabricated by sputtering palladium particles on both sides of a ceramic substrate membrane, which could then act as both anode and cathode (Figure 4) [60]. The Janus membrane showed well retained efficiency for electrochemical activation of PMS towards the oxidative degradation of MB with a η value of 95% after 8 repeated runs [60].
The introduction of both photo and electron energy may further boost the efficiency of the persulfate oxidation process. For instance, a g-C3N4/CNT/ceramic membrane was used in the photo-electrocatalysis process, showing a phenol removal efficiency 2.7 and 2.0 times higher than that with photo irradiation and electron energy supply only, respectively [62]. In another work, TiO2-ZnO deposited on stainless steel support was also utilized in the photo-electrocatalysis process. In the presence of PMS, the reaction system could remove atrazine and MB almost completely [61]. After 5 runs, a loss of 15% and 10% in removal efficiency was observed for MB and atrazine, which was mainly due to the adsorption of oxidation intermediates accumulated over time considering the stable morphology and crystal structure of the catalytic membrane [61].

3.2. Performance Stability of Catalytic Membranes Tested by Long-Time Operation

In this section, the performance stability of catalytic membranes evaluated by long-time operation was summarized and analyzed (Table 2), where the related reports were also classified based on the type of AOP and discussed in sub-sections. The factors listed in Table 2 are similar to those in Table 1. However, it should be noted that when the long-time operation method is used, the operation mode must be single pass while the recirculation mode is not applicable. Thus, the column of operation mode is omitted in Table 2.
For comparison of the performance stability results, another dimensionless factor (η′) was determined as follows:
η′ = Rt/R0 × 100%,
where t stands for operation time of the long-time stability test. Rt represents the removal efficiency of the target pollutant obtained at the end of the test, while R0 stands for the value obtained during the early time of the test. Note that R0 was defined in this way rather than as the removal efficiency of the first permeate sample. This was because in some reports, the membrane was saturated in advance by the pollutant solution. In this case, the removal efficiency was limited due to the adsorbed pollutants when the first sample was withdrawn, and it took some time for the removal efficiency to increase and reach a stable value. Similar to η defined in Section 3.1, η′ should also be in the range of 0–100%, and a larger η′ indicates better performance stability within the same operation time.
As summarized in Table 2, for the performance stability tests of these catalytic membranes tested by long-time operation, the employed operation time is in the range of 2.5–240 h, achieving η′ values in the range of 70–100%. A longer operation time and a higher η′ value should indicate better performance stability. Thus, among all the reported cases summarized in Table 2, the Co-TiOx/mixed cellulose ester membrane and the Co-CuOx/PVDF membrane stood out, both of which could remove their target pollutants completely during a long operation time of 100 h and 120 h, respectively. Another two membranes were tested for prolonged periods of time. The Fe-doped graphitic carbon nitride/cellulose acetate membrane was tested for PMS activation towards bisphenol A (BPA) removal for 170 h, and the Co3O4/Ti membrane was tested for electrochemical removal of COD for an even longer time of 240 h. In both cases, the η′ values at the end of the long-time operations surpassed 90%, demonstrating the outstanding performance stabilities of these two catalytic membranes.
In the following sub-sections, the summarized results in Table 2 are further discussed based on the type of advanced oxidation process.
Table 2. Performance stability of catalytic membranes tested by long-time operation.
Table 2. Performance stability of catalytic membranes tested by long-time operation.
AOP TypeSubstrate MembraneCatalystFabrication MethodMembrane TypeFiltration
Mode
Target
Pollutant
C0Flux
(L/(m2·h))
R0Operation Time (h)η′Ref.
PMSPVDFβ-FeOOHimpregnation and in situ mineralizationflat-sheetcross flowRhB10 mg/L4097.999%1299%[63]
PMSPVDFMoS2filtrationflat-sheetdead-endBPA2 mg/L30.8100%690%[64]
PMSPVDFCo-CuOx filtrationflat-sheetdead-endRA5 mg/L357100%120~100%[65]
PMSPVDFZIF-67-derived Co-carbonflash freezingflat-sheetcross-flowTC10 mg/L40097%84 95%[66]
PMS PVDFCoAl-LDHfiltrationflatdead-endRA2.5 mg/L80.391%29 ~100%[11]
PMSPESCo3O4/C@SiO2phase inversionflat-sheetdead-endBPA10 mg/L22995%42 90%[67]
PMScellulose acetateFe-doped graphitic carbon nitridefiltrationflat-sheetdead-endBPA20 mg/L28.66100%170 90%[68]
PMScellulose acetateCo3O4filtrationflat-sheetdead-endRA5 mg/L158.4100%13 90%[69]
PMSmixed cellulose esterCo-TiOxfiltrationflat-sheetdead-endRA5 mg/L130100%100 ~100%[70]
PMSnylon Co@N CNTfiltrationflat-sheetdead-endTC10 mg/L40100%24 90%[71]
PMSnylonFe2O3@CNTfiltrationflat-sheetdead-endTC0.04 mmol/L16.396.1%4889%[22]
PDSnylonrGO/NCNTfiltrationflat-sheetdead-endNSMX0.5 mg/L46.1590%50 79%[72]
PDSnylonNG/rGO/CNTfiltrationflat-sheetdead-endSMX0.5 mg/L46.1599.8%24 99.8%[23]
PMSceramic membranenitrogen doped carbondip-coating and pyrolysisflat-sheetdead-endphenol0.1 mmol/L689.76100%48 100%[73]
PMSceramic membraneMn2O3solid state sintering and calcinationflat-sheetdead-endBPA
EE2
NP
0.1 mg/L
0.1 mg/L
0.1 mg/L
6092.2%
97.6%
97.3%
8 96%
94%
95%
[74]
PMSceramic membraneMn2O3spraying and calcinatingflat-sheet dead-endacetaminophen1 mg/L6095%24 92%[28]
PMScarbon nanofibrous membraneCoOximpregnation and pyrolysisflat-sheetdead-endTC30 mg/L3098%10 90%[75]
H2O2PVDFPrussian bluein situ growthflat-sheetcross-flowMB30 mg/L30099%24 100%[76]
H2O2PVDFFe3O4phase inversionhollow fibercross-flowMB100 mg/L175.897%12 100%[77]
H2O2PVDFBiOIfiltrationflat-sheetdead-endparacetamol10 mg/L12096%100 88%[78]
H2O2PPPrussian blue/GOfiltrationflat-sheetdead-endMB10 mg/L2798%24 100%[79]
H2O2ceramic membraneMoSxhydrothermaltubulardead-endclothianidin10 mg/L2599.9%60 70%[80]
H2O2ceramic membraneFeOClimpregnation and calcinationflat-sheetdead-endpCBA50 μmol/L100100%120 80%[81]
H2O2ceramic membraneCo3O4/MCM-41packingtubulardead-endAO750 mg/L11.2596%40 103%[82]
PhotoPVDFZnIn2S4filtrationflat-sheetdead-endTC100 μg/L84.0690%36 100%[83]
PhotoPVDFAg2CO3@UiO-66-NH2filtrationflat-sheetdead-endMB20 mg/L50100%3.3100%[84]
PhotoPESB doped-
TiO2-SiO2/CoFe2O4
phase inversionflat-sheetcross-flowCOD of biologically treated palm oil mill effluent1000 mg/L38.1100%12100%[85]
Photohydrogel membraneZnCeOx/graphitic carbon nitridecrosslinkingflat-sheetcross-flowMB20 mg/L64095%497%[86]
Photoceramic membraneTiO2atomic layer deposition and calcinationflat-sheetdead-endMB1 mg/L15040%3375%[87]
ElectroTi membrane (anode)
stainless steel mesh (cathode)
Co3O4 (anode)
–(cathode)
hydrothermal and calcinationtubulardead-endphenol5 mmol/L3.2299% (COD removal)24091%[88]
Electrocoal-based carbon membrane (anode)
titanium plate (cathode)
–(anode)
–(cathode)
tubulardead-endBPA50 mg/L43.7397%2490%[89]
Electrocarbon nanofiber membrane (anode)
titanium mesh (cathode)
RuO2/TiO2 (anode)
–(cathode)
hydrothermal and calcinationflat-sheetdead-endBPA
SMX
500 μg/L
500 μg/L
36098%
98%
72 97%
97%
[90]
Electroceramic membrane/Ti mesh (anode)
Ti mesh (cathode)
TiO2-SnO2-Sb (anode)
–(cathode)
adhesive assembleflat-sheetdead-endPCA1.28 mg/L17.485.5%72 97%[91]
Electroceramic membrane/Ti mesh (anode)
titanium mesh (cathode)
TiO2/SnO2-Sb (anode)
–(cathode)
adhesive assembleflat-sheetdead-end2,4-D1 mg/L27862.4%84 100%[58]
ElectroPt foil (anode)
PTFE (cathode)
–(anode)
GO-COOFe2+ (cathode)
filtrationflat-sheetdead-endflorfenicol1 mg/L21.595%72 95%[92]
Electro + O3titanium mesh (anode)
PTFE (cathode)
–(anode)
CNT (cathode)
filtrationflat-sheetdead-endibuprofen2 mg/L14095.19%10 ~100%[93]
Electro + PMSperforated titanium plate (anode)
PTFE (cathode)
–(anode)
zero valence copper-CNT (cathode)
filtrationflat-sheetdead-endCR10.4 mg/L34.6100%2.5 100%[94]
Abbreviations: AO7: Acid orange 7; BPA: Bisphenol A; BPS: Bisphenol S; CNT: Carbon nanotube; COD: Chemical oxygen demand; CR: Congo red; EE2: 17α-ethynylestradiol; LDH: layered double hydroxide; MB: Methylene blue; NP: Nonylphenol; NSMX: N4-acetyl-SMX; PCA: p-chloroaniline; pCBA: Para-chlorobenzoic acid; PDS: Peroxydisulfate; PES: Polyethersulfone; PMS: Peroxymonosulfate; PP: Polypropylene; PTFE: Polytetrafluoroethylene; PVDF: Polyvinylidene fluoride; RA: Ranitidine; rGO: reduced graphene oxide; RhB: Rhodamine B; SMX: Sulfamethoxazole; TC: Tetracycline; ZIF-67: Zeolitic imidazole framework-67; 2,4-D: 2,4-Dichlorophenoxyacetic acid.

3.2.1. Long-Time Operation Performance of Catalytic Membranes Coupled with Persulfate Oxidation

The long-time operation method is commonly used to test the performance stability of catalytic membranes in the persulfate oxidation process. PVDF, the widely employed organic substrate membrane, was loaded with β-FeOOH, and the composite membrane was tested to remove RhB in the presence of PMS, showing high removal efficiency for 12 h. The flux of the membrane under the same transmembrane pressure also remained stable during that time, indicating the high stability of the membrane [63]. Several other catalysts including MoS2 [64], Co-CuOx [65], and ZIF-67-derived Co-carbon [66] were also loaded on PVDF, and the resulting catalytic membranes all possessed high performance stability with η′ values of 90–100%. The CoAl-LDH/PVDF membrane mentioned in Section 3.1.1 was also tested in the long-time operation by changing the operation mode from recirculation to single pass, and the removal efficiency of RA remained stable during the 29 h operation. Interestingly, the flux of the membrane was observed to increase slightly in the first 2 h, which was possibly due to swelling of the pores caused by hydration of the membrane [11]. Organic substrate membranes made from other materials including PES [67], cellulose acetate [68,69], mixed cellulose ester [70] and nylon [22,71] were also investigated. After loaded with various Co- or Fe-containing catalysts, the fabricated catalytic membranes generally showed high performance stability with η′ values above 90%, and limited metal leaching is generally detected in these reports.
In a recent work, a metal-free catalyst, i.e., rGO/N-doped CNT (NCNT), was loaded on a nylon substrate membrane, which was then employed for the degradation of N4-acetyl-SMX (NSMX), the main refractory metabolite of SMX. The membrane exhibited over 90% NSMX removal efficiency within the first 10 h. However, at the end of the 50 h continuous filtration, the removal efficiency decreased to 70.7%, corresponding to a η′ value of 79% [72]. XPS measurements revealed the accumulation of acid oxygen-containing groups, especially phenolic and carboxylic groups, in the used membrane, which hindered PDS activation with electron-withdrawing effects at the reactive sites. In addition, the partial transformation of graphitic and pyridinic N to pyrrolic N as well as the oxidation of structural defects also contributed to the decreased removal efficiency [72]. Another metal-free catalytic membrane (NG/rGO/CNT/nylon), which was mentioned in Section 3.1.1, was fabricated by the same research group [23]. The membrane showed stable SMX removal in the 24 h long-time operation when deionized water or tap water was used as the background. However, when surface water or effluents of a municipal wastewater treatment plant were utilized, declined removal efficiency was observed. This was attributed to membrane fouling as evidenced from the formation of gel-like cake layers after filtration. In addition, the decrease in the degree of carbon defects also contributed to the decreased removal efficiency [23].
Catalytic membranes constructed of inorganic membranes including ceramic membrane and carbon membrane were also investigated. Nitrogen-doped carbon supported on ceramic membrane was investigated for phenol degradation in a long-time period of 48 h. Specifically, real surface water was used as the background, and phenol removal remained 100% during the whole operation time [73]. In another work, Mn2O3-containing ceramic membrane was prepared by solid state sintering and calcination [74]. The membrane could efficiently remove three organic pollutants including BPA, 17α-ethynylestradiol (EE2), and nonylphenol (NP), showing similar removal efficiencies (92.2–97.6%) and η′ values (94–96%). Another Mn2O3-containing ceramic membrane mentioned in Section 3.1.1 also showed slightly decreased removal efficiency from 95% to 87% (η′ = 92%) in the 24 h long-time operation [28]. Meanwhile, TOC removal during that time was also recorded, which declined from 64% to 58% [28]. CoOx/carbon nanofibrous membrane as a gravity-driven membrane was constructed in another report. The long-term stability of the membrane for TC removal was investigated using consecutive gravity filtration experiments, showing a constantly high removal efficiency over 90% and stable permeability in 10 h [75].

3.2.2. Long-Time Operation Performance of Catalytic Membranes Coupled with H2O2 Oxidation

Two iron-based catalysts, Prussian blue [76] and Fe3O4 [77], were loaded on PVDF substrate membrane and then used for the oxidative degradation of MB in the Fenton process. Both membranes showed stable removal efficiencies in the 24 h and 12 h operation time, respectively, with constant [76] or slightly decreased water flux [77]. In another report, the loaded catalyst was changed to BiOI, and the synthesized membrane also showed stable removal efficiency in the 24 h operation period [78]. Nevertheless, when the operation time was extended to 100 h, the removal efficiency declined with a η′ value of 88%. Leaching of Bi and I was measured throughout the whole operation process. Although the leached concentration of I was much higher than that of Bi, the role of I in the catalytic process was found to be negligible [78]. Instead of PVDF, PP membrane as another organic polymer membrane was utilized as the substrate to load Prussian blue/GO [79], which could stably remove MB for 24 h with almost constant flux.
To modify ceramic membranes for the Fenton oxidation process, MoSx [80] and FeOCl [81] were loaded, and the degradation of clothianidin and para-chlorobenzoic acid (pCBA) was investigated, respectively. During the 24 h operation time in the former case, the removal efficiency could maintain above 90%. When extending the operation time to 60 h, the removal efficiency further declined and approached 70% [80]. In the latter case, the removal efficiency could maintain almost complete for 24 h, but finally reached 80% after another 96 h [81]. Different from the BiOI catalyst mentioned above, for this FeOCl catalyst, the reason for the decreased catalytic activity was attributed to the leaching of Cl rather than Fe. It was noted that the Fenton reaction was initiated by the binding of H2O2 to the Fe site which was coordinated with a labile Cl ligand. Since H2O2 possessed a higher affinity to the original Fe site bound to Cl compared to the Fe site formed after each reaction cycle, the gradual loss of Cl resulted in gradually deteriorated removal efficiency [81].
In a recent report, to test the applicability of the Co3O4/MCM-41 catalyst packed in the lumen of tubular ceramic membrane, humic acid (HA) as a typical natural organic matter was fed together with acid orange 7 (AO7), which was the target organic pollutant in the Fenton oxidation process [82]. During the 40 h operation time, water flux of the integrated membrane gradually declined by 10%, indicating the fouling effect of HA. Interestingly, the removal efficiency of AO7, which was 96% first, was gradually increased to more than 99% at the end of 40 h [82], indicating an unusual η′ higher than 100%. The reason behind this enhanced removal efficiency was not discussed in the original work. It was inferred that HA gradually accumulated on the membrane and led to decreased pore size, which probably benefitted the rejection of AO7 and contributed to the enhanced removal efficiency.

3.2.3. Long-Time Operation Performance of Catalytic Membranes Coupled with Photocatalysis

Using PVDF membrane as the substrate, ZnIn2S4 [83] and Ag2CO3@UiO-66-NH2 [84] were respectively loaded by filtration and tested for the photocatalytic degradation of TC and MB. Stable removal efficiencies were achieved in both cases. In another work, B-doped TiO2-SiO2/CoFe2O4 were loaded in PES substrate membrane by phase inversion [85]. Specifically, different from the single pollutants generally investigated in other reports, the COD of biologically treated palm oil mill effluent was treated by this photocatalysis membrane filtration system. During the operation time of 12 h, complete COD removal could be stably achieved, although a reduction in flux was observed simultaneously [85].
Different from the commonly employed organic polymer substrate membrane, a self-standing hydrogel membrane made from sodium alginate and porous carbon nitride with uniformly distributed ZnCeOx nanoparticles was prepared in a recent report [86]. During the operation time of 4 h, the integrated hydrogel could stably reject MB both with and without visible light irradiation. However, under visible light conditions, the water flux remained stable after a drop by 21%, while it dropped dramatically by 53% in the dark. This clearly demonstrated the indispensable contribution of photocatalytic degradation during the filtration process [86].
A ceramic substrate-based catalytic membrane was constructed after the atomic layer deposition of TiO2 in another report [87]. The as-synthesized membrane was tested for the continuous degradation of MB under UV radiation during a time period of 33 h. The removal efficiency of MB maintained stable at around 40% during the first 10 h. After that, it declined slowly and stabilized again at around 30%. The authors noted that the decrease in MB removal efficiency was unexplained and possibly within system error considering that other recorded variables including temperature, pressure, conductivity, and flow rate were all constant throughout the whole operation time [87].

3.2.4. Long-Time Operation Performance of Catalytic Membranes Coupled with Electrocatalysis

Ti membrane as a conductive substrate was modified with Co3O4 by hydrothermal treatment and calcination, which was then employed for the removal of phenol containing wastewater [88]. The removal efficiency of COD maintained stable during a long operation time of 240 h, and stable hydrogen production could be simultaneously achieved with the reaction system. When further extending the operation time, a clear decline trend was observed, showing a η′ value of 60% at 2400 h. X-ray diffraction (XRD) and scanning electron microscopy (SEM) tests revealed the detachment and collapse of Co3O4 particles on the membrane surface, which was responsible for the deteriorated removal efficiency [88].
Carbon-based membranes as another type of conductive membrane have been investigated as well. Coal as a cheap and abundant precursor was fabricated into a carbon membrane, which was then directly used as the anode for the electrochemical oxidation of BPA, showing a η′ value of 90% after 24 h [89]. In another report, a carbon nanofiber membrane was modified with RuO2/TiO2, and the integrated membrane could stably remove BPA and SMX during an operation time of 72 h [90].
With continuous supply of oxygen, the integrated membrane loaded with TiO2-SnO2-Sb was reported to stably remove p-chloroaniline (PCA) [91] and 2,4-dichlorophenoxyacetic acid [58]. In another report, PTFE membrane modified with GO-COOFe2+ was employed for the electrochemical oxidation of florfenicol, which could stably remove florfenicol during 72 h. The modified membrane served as the cathode, promoting the electro-Fenton process through the production of H2O2 from oxygen, while the contribution of ·OH generated on the anode was negligible [92].
Additional oxidants, such as O3 and persulfates, can be added in the electrocatalysis process to boost the removal efficiency. In a recent work, a CNT modified PTFE membrane served as the cathode, which was responsible for the decomposition of both in situ-generated H2O2 and additionally added O3 [93]. The reaction system could remove ibuprofen and another four organic pollutants including carbamazepine (CBZ), caffeine, triclosan, and acetaminophen with removal efficiencies over 95% and η′ values approaching 100% [93]. Another PTFE membrane modified with zero valence copper and CNT was also employed as the cathode for the electrochemical activation of PMS, maintaining complete CR removal during 2.5 h continuous filtration [94].
Although the performance stability of catalytic membranes is an important issue, information on this is missing for some reported membranes with inspiring catalytic activities. Thus, the practical applications of these membranes may be doubtful. Sometimes the stability was tested by measuring the change in pure water flux. Although pure water flux is an important parameter for membranes, it could not directly reflex the stability of membranes in the catalytic process.
Summarizing the discussions above, it can be found that the performance of catalytic membranes generally deteriorates upon multi-cycle use or long-time operation. In some cases, the removal efficiency of the target pollutant maintains high. However, stable removal efficiencies may not always indicate high performance stability of the membranes.
In the recirculation operation mode, stable removal efficiency may be obtained at the end of the multi-cycle degradation runs, but a decrease in the kinetic constant may take place. In the single pass operation mode, the removal efficiency may also remain stable during the long-time operation, but the increased TMP at constant flow rate, the decreased flow rate at constant TMP, or the increased cell voltage at constant current density may indicate the performance reduction of the membranes, leading to increased operation cost of the filtration process. Without measuring changes in these factors, the stability of catalytic membranes may not be clearly revealed.
In addition, most of the stability tests were conducted in pure water without co-existing ions and natural organic matter. For practical applications, the presence of these species is inevitable, which may lead to membrane fouling and strongly weaken the performance stability of these catalytic membranes.
The removal efficiency of TOC or COD is generally measured for fresh catalytic membranes only. The ability of catalytic membranes for the removal of TOC or COD may also decline upon repeated use, which has been paid little attention.
In summary, from a practical viewpoint, regeneration of the used catalytic membranes is required sooner or later to restore their ability towards efficient and sustained degradation of organic pollutants, which is discussed in the following section.

4. Regeneration Performance of Catalytic Membranes Coupled with Advanced Oxidation Process

As discussed above in Section 3, a series of reasons may lead to decreased activity of the membranes, including adsorption of organic pollutants and its oxidation intermediates, changes in chemical properties of the membrane such as valence states or surface functionalities, as well as leaching of metal species. Faced with these inactivation reasons, various regeneration methods have been developed. To remove the adsorbed organic species, washing the used membrane with water and other solvents is a widely used approach. A heat treatment procedure or an addition oxidation step can also help to remove the adsorbed organic species. To restore the chemical properties of the carbon-based membranes, heat treatment may help to remove the oxygen functionalities on their surface, which were produced during the degradation process. To restore the valence states of a metal-based membrane, electrochemical or chemical reduction of the used membrane may make a contribution. Note that none of these methods can solve the leaching problem, and it is important to carefully regulate the level of metal leaching through optimization of the membrane structure in order to extent the lifespan of metal-based catalytic membranes.
For comparison of the regeneration performance, the dimensionless factor (η″) was determined as follows:
η″ = Rr/R0 × 100%,
where Rr represents the removal efficiency of the target pollutant obtained with the regenerated membrane, while R0 stands for the value obtained with the fresh one.
The recent works concerning the regeneration of catalytic membranes using solvent washing, heat treatment, and other methods are summarized in Table 3, Table 4, and Table 5, respectively. For the solvent washing method, the CoCu-LDH/polyethylene glycol (PEG)/PVDF membrane stood out, which provided stable SMX removal with a η″ value of 100% in 10 regeneration cycles using ethanol as the washing solvent. The two membrane electrodes, carbon black/polyaniline/carbon felt and CNT-doping polypyrrole/PVDF, which were regenerated by washing with water, showed stable performance with η″ values of 95% after 15 and 20 cycles, respectively, indicating their promising regeneration properties. The employed number of cycles for the heat treatment regeneration method (1–5 cycles) was much smaller compared to that of solvent washing (3–20 cycles). Although η″ values approaching 100% were reported for several catalytic membranes, the heat treatment was conducted only once for them, and the trend in more heat treatment cycles remains unknown. For the Mo/ceramic membrane and the CoFe2O4/diatomite/stainless steel mesh, 3 and 5 heat treatment regeneration cycles in air were conducted, respectively, but the achieved η″ values deviated from 100%, reaching 94% and 87%. When other regeneration methods were employed, the number of cycles was in the range of 1–8 cycles. The three membranes of Mn2O3/ceramic membrane, TiO2/Poly(sodium-p-styrenesulfonate) (PSS)/PVDF, and ZIF-67/Ni-foam, which were regenerated by advanced oxidation, all showed outstanding η″ values approaching 100% after 5 cycles. The CoFe2O4/ceramic membrane was also promising, showing a η″ value of 95% after 8 regeneration cycles by advanced oxidation.
In the following sub-sections, the summarized results in Table 3, Table 4 and Table 5 will be separately discussed based on the type of regeneration method.

4.1. Regeneration Performance of Catalytic Membranes by Washing

Washing is probably the most widely employed method for membrane regeneration, during which the adsorbed organic species are washed away with the solvents, and thus, the active sites are restored. The washing process can be conducted in many approaches listed as follows (Figure 5):
(a)
The used membrane remains in the filtration apparatus, and filtration with the washing solvent is conducted. This method is applicable to most of the catalytic membranes.
(b)
The used membrane remains in the filtration apparatus, and back washing with the washing solvent is conducted. This method is not applicable to the catalytic membranes formed by filtration.
(c)
The used membrane is removed from the filtration apparatus and then immersed in the solvent, and sonication, stirring or another mechanical force is introduced to facilitate the desorption/detachment of organic pollutants on the membrane. This method is not applicable to the catalytic membranes formed by filtration and those with weak mechanical strength.
(d)
The used membrane is removed from the filtration apparatus, and the catalytic particles on the membrane were removed from the substrate membrane. After that, the separated catalytic particles were washed with the solvent with the assistance of sonication or stirring. After washing, the regenerated catalytic particles were reloaded on the substrate membrane. This method is only applicable to the catalytic membranes formed by filtration.
Various washing solvents have been utilized, and water is the most extensively investigated one due to its green and economical nature. The recent works concerning regeneration of catalytic membranes using solvent washing are summarized in Table 3. An additional column showing the employed washing solvent is provided in the table. As shown in Table 3, water has been widely utilized for membrane regeneration in PMS oxidation [95,96,97,98,99,100], H2O2 oxidation [101,102,103,104], O3 oxidation [35,105], photocatalytic oxidation [106], and electrochemical oxidation [107,108,109,110,111] processes. After washing with water, the adsorbed/attached pollutants can be removed, resulting in restored catalytic performance. In these reports, high η″ values in the range of 88–100% were obtained, indicating the efficiency of this regeneration method.
However, in most of these reports, the performance of the membrane cannot be completely restored. On the one hand, although the adsorbed/attached pollutants can be removed by washing with water, other reasons may also lead to decreased catalytic activity of the membrane. For example, when the NH2-MIL-88B(Fe)/ceramic membrane was investigated, the catalyst loss was found to be 12.5 mg after the first run, and the value was in the range of 2–5 mg during the subsequent runs [107]. This catalyst loss partly resulted in the decreased activity of the membrane, which could not be recovered by washing. As another example, the ceramic tube membrane loaded with Cu-UiO-66 showed high regeneration ability with a η″ value of 90% after 5 regeneration cycles by washing with water, and the slightly decreased activity was attributed to the leaching of Cu [104]. However, when Cu was replaced by Mn, the Mn-UiO-66/ceramic tube membrane showed dramatically deteriorated phenol removal after 5 cycles with a η″ value of only 11%. This was ascribed to the low stability of the Mn-UiO-66 catalyst, whose decomposition resulted in substantial leaching of Mn and sharply declined removal efficiency [104].
On the other hand, some organic pollutants may have possessed low solubility in water and could not be completely removed after washing, which consumed reactive species in subsequent cycles and led to decreased removal efficiency. Consequently, other washing agents were also investigated for membrane regeneration, including organic solvents as well as acidic and alkaline solutions.
Table 3. Regeneration performance of catalytic membranes by washing.
Table 3. Regeneration performance of catalytic membranes by washing.
AOP TypeSubstrate MembraneCatalystFabrication MethodMembrane TypeFiltration
Mode
Operation ModeTarget
Pollutant
C0Flux
(L/(m2·h))
R0Washing SolventNumber of Washing Cyclesη″Ref.
PMSceramic membraneMnO2ball-milling and calcinationflat-sheetcross-flowsingle pass4-hydroxylbenzoic acid80 mg/L19.50100%water691%[95]
PMSceramic membraneCoFe2O4impregnation and calcinationtubulardead-endsingle passMB25.0 mg·L−115098%water492%[96]
PMSceramic membraneFe-doped CoTiO3impregnation and calcinationflat-sheetdead-endsingle passnimesulide10 mg/L30096.3%water588%[97]
PMSceramic membraneCoOx
granular activated carbon
dipcoating and calcination;
filling the channels
flat-sheetdead-endsingle passBPA10 mg/L3595%water595%[98]
PMSTi membraneCo3O4-Bi2O3impregnation and calcinationflat-sheetdead-endsingle passMB20 μmol/L76.898.7%water4~100%[99]
PMSTi membraneZIF-67electro-deposition and calcinationflat-sheetdead-endsingle passSMX0.04 mmol/L56.996.3%water593%[100]
H2O2bacterial cellulose membranetetra-amino cobalt (II) phthalocyanineimpregnation and in situ reactionflat-sheetdead-endrecirculationred X-3B100 µmol/L1146.5095%water597%[101]
H2O2ceramic membranePDA-β-FeOOHimpregnation and in situ reactiontubularcross-flowsingle passMO2 mg/L3898%water6~100%[102]
H2O2CNT membraneFe0wet-spinning and calcinationhollow fiberdead-endsingle passBPA10 mg/L545.997.8%water597%[103]
H2O2ceramic
membrane
Cu-UiO-66
Mn-UiO-66
solvothermal and
impregnation
tubulardead-endsingle passphenol100 mg/L3.18100%
90%
water590%
11%
[104]
O3ceramic membraneMn/FeOxco-precipitation and calcinationflat-sheetdead-endsingle passtrimethoprim5 mg/L1098.6%water591%[105]
O3ceramic membraneN-rGOpneumatic methodtubularcross-flowsingle passbenzotriazole0.084 mol/L660.86100%water195%[35]
Photoceramic membraneTiO2dip-coating and calcinationtubularcross-flowrecirculationOTC5 mg/L96099%water3~100%[106]
Electroceramic
membrane
NH2-MIL-88B(Fe)hydrothermalflat-sheetdead-endrecirculationnaproxen0.060 mmol/L15.997%water589%[107]
Electrocoal-based carbon membrane (anode)
the titanium plate (cathode)
polyaniline (anode)
–(cathode)
electrochemical polymerization depositiontubulardead-endsingle passphenol50 mg/L77.4299%water1097%[108]
Electrocoal-based carbon membrane (anode)
the titanium plate (cathode)
polyaniline (anode)
–(cathode)
electrochemical polymerization depositiontubulardead-endsingle passphenol50 mg/L99.3%water696%[109]
Electrotitanium (anode)
carbon felt (cathode)
RuO2-IrO2 (anode)
carbon black and polyaniline (cathode)
phase inversionflat-sheetdead-endrecirculationTC50 mg/L3857.1492.9%water1595%[110]
Electro + PMStitanium plate (anode)
PVDF (cathode)
–(anode)
CNT doping polypyrrole (cathode)
phase
inversion
flat-sheetcross-flowsingle passCBZ2 mg/L142.595%water2095%[111]
PMSPTFEMn3O4/CNNSfiltrationflat-sheetdead-endrecirculation4-chlorophenol0.1 mM15.9290%ethanol698%[112]
PMSPTFENiCo@NCNTfiltrationflat-sheetdead-endsingle passibuprofen20 mg/L305.73100%ethanol395%[113]
PMSPVDFFe-Co@NCNTphase inversionflat-sheetcross-flowsingle passBPA30 mg/L45.5100%ethanol1060%[114]
PMSPESCoCu-LDH/PEG/calotropis gigantean fiberfiltrationflat-sheetdead-endsingle passSMX10 mg/L124.7893%ethanol10100%[115]
PMSnylonMnOOHfiltrationflat-sheetdead-endrecirculation2,4-Dichlorophenol25 mg/L238.897.9%ethanol499%[116]
Electrocoal-based carbon membrane (anode)
Ti plate (cathode)
–(anode)
–(cathode)
tubulardead-endsingle passRhB250 mg/L88.7100%ethanol6~100%[117]
Electrocoal-based carbon membrane (anode)
Ti plate (cathode)
CuO (anode)
–(cathode)
electrodepositiontubulardead-endsingle passRhB300 mg/L66.5199.9%ethanol4~100%[118]
PMSAg-La0.8Ca0.2Fe0.94O3−δphase inversion and calcinationhollow fiberdead-endrecirculationMB10 mg/L2228.5790%H2SO4578%[119]
H2O2ceramic membraneFeOClimpregnation and calcinationflat-sheetdead-endsingle passpCBA50 μmol/L100100%HCl480%[81]
PhotoPVDFPDA/BiOCl0.875Br0.125pneumatic methodflat-sheetdead-endrecirculationroxarsone17.5 mg/L380.4100%NaOH479%[120]
Photo +H2O2ceramic membraneFeOClcross-linkingflat-sheetdead-endsingle passnitrobenzene10 μmol/L69.2100%NaOH5100%[121]
ElectroTi sheet (anode)
PTFE (cathode)

(anode)
CNT@MIL-101(Fe) (cathode)
filtrationflat-sheetdead-endrecirculationTC17.8 mg/L25.3594.3%NaOH488%[122]
Electro + PMSPTFECNT-Fe3O4filtration and hydrothermalflat-sheetdead-emdrecirculationroxarsone1.5 mg/L91.8%NaOH490%[123]
Abbreviations: BP-3: Benzophenone-3; BPA: Bisphenol A; CBZ: Carbamazepine; CNT: Carbon nanotube; LDH: layered double hydroxide; MB: Methylene blue; MO: Methyl orange; NCNT: N-doped CNT; OTC: Oxytetracycline; pCBA: para-chlorobenzoic acid; PDA: Polydopamine; PES: Polyethersulfone; PMS: Peroxymonosulfate; PTFE: Polytetrafluoroethylene; PVDF: Polyvinylidene fluoride; rGO: reduced graphene oxide; RhB: Rhodamine B; SDZ: Sulfadiazine; SMX: Sulfamethoxazole; TC: Tetracycline; ZIF-67: Zeolitic imidazole framework-67.
Ethanol as a cheap and green organic solvent has been widely investigated [112,113,114,115,116,117,118]. In some cases, the washing procedure was conducted under solvothermal conditions. For example, for the regeneration of Mn3O4 nanodots-g-C3N4 nanosheets (Mn3O4/CNNS), the used membrane was washed with ethanol and water and then used for the next run, and the η″ approached 89% after 5 cycles. After that, the used membrane was further treated with ethanol at 150 °C for 3 h, obtaining an increased η″ value of 98% [112]. This may be ascribed to the enhanced solubility of organic pollutants in ethanol under the increased treatment temperature.
Acid and alkaline solutions have long been used for the chemical cleaning of fouled membranes. The change in pH may alter the charge status of both the membrane surface and the adsorbed pollutants, enhancing desorption of the adsorbed pollutants and, thus, the regeneration efficiency. Various acids including H2SO4, HCl, and H3PO4 have been employed. H2SO4 was used for the regeneration of perovskite oxide membrane, achieving a η″ value of 78% after 5 cycles. The declined catalytic activity was mainly ascribed to leaching of iron, which changed the surface composition of perovskite and caused a loss in activity over time [119]. For the regeneration of FeOCl/ceramic membrane, HCl was utilized [81]. As mentioned above in Section 3.1.1, the decreased activity of the membrane was partly ascribed to the leaching of Cl. It was reported that the activity of the used membrane could be effectively restored by HCl treatment and subsequent heat treatment at 220 °C. HCl treatment helped to remove the pollutants on the used membrane. More importantly, the regeneration procedure led to Cl intercalation and recovery of the Fe-Cl bonds. The heat treatment step was crucial to restore the Fe-Cl bonds, as evidenced by the low regeneration efficiency with HCl treatment only [81]. For alkaline treatment, NaOH is commonly employed. For example, NaOH was used to regenerate the PDA/BiOCl0.875Br0.125/PVDF membrane after the treatment of roxarsone, which was an arsenic-containing organic pollutant [120]. During the photocatalytic process, the released inorganic arsenic could be converted to As (V) and then immobilized on the membrane surface. Upon NaOH treatment, OH could replace the adsorbed As (V) anions, while Na+ helped to weaken the electrostatic interaction between PDA and As (V) anions [120]. NaOH was also used to regenerate other catalytic membranes in several photocatalytic [121] and electrochemical catalytic processes [122,123]. To further improve the regeneration performance, acid and alkaline treatment can be used in combination, such as in the case of α-FeOOH/ceramic membrane where NaOH and H3PO4 washing was performed successively [124].
It should be noted that when organic solvents or acid/alkaline solutions were used for membrane cleaning, a final washing step with water is generally required to remove the residue detergent, avoiding the possible interference in subsequent operations. In addition, although the utilization of these agents could lead to improved regeneration performance, the increased cost and potential environmental pollution should also be taken into consideration.
Solvent washing can help to remove the adsorbed pollutants to some extent. However, even after solvent washing with different solvents, some of the pollutants may still remain on the membranes. Another method to remove the adsorbed pollutants in the heat treatment approach, which is discussed in the following section.

4.2. Regeneration Performance of Catalytic Membranes by Heat Treatment

During heat treatment, the adsorbed species volatilize or decompose into gaseous products under the applied high temperature and leave the occupied active sites on the membrane. The recent works concerning regeneration of catalytic membranes using heat treatment are summarized in Table 4. The employed treatment temperature and atmosphere were provided in two additional columns in the table. As shown in Table 4, different types of atmospheres, including nitrogen [125], argon [113], and air [126,127,128,129,130,131], have been applied for the thermal regeneration of catalytic membranes.
Table 4. Regeneration performance of catalytic membranes by heat treatment.
Table 4. Regeneration performance of catalytic membranes by heat treatment.
AOP TypeSubstrate MembraneCatalystFabrication MethodMembrane TypeFiltration
Mode
Operation ModeTarget
Pollutant
C0Flux
(L/(m2·h))
R0Treatment
Temperature/°C
AtmosphereNumber of Heat Treatment Cyclesη″Ref.
PMSPTFENiCo@NCNTfiltrationflat-sheetdead-endsingle passibuprofen20 mg/L305.73100%350Ar1100%[113]
PDSCNTwet-spinning and pyrolyzationhollow fiberdead-endsingle passphenol10 mg/L97%800N21100%[125]
PMSceramic membraneCNT@nitrogen-doped carbonfiltrationflat-sheetdead-endsingle passSMX20 mg/L6.865%250air1100%[126]
PMSnylonnitrogen-doped rGO/CNTfiltrationflat-sheetdead-endsingle pass4-chlorophenol0.16 mM72.55100%90air194%[127]
PMSPVDFCoxFe3−xO4filtrationflat-sheetdead-endsingle passBPA10 mg/L10099.9%100air1~100%[128]
PMSsteel mesh/non-woven fabricLaFexCo1−xO3−λ/SiO2filtartionflat-sheetdead-endrecirculationTC20 mg/L1064.599%450air1~100%[129]
PMSstainless steel meshCoFe2O4/diatomitefiltrationflat-sheetdead-endrecirculationRhB50 mg/L98.6899.9%200air587%[130]
Airceramic membraneMohydrothermal and calcinationflat-sheetdead-endsingle passsafranine O10 mg/L32496%300air394%[131]
Abbreviations: BPA: Bisphenol A; CNT: Carbon nanotube; NCNT: N-doped CNT; PDS: Peroxydisulfate; PMS: Peroxymonosulfate; PTFE: Polytetrafluoroethylene; PVDF: Polyvinylidene fluoride; rGO: reduced graphene oxide; RhB: Rhodamine B; SMX: Sulfamethoxazole; TC: Tetracycline.
When conducting thermal regeneration, the employed atmosphere and heating temperature should be carefully designed considering the properties of the catalytic membrane. For carbon-based membranes, the regeneration process should be generally conducted under inert atmosphere to avoid the oxidation of the membrane. For example, the regeneration of NiCo@NCNT and CNT membranes was conducted under argon [113] and nitrogen [125] atmospheres, respectively. The air atmosphere can be applied as well, but the employed temperature should be much lower. For example, the regeneration of CNT@nitrogen-doped carbon and nitrogen-doped rGO/CNT membranes was performed at 250 °C [126] and 90 °C [127], respectively, lower than the temperatures of 350 °C and 800 °C applied in the former two cases. To avoid the oxidation of the substrate membrane, the loaded catalyst could be separated first and then treated by heat treatment. After heat treatment, the regenerated catalytic particles could be reloaded on the substrate membrane. For example, the LaFexCo1−xO3−λ/SiO2 catalyst could withstand the high temperature treatment, but the substrate membrane of steel mesh/non-woven fabric could not. Therefore, the functional layer formed by the catalyst was separated and calcined in air at 450 °C, and then, the catalytic membrane was reformed by filtration, achieving a high η″ value close to 100% [129]. In another case where CoFe2O4/diatomite was supported on stainless steel mesh, a similar separation, calcination, and reforming procedure was applied [130].
Although the thermal regeneration method may be highly efficient, the energy cost may be a potential concern. In addition, a higher treatment temperature may lead to better regeneration performance, but the cost is also considerably increased. Consequently, rather than conducting heat treatment after each catalytic run, it can be performed periodically in the whole lifespan of the membrane to make a compromise between regeneration cost and removal efficiency. For example, when the CNT membrane was employed in combination with PDS for phenol degradation, it was used for five times without regeneration, after which thermal regeneration was conducted [125]. Similarly, the CNT@nitrogen-doped carbon/ceramic membrane was used for three times without regeneration followed by a calcination step in air [126]. The thermal regeneration method can also be used in combination with the solvent washing approach. For instance, the nitrogen-doped rGO/CNT membrane was regenerated by heating at 90 °C and subsequent washing with NaOH [127]. Among the summarized reports, the heating temperature in this work was the lowest. This may be attributed to the assistance of NaOH washing, which shared some responsibility. In another work [113], for the regeneration of the NiCo@NCNT membrane, the thermal method and solvent washing method were also used in combination but operated in a different way. After each of the first three runs, solvent washing with ethanol was conducted, while heat treatment was performed after the fourth run [113]. Twelve runs were performed in total, with nine times of solvent washing and three times of thermal regeneration, achieving an acceptable η″ value of 78% in the end.
Besides solvent washing and heat treatment, other methods have also been employed for membrane regeneration, which are discussed in the following section.

4.3. Regeneration Performance of Catalytic Membranes by Other Methods

The recent works concerning regeneration of catalytic membranes by other methods are summarized in Table 5, where the employed regeneration method is shown in a separated column.
Table 5. Regeneration performance of catalytic membranes by other methods.
Table 5. Regeneration performance of catalytic membranes by other methods.
AOP TypeSubstrate MembraneCatalystFabrication MethodMembrane TypeFiltration
Mode
Operation ModeTarget
Pollutant
C0Flux
(L/(m2·h))
R0Regeneration MethodNumber of Regeneration Cyclesη″Ref.
PMSceramic membraneCoFe2O4hydrothermalflat-sheetdead-endsingle passofloxacin40 μmol/L98100%PMS895%[132]
PMSceramic membraneCoFe2O4impregnation and calcinationflat-sheetdead-end
single passSMX10 mg/L23650%PMS395%[133]
PMS
ceramic membraneCo3O4impregnation and calcinationflat-sheetdead-endsingle passSMX0.08 mmol/L23646%PMS354%[134]
PMSceramic membraneCo3O4in situ reaction and calcinationflat-sheetdead-end
recirculationSMX10 mg/L211.6875%PMS195%[135]
O3ceramic membraneMn2O3impregnation and calcinationflat-sheetdead-end
single passSMX20 mg/L3581.3%O35~100%[136]
PhotoPVDFTiO2/PSSlayer-by-layer depositionflat-sheetcross-flowrecirculationLanasol Blue 3R30 mg/L30091.42%photo5~100% [137]
PMSNi-foamZIF-67electrodeposion and calcinationflat-sheetdead-endrecirculationBPA0.044 mmol/L12.68100%photo5~100%[138]
Electroporous Ti plate (anode)
porous Ti plate (cathode)
blue TiO2 (anode)
blue TiO2 (cathode)
electrooxidation, calcination and electroreductionflat-sheetdead-endrecirculationCBZ10 mg/L11464.9795.3%electrochemical reduction499%[139]
ElectroTi membrane (anode)
Ti mesh (cathode)
blue TiO2 (anode)
–(cathode)
electrooxidation, calcination and electroreductionflat-sheetdead-endsingle passMB10 mg/L99.5%electrochemical reduction2~100%[140]
Electronceramic membrane (anode)
stainless steel mesh (cathode)
CNT (anode)
–(cathode)
filtration and pyrolyzationtubularcross-flowsingle passTOC of water from Lingshui Reservoir3.281 mg/L347.8490%electrochemical back washing495%[141]
H2O2PVDFFe/PDA/PEIdeposition, filtration and reductionflat-sheetcross-flowsingle passRhB10 mg/L40095%chemical reduciton with NaBH41~100%[142]
Abbreviations: BPA: Bisphenol A; CBZ: Carbamazepine; CNT: Carbon nanotube; MB: Methylene blue; PDA: Polydopamine PMS: Peroxymonosulfate; PSS: polystyrene; PVDF: Polyvinylidene fluoride; RhB: Rhodamine B; SMX: Sulfamethoxazole; TOC: Total organic carbon; ZIF-67: Zeolitic imidazole framework-67.
Since AOP can help to remove the pollutants in the feed water, it can also be used to remove the pollutants on the used catalytic membranes. In the regeneration process, no additional catalyst is required because the membrane itself can act as the catalyst, and the oxidant or external power is provided in the absence of pollutants in feed water. Generally, the same oxidant or external power as that employed for the treatment of polluted water is used for membrane regeneration. This is because the active species in different AOPs may differ from each other, and the used membrane after a specific AOP may not act as an efficient catalyst during regeneration by another AOP. For instance, after being used in the PMS oxidation process, several membranes loaded with Co-based catalytic species were regenerated by PMS oxidation [132,133,134,135,143]. After PMS treatment, the accumulated pollutants on the membranes were further degraded and removed, leading to high η″ values over 90%. The relatively low η″ reported in Ref. [134] was ascribed to the co-existence of humic acid in feed water, which could not be efficiently degraded in the catalytic oxidation process. Similarly, the Mn2O3/ceramic membrane was employed in the catalytic ozonation process and regenerated by ozone oxidation [136]. The FeOCl/PVDF membrane was tested for Fenton oxidation and regenerated by Fenton oxidation [144]. TiO2/PSS/PVDF was utilized in the photocatalytic process and regenerated by photo treatment [137]. Interestingly, after ZIF-67/Ni-foam was tested for PMS activation, its regeneration was performed not by PMS oxidation but by photo irradiation, achieving a high η″ close to 100% after 5 regeneration cycles [138]. This may be ascribed to the specific feature ofZIF-67, which could act as the catalyst not only for PMS activation but also in the photocatalytic process.
Several catalytic membranes, after being used for electrochemical oxidation, were regenerated by electrochemical treatment. However, different from the cases in other AOPs discussed above, the mechanism during the regeneration process is not the same as that in the degradation process. For example, the Ti-based membrane anodes loaded with blue TiO2 were regenerated by electrochemical reduction, during which the active Ti3+ sites were restored, leading to recovered catalytic activities [139,140]. In another two cases, both the ZnO/coal-based carbon membrane [145] and the CNT/ceramic membrane [141] were used as anodes in the electrochemical oxidation process, but their regeneration was performed by electrochemical back washing where they acted as cathodes.
Chemical reduction with reduction agents such as NaBH4 has also been employed for metal-based membranes. For example, after being used in the Fenton process, the membrane of Fe/PDA/PEI/PVDF was regenerated by NaBH4 (Figure 6) [142]. Similarly, the FexOy/PVDF membrane used for PDS oxidation was also regenerated by NaBH4 treatment [146]. In both cases, chemical reduction with NaBH4 helped to reduce the high-valent iron species, resulting in restored catalytic activity.
Summarizing the above discussions, the regeneration properties of catalytic membranes by various methods have been investigated in many references, but there are still some problems. As mentioned above, the regeneration method by solvent washing can be conducted in several approaches, considering the fabrication method and the mechanical strength of the catalytic membranes. However, in many previous works, the regeneration method is simply described as “washing” without the detailed procedures. Similarly, detailed information for the thermal regeneration process, such as the temperature and calcination time, is also missing in some references.
In addition, similar to the case when evaluating the performance stability, the removal efficiency of the target pollutant is also the main (in most cases the only) index to evaluate the regeneration property. More attention should be paid to other factors including flux, TMP, and TOC/COD removals. The presence of inorganic ions and natural organic matter in the water matrix may lead to more serious membrane fouling, making the current regeneration methods less effective.
Up until now, when AOP was used for membrane regeneration, the same oxidant or external power as that employed for the treatment of polluted water was generally utilized. Meanwhile, none of the reported regeneration methods can solve the problem of metal leaching.

5. Recommendations for Future Works

Although the performance stability and regeneration of catalytic membranes have been extensively investigated, further research in this field is still required. Several recommendations for future works are listed as follows.
  • In future works, it is suggested to always include the performance stability or regeneration property of catalytic membranes when evaluating their abilities for pollutant removal. Catalytic membranes with high activity but poor stability and regeneration property are far away from practical applications.
  • When evaluating the stability and regeneration property, it is suggested to clearly describe the process with all the details provided. Besides the removal efficiency of pollutants, other factors including change in kinetic constant, TMP, flux, and TOC/COD removal should also be paid attention to present a more comprehensive profile.
  • Although the effects of inorganic ions and natural organic matter for the activity of fresh catalytic membranes are generally reported, their effects on the performance stability and regeneration property have been rarely investigated. In future works, more attention should be paid to this issue because these species may lead to serious membrane fouling, weakening the performance stability and making the current regeneration methods less effective.
  • The number of cycles in the multi-cycle run and the operation time in the long-time run summarized in Table 1 and Table 2 have been statistically analyzed. As shown in Figure 7a, for the number of cycles, most of the data were in the range of 5–10 cycles, and the median and average values were 5 and 6.3 cycles (n = 52). The operation time mostly fell in the range of 0–25 h with a median value of 33 h and an average value of 50.2 h (n = 37) (Figure 7b). For future works, it is suggested to select suitable cycle numbers and operation times based on these statistics. When possible, a large cycle number and a longer operation time are suggested, which can provide better references for potential practical applications of the investigated catalytic membranes.
  • The types and distributions of organic pollutants on the fouled membrane are different from those in the polluted water, and the utilization of another AOP for its regeneration may be more efficient. In other words, a certain AOP may be suitable to regenerate the fouled membrane, although it is not suitable to treat the polluted water. Through the combination of different AOPs, better process optimization and lower whole process cost may be achieved.
  • Faced with the decreased activity of metal-based membranes caused by metal-leaching, re-loading of metal species may be a possible solution, which is more economic than simply discarding the used membrane. However, the organic pollutants on the used membrane may impact the metal loading process, and more research on this issue is still in demand.
  • Almost all of the reports summarized and discussed above were conducted at bench-scale. To the best of our knowledge, there is no industrial scale application of this new technology at the current stage. Only several pilot-scale studies are available, where catalytic membranes were coupled with ozone oxidation [147,148,149], electrochemical oxidation [150,151], and electro-Fenton [152] processes. In these studies, to increase the throughput of the catalytic filtration process, two strategies have been utilized. One is to build catalytic membranes with larger sizes, as in the cases of Ti-Mn/ceramic membrane (Figure 8a) [147] and PbO2/Ti membrane (Figure 8b) [151]. Another is the numbering up strategy where multiple bench-scale membranes are integrated, such as in the cases of Ti-Mn/TiO2/ceramic membrane (Figure 8c) [148]. In these pilot studies, natural water sources with complex compositions were treated, leading to membrane fouling in the long run and making membrane regeneration inevitable. From a practical viewpoint, the large-sized membranes are difficult to fabricate and difficult to regenerate by some of the above-mentioned methods such as the heat treatment method, during which large-sized furnaces are required to fit the large-sized membranes, increasing the equipment investment. The numbering up strategy is convenient for membrane fabrication, but complex pipe systems would be required, making it difficult to maintain, disassemble, and regenerate the integrated multiple membranes. In future works, more research is required regarding the stability and regeneration property of pilot-scale catalytic membranes.
In summary, the current review further highlights the current achievements and hurdles in the performance stability and regeneration property of catalytic membranes, proposing recommendations for future works, including more comprehensive evaluation of the performance stability and regeneration property based on changes in the kinetic constant, TMP/flux, and TOC/COD removals as well as the effects of inorganic ions and natural organic matter, the selection of suitable cycle numbers and operation time, the utilization of a different AOP for membrane regeneration, the development of new regeneration methods, and more studies on the pilot and larger scales.

6. Conclusions

Catalytic membranes coupling AOP and membrane filtration have been widely employed for the treatment of various organic pollutants in water. In this review, the recent works concerning performance stability and regeneration property of these membranes have been summarized in terms of the different reaction systems including persulfate oxidation, Fenton oxidation, ozone oxidation, photocatalytic oxidation, and electrochemical oxidation. The performance stability of catalytic membranes was commonly evaluated by either multi-cycle experiments or long-time filtration, and the main deactivation reasons included the adsorption of pollutants and its oxidation intermediates, the changes in chemical properties of the membranes (valence state or surface functionalities), as well as leaching of metal species. Faced with these inactivation reasons, various regeneration methods have been developed. Washing with solvents is the most extensively used regeneration method, and different solvents including water, ethanol, and acidic/alkaline solutions have been investigated. Heat treatment is another popular approach, during which the employed atmosphere and heating temperature should be carefully selected considering the properties of the catalytic membrane. Other regeneration methods were also discussed, including AOP, electrochemical reduction, electrochemical backwashing, and chemical reduction. This review further highlights the current achievements and hurdles in the performance stability and regeneration property of catalytic membranes, proposing recommendations for future works.

Author Contributions

Conceptualization, Y.S. (Yawei Shi); methodology, Y.S. (Yawei Shi), T.Z. and Q.C.; investigation, Y.S. (Yawei Shi), T.Z., Q.C., C.M. and Y.Y.; resources, Y.S. (Yawei Shi), S.W., Z.P., Y.S. (Ya Sun) and G.D.; writing—original draft preparation, Y.S. (Yawei Shi); writing—review and editing, Y.S. (Yawei Shi); funding acquisition, Y.S. (Yawei Shi). All authors have read and agreed to the published version of the manuscript.

Funding

This work is financially supported by the National Natural Science Foundation of China (51908409), the Foundation (No. BCERE202201) of Tianjin Key Laboratory of Brine Chemical Engineering and Resource Eco-utilization (Tianjin University of Science & Technology), China, and the Fundamental Research Funds for the Central Universities (3132023162).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data analyzed during this study are included in this published article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Giwa, A.; Yusuf, A.; Balogun, H.A.; Sambudi, N.S.; Bilad, M.R.; Adeyemi, I.; Chakraborty, S.; Curcio, S. Recent advances in advanced oxidation processes for removal of contaminants from water: A comprehensive review. Process. Saf. Environ. 2021, 146, 220–256. [Google Scholar] [CrossRef]
  2. Wang, J.; Zhuan, R. Degradation of antibiotics by advanced oxidation processes: An overview. Sci. Total. Environ. 2020, 701, 135023. [Google Scholar] [CrossRef] [PubMed]
  3. Asif, M.B.; Kang, H.; Zhang, Z. Gravity-driven layered double hydroxide nanosheet membrane activated peroxymonosulfate system for micropollutant degradation. J. Hazard. Mater. 2022, 425, 127988. [Google Scholar] [CrossRef] [PubMed]
  4. Yue, R.; Sun, X. A Self-cleaning, catalytic titanium carbide (MXene) membrane for efficient tetracycline degradation through peroxymonosulfate activation: Performance evaluation and mechanism study. Sep. Purif. Technol. 2021, 279, 119796. [Google Scholar] [CrossRef]
  5. Pan, Z.; Song, C.; Li, L.; Wang, H.; Pan, Y.; Wang, C.; Li, J.; Wang, T.; Feng, X. Membrane technology coupled with electrochemical advanced oxidation processes for organic wastewater treatment: Recent advances and future prospects. Chem. Eng. J. 2019, 376, 120909. [Google Scholar] [CrossRef]
  6. Zhu, X.; Jassby, D. Electroactive Membranes for Water Treatment: Enhanced Treatment Functionalities, Energy Considerations, and Future Challenges. Acc. Chem. Res. 2019, 52, 1177–1186. [Google Scholar] [CrossRef]
  7. Wang, X.; Li, F.; Hu, X.; Hua, T. Electrochemical advanced oxidation processes coupled with membrane filtration for degrading antibiotic residues: A review on its potential applications, advances, and challenges. Sci. Total. Environ. 2021, 784, 146912. [Google Scholar] [CrossRef]
  8. Yan, H.; Lai, C.; Wang, D.; Liu, S.; Li, X.; Zhou, X.; Yi, H.; Li, B.; Zhang, M.; Li, L.; et al. In situ chemical oxidation: Peroxide or persulfate coupled with membrane technology for wastewater treatment. J. Mater. Chem. A 2021, 9, 11944–11960. [Google Scholar] [CrossRef]
  9. Yu, C.; Xiong, Z.; Zhou, H.; Zhou, P.; Zhang, H.; Huang, R.; Yao, G.; Lai, B. Marriage of membrane filtration and sulfate radical-advanced oxidation processes (SR-AOPs) for water purification: Current developments, challenges and prospects. Chem. Eng. J. 2021, 433, 133802. [Google Scholar] [CrossRef]
  10. Zhou, M.; Chen, J.; Yu, S.; Chen, B.; Chen, C.; Shen, L.; Li, B.; Lin, H. The coupling of persulfate activation and membrane separation for the effective pollutant degradation and membrane fouling alleviation. Chem. Eng. J. 2023, 451, 139009. [Google Scholar] [CrossRef]
  11. Li, N.; Lu, X.; He, M.; Duan, X.; Yan, B.; Chen, G.; Wang, S. Catalytic membrane-based oxidation-filtration systems for organic wastewater purification: A review. J. Hazard. Mater. 2021, 414, 125478. [Google Scholar] [CrossRef] [PubMed]
  12. Titchou, F.E.; Zazou, H.; Afanga, H.; Gaayda, J.E.; Akbour, R.A.; Nidheesh, P.V.; Hamdani, M. Removal of organic pollutants from wastewater by advanced oxidation processes and its combination with membrane processes. Chem. Eng. Process.-Process. Intensif. 2021, 169, 108631. [Google Scholar] [CrossRef]
  13. Sun, X.; Zheng, H.; Jiang, S.; Zhu, M.; Zhou, Y.; Wang, D.; Fan, Y.; Zhang, D.; Zhang, L. Fabrication of FeOCl/MoS2 catalytic membranes for pollutant degradation and alleviating membrane fouling with peroxymonosulfate activation. J. Environ. Chem. Eng. 2022, 10, 107717. [Google Scholar] [CrossRef]
  14. Chen, B.; Hu, X.; Wang, J.; Li, R.; Shen, L.; Xu, Y.; Zhang, M.; Hong, H.; Lin, H. Novel catalytic self-cleaning membrane with peroxymonosulfate activation for dual-function wastewater purification: Performance and mechanism. J. Clean. Prod. 2022, 355, 131858. [Google Scholar] [CrossRef]
  15. Ye, J.; Dai, J.; Yang, D.; Li, C.; Yan, Y.; Wang, Y. 2D/2D confinement graphene-supported bimetallic Sulfides/g-C3N4 composites with abundant sulfur vacancies as highly active catalytic self-cleaning membranes for organic contaminants degradation. Chem. Eng. J. 2021, 418, 129383. [Google Scholar] [CrossRef]
  16. Han, Y.; Jiang, B.; Zhang, C.; Zhang, L.; Zhang, L.; Sun, Y.; Yang, N. Co@N-C nanocatalysts anchored in confined membrane pores for instantaneous pollutants degradation and antifouling via peroxymonosulfate activation. J. Water Process Eng. 2022, 47, 102639. [Google Scholar] [CrossRef]
  17. Ye, J.; Dai, J.; Li, C.; Yan, Y. Lawn-like Co3O4@N-doped carbon-based catalytic self-cleaning membrane with peroxymonosulfate activation: A highly efficient singlet oxygen dominated process for sulfamethoxazole degradation. Chem. Eng. J. 2021, 421, 102639. [Google Scholar] [CrossRef]
  18. Ye, J.; Wang, Y.; Li, Z.; Yang, D.; Li, C.; Yan, Y.; Dai, J. 2D confinement freestanding graphene oxide composite membranes with enriched oxygen vacancies for enhanced organic contaminants removal via peroxymonosulfate activation. J. Hazard. Mater. 2021, 417, 126028. [Google Scholar] [CrossRef]
  19. Pedrosa, M.; Drazic, G.; Tavares, P.B.; Figueiredo, J.L.; Silva, A.M.T. Metal-free graphene-based catalytic membrane for degradation of organic contaminants by persulfate activation. Chem. Eng. J. 2019, 369, 223–232. [Google Scholar] [CrossRef]
  20. Zhang, L.; Yang, N.; Han, Y.; Wang, X.; Liu, S.; Zhang, L.; Sun, Y.; Jiang, B. Development of polyacrylonitrile/perovskite catalytic membrane with abundant channel-assisted reaction sites for organic pollutant removal. Chem. Eng. J. 2022, 437, 135163. [Google Scholar] [CrossRef]
  21. Liu, W.; Zhang, Y.Q.; Zhang, L.F.; Guo, J.Q.; Wei, J.H. Polysulfone Ultrafiltration Membrane Promoted by Brownmillerite SrCuxCo1−xO3−λ-Deposited MCM-41 for Industrial Wastewater Decontamination: Catalytic Oxidation and Antifouling Properties. Ind. Eng. Chem. Res. 2020, 59, 7805–7815. [Google Scholar] [CrossRef]
  22. Ma, H.; Zhang, X.; Feng, G.; Ren, B.; Pan, Z.; Shi, Y.; Xu, R.; Wang, P.; Liu, Y.; Wang, G.; et al. Carbon nanotube membrane armed with confined iron for peroxymonosulfate activation towards efficient tetracycline removal. Sep. Purif. Technol. 2023, 312, 123319. [Google Scholar] [CrossRef]
  23. Qian, F.; Luo, J.; Yin, H.; Liu, F.; Gao, S.; Gu, X. Carbonaceous composite membranes for peroxydisulfate activation to remove sulfamethoxazole in a real water matrix. Chemosphere 2022, 288, 132597. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, S.; Tian, J.; Wang, Q.; Zhao, Z.; Cui, F.; Li, G. Low-temperature sintered high-strength CuO doped ceramic hollow fiber membrane: Preparation, characterization and catalytic activity. J. Membr. Sci. 2019, 570–571, 333–342. [Google Scholar] [CrossRef]
  25. Wang, S.X.; Tian, J.Y.; Wang, Z.H.; Wang, Q.; Jia, J.L.; Hao, X.J.; Gao, S.S.; Cui, F.Y. Integrated process for membrane fouling mitigation and organic pollutants removal using copper oxide modified ceramic hollow fiber membrane with in-situ peroxymonosulfate activation. Chem. Eng. J. 2020, 396, 11. [Google Scholar] [CrossRef]
  26. Wang, S.; Tian, J.; Wang, Q.; Xiao, F.; Gao, S.; Shi, W.; Cui, F. Development of CuO coated ceramic hollow fiber membrane for peroxymonosulfate activation: A highly efficient singlet oxygen-dominated oxidation process for bisphenol a degradation. Appl. Catal. B Environ. 2019, 256, 117783. [Google Scholar] [CrossRef]
  27. Cheng, X.; Zhang, Y.; Fan, Q.; Wang, L.; Shi, S.; Luo, X.; Zhu, X.; Wu, D.; Liang, H. Preparation of Co3O4@carbon nanotubes modified ceramic membrane for simultaneous catalytic oxidation and filtration of secondary effluent. Chem. Eng. J. 2023, 454, 140450. [Google Scholar] [CrossRef]
  28. Chen, L.; Maqbool, T.; Fu, W.; Yang, Y.; Hou, C.; Guo, J.; Zhang, X. Highly efficient manganese (III) oxide submerged catalytic ceramic membrane for nonradical degradation of emerging organic compounds. Sep. Purif. Technol. 2022, 295, 121272. [Google Scholar] [CrossRef]
  29. Wang, S.; Wu, Q.; Yan, B.; Guo, Y.; Xia, W.; Li, J.; Cui, F.; Tian, J. A novel integrated process of ceramic membrane filtration coupled with peroxymonosulfate activation and adsorption for water treatment. Sep. Purif. Technol. 2022, 291, 120874. [Google Scholar] [CrossRef]
  30. Zhao, J.; Zhang, H.; Huang, Q.; Xiao, C. Poly(tetrafluoroethylene-co-hexafluoropropylene)/Ferric Oxide Hybrid Membranes for High Concentration of Dye Wastewater Treatment by Heterogeneous Fenton-Like Catalysis. Catal. Lett. 2021, 151, 3020–3029. [Google Scholar] [CrossRef]
  31. Karimnezhad, H.; Navarchian, A.H.; Tavakoli Gheinani, T.; Zinadini, S. Amoxicillin removal by Fe-based nanoparticles immobilized on polyacrylonitrile membrane: Individual nanofiltration or Fenton reaction, vs. engineered combined process. Chem. Eng. Res. Des. 2020, 153, 187–200. [Google Scholar] [CrossRef]
  32. Scaratti, G.; De Noni Junior, A.; Jose, H.J.; Peralta Muniz Moreira, R.D.F. 1,4-Dioxane removal from water and membrane fouling elimination using CuO-coated ceramic membrane coupled with ozone. Environ. Sci. Pollut. Res. 2020, 27, 22144–22154. [Google Scholar] [CrossRef] [PubMed]
  33. He, Y.; Wang, L.; Chen, Z.; Huang, X.; Wang, X.; Zhang, X.; Wen, X. Novel catalytic ceramic membranes anchored with MnMe oxide and their catalytic ozonation performance towards atrazine degradation. J. Membr. Sci. 2022, 648, 120362. [Google Scholar] [CrossRef]
  34. Guo, Y.; Song, Z.; Xu, B.; Li, Y.; Qi, F.; Croue, J.P.; Yuan, D. A novel catalytic ceramic membrane fabricated with CuMn2O4 particles for emerging UV absorbers degradation from aqueous and membrane fouling elimination. J. Hazard. Mater. 2018, 344, 1229–1239. [Google Scholar] [CrossRef] [PubMed]
  35. Song, Z.; Sun, J.; Wang, W.; Wang, Z.; Zhang, Y.; Xu, B.; Qi, F. Stable synergistic decontamination and self-cleaning performance of powerful N-rGO catalytic ozonation membrane: Clustering effect of free electrons and role of interface properties. Appl. Catal. B Environ. 2021, 283, 119662. [Google Scholar] [CrossRef]
  36. Dekkouche, S.; Morales-Torres, S.; Ribeiro, A.R.; Faria, J.L.; Fontàs, C.; Kebiche-Senhadji, O.; Silva, A.M.T. In situ growth and crystallization of TiO2 on polymeric membranes for the photocatalytic degradation of diclofenac and 17α-ethinylestradiol. Chem. Eng. J. 2022, 427, 131476. [Google Scholar] [CrossRef]
  37. Mastropietro, T.F.; Meringolo, C.; Poerio, T.; Scarpelli, F.; Godbert, N.; Di Profio, G.; Fontananova, E. Multistimuli Activation of TiO2/alpha-Alumina Membranes for Degradation of Methylene Blue. Ind. Eng. Chem. Res. 2017, 56, 11049–11057. [Google Scholar] [CrossRef]
  38. Li, C.; Lu, Z.; Ao, X.; Sun, W.; Huang, X. Degradation kinetics and removal efficiencies of pharmaceuticals by photocatalytic ceramic membranes using ultraviolet light-emitting diodes. Chem. Eng. J. 2022, 427, 130828. [Google Scholar] [CrossRef]
  39. Nair, A.K.; JagadeeshBabu, P.E. TiO2 nanosheet-graphene oxide based photocatalytic hierarchical membrane for water purification. Surf. Coat. Technol. 2017, 320, 259–262. [Google Scholar] [CrossRef]
  40. Chen, W.; Ye, T.; Xu, H.; Chen, T.; Geng, N.; Gao, X. An ultrafiltration membrane with enhanced photocatalytic performance from grafted N–TiO2/graphene oxide. RSC Adv. 2017, 7, 9880–9887. [Google Scholar] [CrossRef]
  41. Gao, Y.; Yan, N.; Jiang, C.; Xu, C.; Yu, S.; Liang, P.; Zhang, X.; Liang, S.; Huang, X. Filtration-enhanced highly efficient photocatalytic degradation with a novel electrospun rGO@TiO2 nanofibrous membrane: Implication for improving photocatalytic efficiency. Appl. Catal. B Environ. 2020, 268, 118737. [Google Scholar] [CrossRef]
  42. Liu, T.; Wang, L.; Liu, X.; Sun, C.; Lv, Y.; Miao, R.; Wang, X. Dynamic photocatalytic membrane coated with ZnIn2S4 for enhanced photocatalytic performance and antifouling property. Chem. Eng. J. 2020, 379, 122379. [Google Scholar] [CrossRef]
  43. Yan, C.; Cheng, Z.; Wei, J.; Xu, Q.; Zhang, X.; Wei, Z. Efficient degradation of antibiotics by photo-Fenton reactive ceramic membrane with high flux by a facile spraying method under visible LED light. J. Clean. Prod. 2022, 366, 132849. [Google Scholar] [CrossRef]
  44. Lv, Y.; Zhang, C.; He, A.; Yang, S.J.; Wu, G.P.; Darling, S.B.; Xu, Z.K. Photocatalytic Nanofiltration Membranes with Self-Cleaning Property for Wastewater Treatment. Adv. Funct. Mater. 2017, 27, 8. [Google Scholar] [CrossRef]
  45. Deng, R.; Xia, X.-Z.; Han, J.-C.; Wu, Q.-Y.; Yang, H.-C. Siphon-driven interfacial photocatalytic reactors enhanced by capillary flow for continuous wastewater treatment. Sep. Purif. Technol. 2022, 300, 121835. [Google Scholar] [CrossRef]
  46. Wang, Y.; Yang, T.; Chen, J.; Wen, S.; Li, D.; Wang, B.; Zhang, Q. Multifunctional ferrocene-based photo-Fenton membrane: An efficient integration of rejection and catalytic process. Sep. Purif. Technol. 2022, 298, 121557. [Google Scholar] [CrossRef]
  47. Li, N.; Chen, G.Y.; Zhao, J.H.; Yan, B.B.; Cheng, Z.J.; Meng, L.J.; Chen, V. Self-cleaning PDA/ZIF-67@PP membrane for dye wastewater remediation with peroxymonosulfate and visible light activation. J. Membr. Sci. 2019, 591, 9. [Google Scholar] [CrossRef]
  48. Zhu, Z.S.; Qu, J.; Hao, S.M.; Han, S.; Jia, K.L.; Yu, Z.Z. alpha-Fe2O3 Nanodisk/Bacterial Cellulose Hybrid Membranes as High-Performance Sulfate-Radical-Based Visible Light Photocatalysts under Stirring/Flowing States. ACS Appl. Mater. Interfaces 2018, 10, 30670–30679. [Google Scholar] [CrossRef]
  49. Zhang, Y.Z.; Chen, Q.Y.; Liu, L.F.; Wang, Y.; Leung, M.K.H. Activation of peroxymonosulfate and recycled effluent filtration over cathode membrane CNFs-CoFe2O4/PVDF in a photocatalytic fuel cell for water pollution control. Chem. Eng. J. 2020, 399, 125731. [Google Scholar] [CrossRef]
  50. Wang, G.; Wang, D.; Dong, X.; Zhang, X.; Ma, H. Sodium persulfate based PVDF membrane for concurrent advanced oxidation and ultrafiltration of ofloxacin in water. Chem. Eng. J. 2017, 315, 509–515. [Google Scholar] [CrossRef]
  51. Ren, L.; Chen, M.; Ma, J.; Li, Y.; Wang, Z. Pd–O2 interaction and singlet oxygen formation in a novel reactive electrochemical membrane for ultrafast sulfamethoxazole oxidation. Chem. Eng. J. 2022, 428, 131194. [Google Scholar] [CrossRef]
  52. Yu, S.; Gao, Y.; Khan, R.; Liang, P.; Zhang, X.; Huang, X. Electrospun PAN-based graphene/SnO2 carbon nanofibers as anodic electrocatalysis microfiltration membrane for sulfamethoxazole degradation. J. Membr. Sci. 2020, 614, 118368. [Google Scholar] [CrossRef]
  53. Qian, X.; Xu, L.; Zhu, Y.; Yu, H.; Niu, J. Removal of aqueous triclosan using TiO2 nanotube arrays reactive membrane by sequential adsorption and electrochemical degradation. Chem. Eng. J. 2021, 420, 127615. [Google Scholar] [CrossRef]
  54. Tan, T.-Y.; Zeng, Z.-T.; Zeng, G.-M.; Gong, J.-L.; Xiao, R.; Zhang, P.; Song, B.; Tang, W.-W.; Ren, X.-Y. Electrochemically enhanced simultaneous degradation of sulfamethoxazole, ciprofloxacin and amoxicillin from aqueous solution by multi-walled carbon nanotube filter. Sep. Purif. Technol. 2020, 235, 116167. [Google Scholar] [CrossRef]
  55. Pan, M.; Wang, J.; Gao, G.; Chew, J.W. Incorporation of single cobalt active sites onto N-doped graphene for superior conductive membranes in electrochemical filtration. J. Membr. Sci. 2020, 602, 117966. [Google Scholar] [CrossRef]
  56. Malekabadi, F.K.; Yousefi, F.; Karimi, R.; Ghaedi, M.; Dashtian, K. Electrocatalytic membrane containing CuFeO2/nanoporous carbon for organic dye removal application. Chem. Eng. Res. Des. 2022, 183, 345–356. [Google Scholar] [CrossRef]
  57. Wei, G.; Quan, X.; Fan, X.; Chen, S.; Zhang, Y. Carbon-nanotube-based sandwich-like hollow fiber membranes for expanded microcystin-LR removal applications. Chem. Eng. J. 2017, 319, 212–218. [Google Scholar] [CrossRef]
  58. Chen, M.; Zheng, J.; Dai, R.; Wu, Z.; Wang, Z. Preferential removal of 2,4-dichlorophenoxyacetic acid from contaminated waters using an electrocatalytic ceramic membrane filtration system: Mechanisms and implications. Chem. Eng. J. 2020, 387, 124132. [Google Scholar] [CrossRef]
  59. Wang, X.; Wang, H.; Li, F.; Hu, X.; Xie, Z.; Hua, T. Activation of peroxymonosulfate in an electrochemical filter by MnFe2O4-rGO electro-assisted catalytic membrane for the degradation of oxytetracycline. J. Environ. Chem. Eng. 2022, 10, 107008. [Google Scholar] [CrossRef]
  60. Zhao, Y.; Sun, M.; Zhao, Y.; Wang, L.; Lu, D.; Ma, J. Electrified ceramic membrane actuates non-radical mediated peroxymonosulfate activation for highly efficient water decontamination. Water Res. 2022, 225, 119140. [Google Scholar] [CrossRef]
  61. Kumari, P.; Bahadur, N.; Conlan, X.A.; Zeng, X.; Kong, L.; O’Dell, L.A.; Sadek, A.; Merenda, A.; Dumée, L.F. Stimuli-responsive heterojunctions based photo-electrocatalytic membrane reactors for reactive filtration of persistent organic pollutants. Chem. Eng. J. 2023, 452, 139374. [Google Scholar] [CrossRef]
  62. Wang, X.; Wang, G.; Chen, S.; Fan, X.; Quan, X.; Yu, H. Integration of membrane filtration and photoelectrocatalysis on g-C3N4/CNTs/Al2O3 membrane with visible-light response for enhanced water treatment. J. Membr. Sci. 2017, 541, 153–161. [Google Scholar] [CrossRef]
  63. Zhang, L.; Yang, N.; Han, Y.; Wang, X.; Zhang, L.; Sun, Y.; Jiang, B. Highly dispersed beta-FeOOH nanocatalysts anchored in confined membrane pores for simultaneously improving catalytic and separation performance. Sep. Purif. Technol. 2021, 279, 119684. [Google Scholar] [CrossRef]
  64. Chen, Y.; Zhang, G.; Liu, H.; Qu, J. Confining Free Radicals in Close Vicinity to Contaminants Enables Ultrafast Fenton-like Processes in the Interspacing of MoS2 Membranes. Angew. Chem. Int. Ed. Engl. 2019, 58, 8134–8138. [Google Scholar] [CrossRef]
  65. Meng, C.; Wang, Z.; Zhang, W.; Cui, L.; Yang, B.; Xie, H.; Zhang, Z. Laminar membranes assembled by ultrathin cobalt-copper oxide nanosheets for nanoconfined catalytic degradation of contaminants. Chem. Eng. J. 2022, 449, 137811. [Google Scholar] [CrossRef]
  66. Lu, N.; Lin, H.; Xu, S.; Wang, J.; Han, Q.; Liu, F. Bone/muscle-inspired polymer porous matrix toughened carbon nanofibrous catalytic membranes for robust emerging contaminants removal. Chem. Eng. J. 2022, 442, 136069. [Google Scholar] [CrossRef]
  67. Xie, J.; Liao, Z.P.; Zhang, M.; Ni, L.H.; Qi, J.W.; Wang, C.H.; Sun, X.Y.; Wang, L.J.; Wang, S.B.; Li, J.S. Sequential Ultrafiltration-Catalysis Membrane for Excellent Removal of Multiple Pollutants in Water. Environ. Sci. Technol. 2021, 55, 2652–2661. [Google Scholar] [CrossRef]
  68. Zhang, Y.; Chen, X.; Liang, C.; Yin, L.; Yang, Y. Reconstructing the coordination environment of single atomic Fe-catalysts for boosting the Fenton-like degradation activities. Appl. Catal. B Environ. 2022, 315, 121536. [Google Scholar] [CrossRef]
  69. Wang, Z.; Meng, C.; Zhang, W.; Zhang, S.; Yang, B.; Zhang, Z. Honeycomb-like holey Co3O4 membrane triggered peroxymonosulfate activation for rapid degradation of organic contaminants. Sci. Total Environ. 2022, 814, 152698. [Google Scholar] [CrossRef]
  70. Meng, C.; Ding, B.; Zhang, S.; Cui, L.; Ostrikov, K.K.; Huang, Z.; Yang, B.; Kim, J.H.; Zhang, Z. Angstrom-confined catalytic water purification within Co-TiOx laminar membrane nanochannels. Nat. Commun. 2022, 13, 4010. [Google Scholar] [CrossRef]
  71. Li, Y.; He, Y.; Zhuang, J.; Shi, H. Hierarchical microsphere encapsulated in graphene oxide composite for durable synergetic membrane separation and Fenton-like degradation. Chem. Eng. J. 2022, 430, 133124. [Google Scholar] [CrossRef]
  72. Luo, J.; Liu, T.; Qian, F.; Xia, X.; Zhou, X.; Zhu, Y. Boosting non-radical oxidation in peroxydisulfate activation with carbonaceous catalytic membranes by coupling structural defects and nitrogen doping sites. J. Environ. Chem. Eng. 2022, 10, 108101. [Google Scholar] [CrossRef]
  73. Zhang, C.; Wang, G.; Gao, Y.; Ma, H.; Dong, X.; Zhang, X. Surface and pore co-functionalized ceramic membrane with nitrogen doped carbon for enhanced water treatment through coupling peroxymonosulfate activation. Sep. Purif. Technol. 2022, 292, 120996. [Google Scholar] [CrossRef]
  74. Chen, L.; Maqbool, T.; Nazir, G.; Hou, C.; Yang, Y.; Guo, J.; Zhang, X. Developing the large-area manganese-based catalytic ceramic membrane for peroxymonosulfate activation: Applications in degradation of endocrine disrupting compounds in drinking water. J. Membr. Sci. 2022, 655, 120602. [Google Scholar] [CrossRef]
  75. Lu, N.; Lin, H.; Li, G.; Wang, J.; Han, Q.; Liu, F. ZIF-67 derived nanofibrous catalytic membranes for ultrafast removal of antibiotics under flow-through filtration via non-radical dominated pathway. J. Membr. Sci. 2021, 639, 119782. [Google Scholar] [CrossRef]
  76. Lin, H.; Fang, Q.; Wang, W.; Li, G.; Guan, J.; Shen, Y.; Ye, J.; Liu, F. Prussian blue/PVDF catalytic membrane with exceptional and stable Fenton oxidation performance for organic pollutants removal. Appl. Catal. B Environ. 2020, 273, 119047. [Google Scholar] [CrossRef]
  77. Huang, Z.H.; Zhang, X.; Wang, Y.X.; Sun, J.Y.; Zhang, H.; Liu, W.L.; Li, M.P.; Ma, X.H.; Xu, Z.L. Fe3O4/PVDF catalytic membrane treatment organic wastewater with simultaneously improved permeability, catalytic property and anti-fouling. Environ. Res. 2020, 187, 8. [Google Scholar] [CrossRef]
  78. Qu, W.; Chen, C.; Tang, Z.; Xia, D.; Ma, D.; Huang, Y.; Lian, Q.; He, C.; Shu, D.; Han, B. Electron-rich/poor reaction sites enable ultrafast confining Fenton-like processes in facet-engineered BiOI membranes for water purification. Appl. Catal. B Environ. 2022, 304, 120970. [Google Scholar] [CrossRef]
  79. Song, H.-M.; Zhu, L.-J.; Wang, Y.; Wang, G.; Zeng, Z.-X. Fe-based Prussian blue cubes confined in graphene oxide nanosheets for the catalytic degradation of dyes in wastewater. Sep. Purif. Technol. 2022, 288, 120676. [Google Scholar] [CrossRef]
  80. Yi, Q.; Li, Y.; Dai, R.; Li, X.; Li, Z.; Wang, Z. Efficient removal of neonicotinoid by singlet oxygen dominated MoSx/ceramic membrane-integrated Fenton-like process. J. Hazard. Mater. 2022, 439, 129672. [Google Scholar] [CrossRef]
  81. Zhang, S.; Hedtke, T.; Zhu, Q.; Sun, M.; Weon, S.; Zhao, Y.; Stavitski, E.; Elimelech, M.; Kim, J.H. Membrane-Confined Iron Oxychloride Nanocatalysts for Highly Efficient Heterogeneous Fenton Water Treatment. Environ. Sci. Technol. 2021, 55, 9266–9275. [Google Scholar] [CrossRef] [PubMed]
  82. Ding, J.; Sarrigani, G.V.; Qu, J.; Ebrahimi, A.; Zhong, X.; Hou, W.-C.; Cairney, J.M.; Huang, J.; Wiley, D.E.; Wang, D.K. Designing Co3O4/silica catalysts and intensified ultrafiltration membrane-catalysis process for wastewater treatment. Chem. Eng. J. 2021, 419, 129465. [Google Scholar] [CrossRef]
  83. Gao, B.; Chen, W.; Liu, J.; An, J.; Wang, L.; Zhu, Y.; Sillanpää, M. Continuous removal of tetracycline in a photocatalytic membrane reactor (PMR) with ZnIn2S4 as adsorption and photocatalytic coating layer on PVDF membrane. J. Photochem. Photobiol. A Chem. 2018, 364, 732–739. [Google Scholar] [CrossRef]
  84. Zeng, H.J.; Yu, Z.X.; Shao, L.Y.; Li, X.H.; Zhu, M.; Liu, Y.C.; Feng, X.F.; Zhu, X.M. Ag2CO3@UiO-66-NH2 embedding graphene oxide sheets photocatalytic membrane for enhancing the removal performance of Cr(VI) and dyes based on filtration. Desalination 2020, 491, 11. [Google Scholar] [CrossRef]
  85. Zangeneh, H.; Zinatizadeh, A.A.; Zinadini, S.; Feyzi, M.; Bahnemann, D.W. Preparation and characterization of a novel photocatalytic self-cleaning PES nanofiltration membrane by embedding a visible-driven photocatalyst boron doped-TiO2-SiO2/CoFe2O4 nanoparticles. Sep. Purif. Technol. 2019, 209, 764–775. [Google Scholar] [CrossRef]
  86. Lu, M.; Zhang, H. Preparation and decontamination performance of a flexible self-standing hydrogel photocatalytic membrane. J. Membr. Sci. 2022, 644, 119979. [Google Scholar] [CrossRef]
  87. Berger, T.E.; Regmi, C.; Schäfer, A.I.; Richards, B.S. Photocatalytic degradation of organic dye via atomic layer deposited TiO2 on ceramic membranes in single-pass flow-through operation. J. Membr. Sci. 2020, 604, 118015. [Google Scholar] [CrossRef]
  88. Yin, Z.; Zhang, K.; Ma, N.; Liu, X.; Yin, Z.; Wang, H.; Yang, X.; Wang, Y.; Qin, X.; Cheng, D.; et al. Catalytic membrane electrode with Co3O4 nanoarrays for simultaneous recovery of water and generation of hydrogen from wastewater. Sci. China Mater. 2022, 66, 651–663. [Google Scholar] [CrossRef]
  89. Pan, Z.L.; Yu, F.P.; Li, L.; Song, C.W.; Yang, J.W.; Wang, C.L.; Pan, Y.Q.; Wang, T.H. Electrochemical microfiltration treatment of bisphenol A wastewater using coal-based carbon membrane. Sep. Purif. Technol. 2019, 227, 8. [Google Scholar] [CrossRef]
  90. Li, X.; Shao, S.; Yang, Y.; Mei, Y.; Qing, W.; Guo, H.; Peng, L.E.; Wang, P.; Tang, C.Y. Engineering Interface with a One-Dimensional RuO2/TiO2 Heteronanostructure in an Electrocatalytic Membrane Electrode: Toward Highly Efficient Micropollutant Decomposition. Acs Appl. Mater. Interfaces 2020, 12, 21596–21604. [Google Scholar] [CrossRef]
  91. Zheng, J.J.; Wang, Z.W.; Ma, J.X.; Xu, S.P.; Wu, Z.C. Development of an Electrochemical Ceramic Membrane Filtration System for Efficient Contaminant Removal from Waters. Environ. Sci. Technol. 2018, 52, 4117–4126. [Google Scholar] [CrossRef] [PubMed]
  92. Jiang, W.L.; Xia, X.; Han, J.L.; Ding, Y.C.; Haider, M.R.; Wang, A.J. Graphene Modified Electro-Fenton Catalytic Membrane for in Situ Degradation of Antibiotic Florfenicol. Environ. Sci. Technol. 2018, 52, 9972–9982. [Google Scholar] [CrossRef] [PubMed]
  93. 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]
  94. Zheng, W.; Liu, Y.; Liu, W.; Ji, H.; Li, F.; Shen, C.; Fang, X.; Li, X.; Duan, X. A novel electrocatalytic filtration system with carbon nanotube supported nanoscale zerovalent copper toward ultrafast oxidation of organic pollutants. Water Res. 2021, 194, 116961. [Google Scholar] [CrossRef] [PubMed]
  95. Wu, H.; Xu, X.; Shi, L.; Yin, Y.; Zhang, L.-C.; Wu, Z.; Duan, X.; Wang, S.; Sun, H. Manganese oxide integrated catalytic ceramic membrane for degradation of organic pollutants using sulfate radicals. Water Res. 2019, 167, 115110. [Google Scholar] [CrossRef]
  96. Wang, X.; Li, Y.; Yu, H.; Yang, F.; Tang, C.Y.; Quan, X.; Dong, Y. High-flux robust ceramic membranes functionally decorated with nano-catalyst for emerging micro-pollutant removal from water. J. Membr. Sci. 2020, 611, 118281. [Google Scholar] [CrossRef]
  97. Wei, J.; Bi, J.; Zhang, L.; Han, D.; Gong, J. Gravity-driven Fe-doped CoTiO3/SiO2 fiber membrane with open catalytic network: Activation of peroxymonosulfate and efficient pollutants removal. Sep. Purif. Technol. 2022, 280, 119975. [Google Scholar] [CrossRef]
  98. Wang, S.; Wu, Q.; Huang, R.; Guo, Y.; Yan, B.; Hao, X.; Li, J.; Xia, W.; Tian, J. Catalytic ceramic membrane integrated with granular activated carbon for efficient removal of organic pollutants. J. Water Process Eng. 2022, 47, 102751. [Google Scholar] [CrossRef]
  99. Wang, L.; Wang, L.; Shi, Y.; Zhu, J.; Zhao, B.; Zhang, Z.; Ding, G.; Zhang, H. Fabrication of Co3O4-Bi2O3-Ti catalytic membrane for efficient degradation of organic pollutants in water by peroxymonosulfate activation. J. Colloid Interface Sci. 2022, 607, 451–461. [Google Scholar] [CrossRef]
  100. Ma, H.; Li, X.; Pan, Z.; Xu, R.; Wang, P.; Li, H.; Shi, Y.; Fan, X.; Song, C. MOF derivative functionalized titanium-based catalytic membrane for efficient sulfamethoxazole removal via peroxymonosulfate activation. J. Membr. Sci. 2022, 661, 120924. [Google Scholar] [CrossRef]
  101. Chen, S.; Teng, Q. Quantitative Immobilization of Phthalocyanine onto Bacterial Cellulose for Construction of a High-Performance Catalytic Membrane Reactor. Materials 2017, 10, 846. [Google Scholar] [CrossRef] [PubMed]
  102. Gao, N.; Liang, F.; Wang, X.; Li, B. Loose composite nanofiltration membrane with in-situ immobilized β-FeOOH film for effective dyes degradation and separation. Colloids Surf. A Physicochem. Eng. Asp. 2022, 654, 130115. [Google Scholar] [CrossRef]
  103. Sun, M.H.; Zou, L.Z.; Wang, P.C.; Fan, X.F.; Pan, Z.L.; Liu, Y.M.; Song, C.W. Nano valent zero iron (NZVI) immobilized CNTs hollow fiber membrane for flow-through heterogeneous Fenton process. J. Environ. Chem. Eng. 2022, 10, 9. [Google Scholar] [CrossRef]
  104. Jiang, S.; Zhao, Z.; Cui, K.; Tang, Y.; Du, X.; He, B.; Li, M.; Feng, J.; Yu, B.; Xiong, W. Catalytic wet peroxide oxidation of phenolic wastewater on novel Cu/Mn-UiO-66@Al2O3 ceramic tube membrane catalysts. Chem. Eng. J. 2022, 430, 132787. [Google Scholar] [CrossRef]
  105. Li, M.; Yang, K.; Huang, X.; Liu, S.; Jia, Y.; Gu, P.; Miao, H. Efficient degradation of trimethoprim by catalytic ozonation coupled with Mn/FeOx-functionalized ceramic membrane: Synergic catalytic effect and enhanced anti-fouling performance. J. Colloid Interface Sci. 2022, 616, 440–452. [Google Scholar] [CrossRef]
  106. Espindola, J.C.; Cristovao, R.O.; Mendes, A.; Boaventura, R.A.R.; Vilar, V.J.P. Photocatalytic membrane reactor performance towards oxytetracycline removal from synthetic and real matrices: Suspended vs immobilized TiO2-P25. Chem. Eng. J. 2019, 378, 122114. [Google Scholar] [CrossRef]
  107. Ye, Z.; Oriol, R.; Yang, C.; Sirés, I.; Li, X.-Y. A novel NH2-MIL-88B(Fe)-modified ceramic membrane for the integration of electro-Fenton and filtration processes: A case study on naproxen degradation. Chem. Eng. J. 2022, 433, 133547. [Google Scholar] [CrossRef]
  108. Pan, Z.L.; Xin, H.; Xu, S.; Xu, R.S.; Wang, P.C.; Yuan, Y.; Fan, X.F.; Song, Y.X.; Song, C.W.; Wang, T.H. Preparation and performance of polyaniline modified coal-based carbon membrane for electrochemical filtration treatment of organic wastewater. Sep. Purif. Technol. 2022, 287, 13. [Google Scholar] [CrossRef]
  109. Pan, Z.; Xu, S.; Xin, H.; Yuan, Y.; Xu, R.; Wang, P.; Yan, X.; Fan, X.; Song, C.; Wang, T. High performance polypyrrole coated carbon-based electrocatalytic membrane for organic contaminants removal from aqueous solution. J. Colloid Interface Sci. 2022, 626, 283–295. [Google Scholar] [CrossRef]
  110. Song, X.; Jo, C.; Zhou, M. Enhanced tetracycline removal using membrane-like air-cathode with high flux and anti-fouling performance in flow-through electro-filtration system. Water Res. 2022, 224, 119054. [Google Scholar] [CrossRef]
  111. Xu, Q.; Liu, Y.; Wang, Y.; Song, Y.; Zhao, C.; Han, L. Synergistic oxidation-filtration process of electroactive peroxydisulfate with a cathodic composite CNT-PPy/PVDF ultrafiltration membrane. Water Res. 2022, 210, 117971. [Google Scholar] [CrossRef] [PubMed]
  112. Chen, C.C.; Xie, M.; Kong, L.S.; Lu, W.H.; Feng, Z.Y.; Zhan, J.H. Mn3O4 nanodots loaded g-C3N4 nanosheets for catalytic membrane degradation of organic contaminants. J. Hazard. Mater. 2020, 390, 122146. [Google Scholar] [CrossRef] [PubMed]
  113. Kang, J.; Zhang, H.Y.; Duan, X.G.; Sun, H.Q.; Tan, X.Y.; Liu, S.M.; Wang, S.B. Magnetic Ni-Co alloy encapsulated N-doped carbon nanotubes for catalytic membrane degradation of emerging contaminants. Chem. Eng. J. 2019, 362, 251–261. [Google Scholar] [CrossRef]
  114. Ye, J.; Dai, J.; Wang, L.; Li, C.; Yan, Y.; Yang, G. Investigation of catalytic self-cleaning process of multiple active species decorated macroporous PVDF membranes through peroxymonosulfate activation. J. Colloid Interface Sci. 2021, 586, 178–189. [Google Scholar] [CrossRef]
  115. Guo, R.N.; Li, Y.H.; Chen, Y.; Liu, Y.; Niu, B.H.; Gou, J.F.; Cheng, X.W. Efficient degradation of sulfamethoxazole by CoCu LDH composite membrane activating peroxymonosulfate with decreased metal ion leaching. Chem. Eng. J. 2021, 417, 127887. [Google Scholar] [CrossRef]
  116. Zhang, H.R.; Wang, X.S.; Li, Y.C.; Zuo, K.C.; Lyu, C. A novel MnOOH coated nylon membrane for efficient removal of 2,4-dichlorophenol through peroxymonosulfate activation. J. Hazard. Mater. 2021, 414, 125526. [Google Scholar] [CrossRef]
  117. Sun, M.H.; An, J.S.; Pan, Z.L.; Feng, G.Q.; Fan, X.F.; Song, C.W.; Wang, T.H. Enhanced organic wastewater treatment performance in electrochemical filtration process of coal-based carbon membrane via the simple Fe2+ addition. Sep. Purif. Technol. 2021, 276, 10. [Google Scholar] [CrossRef]
  118. Li, C.; Feng, G.Q.; Pan, Z.L.; Song, C.W.; Fan, X.F.; Tao, P.; Wang, T.H.; Shao, M.H.; Zhao, S.F. High-performance electrocatalytic microfiltration CuO/Carbon membrane by facile dynamic electrodeposition for small-sized organic pollutants removal. J. Membr. Sci. 2020, 601, 13. [Google Scholar] [CrossRef]
  119. Ma, T.; Liu, L.; Meng, B.; Gao, J.; Wang, S.; Liu, S. Heterogeneous activation of peroxymonosulfate via a Ag-La0.8Ca0.2Fe0.94O3−δ perovskite hollow fibre membrane reactor for dye degradation. Sep. Purif. Technol. 2019, 211, 298–302. [Google Scholar] [CrossRef]
  120. Zhou, J.; Zhao, Z.; Wang, Y.; Ding, Z.; Xu, X.; Peng, W.; Fan, J.; Zhou, X.; Liu, J. BiOCl0.875Br0.125/polydopamine functionalized PVDF membrane for highly efficient visible-light-driven photocatalytic degradation of roxarsone and simultaneous arsenic immobilization. Chem. Eng. J. 2020, 402, 126048. [Google Scholar] [CrossRef]
  121. Liu, F.; Yao, H.; Sun, S.; Tao, W.; Wei, T.; Sun, P. Photo-Fenton activation mechanism and antifouling performance of an FeOCl-coated ceramic membrane. Chem. Eng. J. 2020, 402, 125477. [Google Scholar] [CrossRef]
  122. Dai, Y.; Yao, Y.; Li, M.; Fang, X.; Shen, C.; Li, F.; Liu, Y. Carbon nanotube filter functionalized with MIL-101(Fe) for enhanced flow-through electro-Fenton. Environ. Res. 2022, 204, 112117. [Google Scholar] [CrossRef] [PubMed]
  123. Zhao, Z.; Zheng, W.; Jin, L.; Zhang, S.; You, S.; Liu, Y. Effective peroxymonosulfate activation using electrified nanohybrid filter towards one-step decontamination of roxarsone: Performance and mechanism. J. Environ. Chem. Eng. 2022, 10, 108643. [Google Scholar] [CrossRef]
  124. Sun, S.; Yao, H.; Fu, W.; Hua, L.; Zhang, G.; Zhang, W. Reactive Photo-Fenton ceramic membranes: Synthesis, characterization and antifouling performance. Water Res. 2018, 144, 690–698. [Google Scholar] [CrossRef] [PubMed]
  125. Xu, Y.L.; Yu, Y.L.; Yang, Y.; Sun, T.J.; Dong, S.J.; Yang, H.L.; Liu, Y.M.; Fan, X.F.; Song, C.W. Improved separation performance of carbon nanotube hollow fiber membrane by peroxydisulfate activation. Sep. Purif. Technol. 2021, 276, 8. [Google Scholar] [CrossRef]
  126. Ma, H.; Wang, G.; Miao, Z.; Dong, X.; Zhang, X. Integration of membrane filtration and peroxymonosulfate activation on CNT@nitrogen doped carbon/Al2O3 membrane for enhanced water treatment: Insight into the synergistic mechanism. Sep. Purif. Technol. 2020, 252, 117479. [Google Scholar] [CrossRef]
  127. Ma, H.; Wang, G.; Xu, Z.; Dong, X.; Zhang, X. Confining peroxymonosulfate activation in carbon nanotube intercalated nitrogen doped reduced graphene oxide membrane for enhanced water treatment: The role of nanoconfinement effect. J. Colloid Interface Sci. 2022, 608, 2740–2751. [Google Scholar] [CrossRef]
  128. Qu, W.; Wen, H.; Qu, X.; Guo, Y.; Hu, L.; Liu, W.; Tian, S.; He, C.; Shu, D. Enhanced Fenton-like catalysis for pollutants removal via MOF-derived CoxFe3-xO4 membrane: Oxygen vacancy-mediated mechanism. Chemosphere 2022, 303, 135301. [Google Scholar] [CrossRef]
  129. Zhang, L.F.; Zhang, Y.Q.; Wei, J.H.; Liu, W. Perovskite LaFexCo1−xO3−λ deposited SiO2 catalytic membrane for deeply cleaning wastewater. Chem. Eng. J. 2021, 403, 126386. [Google Scholar] [CrossRef]
  130. Wang, Y.; Zhang, Y.Q.; Wang, J.Q. Nano spinel CoFe2O4 deposited diatomite catalytic separation membrane for efficiently cleaning wastewater. J. Membr. Sci. 2020, 615, 118559. [Google Scholar] [CrossRef]
  131. Bao, Y.; Lee, W.J.; Wang, P.; Xing, J.; Liang, Y.N.; Lim, T.-T.; Hu, X. A novel molybdenum-based nanocrystal decorated ceramic membrane for organics degradation via catalytic wet air oxidation (CWAO) at ambient conditions. Catal. Today 2021, 364, 276–284. [Google Scholar] [CrossRef]
  132. Fan, Y.; Zhou, Y.; Feng, Y.; Wang, P.; Li, X.; Shih, K. Fabrication of reactive flat-sheet ceramic membranes for oxidative degradation of ofloxacin by peroxymonosulfate. J. Membr. Sci. 2020, 611, 118302. [Google Scholar] [CrossRef]
  133. Bao, Y.; Lim, T.-T.; Wang, R.; Webster, R.D.; Hu, X. Urea-assisted one-step synthesis of cobalt ferrite impregnated ceramic membrane for sulfamethoxazole degradation via peroxymonosulfate activation. Chem. Eng. J. 2018, 343, 737–747. [Google Scholar] [CrossRef]
  134. Bao, Y.; Lee, W.J.; Lim, T.-T.; Wang, R.; Hu, X. Pore-functionalized ceramic membrane with isotropically impregnated cobalt oxide for sulfamethoxazole degradation and membrane fouling elimination: Synergistic effect between catalytic oxidation and membrane separation. Appl. Catal. B Environ. 2019, 254, 37–46. [Google Scholar] [CrossRef]
  135. Bao, Y.; Oh, W.-D.; Lim, T.-T.; Wang, R.; Webster, R.D.; Hu, X. Surface-nucleated heterogeneous growth of zeolitic imidazolate framework—A unique precursor towards catalytic ceramic membranes: Synthesis, characterization and organics degradation. Chem. Eng. J. 2018, 353, 69–79. [Google Scholar] [CrossRef]
  136. Yang, Y.; Fu, W.; Chen, X.; Chen, L.; Hou, C.; Tang, T.; Zhang, X. Ceramic nanofiber membrane anchoring nanosized Mn2O3 catalytic ozonation of sulfamethoxazole in water. J. Hazard. Mater. 2022, 436, 129168. [Google Scholar] [CrossRef]
  137. Luo, J.; Chen, W.W.; Song, H.W.; Liu, J.R. Fabrication of hierarchical layer-by-layer membrane as the photocatalytic degradation of foulants and effective mitigation of membrane fouling for wastewater treatment. Sci. Total. Environ. 2020, 699, 12. [Google Scholar] [CrossRef]
  138. Li, M.; You, S.; Duan, X.; Liu, Y. Selective formation of reactive oxygen species in peroxymonosulfate activation by metal-organic framework-derived membranes: A defect engineering-dependent study. Appl. Catal. B Environ. 2022, 312, 121419. [Google Scholar] [CrossRef]
  139. Xu, L.; Niu, J.; Xie, H.; Ma, X.; Zhu, Y.; Crittenden, J. Effective degradation of aqueous carbamazepine on a novel blue-colored TiO2 nanotube arrays membrane filter anode. J. Hazard. Mater. 2021, 402, 123530. [Google Scholar] [CrossRef]
  140. Wang, L.; Wang, L.; Shi, Y.; Zhao, B.; Zhang, Z.; Ding, G.; Zhang, H. Blue TiO2 nanotube electrocatalytic membrane electrode for efficiency electrochemical degradation of organic pollutants. Chemosphere 2022, 306, 135628. [Google Scholar] [CrossRef]
  141. Sun, M.H.; Cui, M.R.; Wang, Y.B.; Fan, X.F.; Song, C.W. Enhanced Permeability and Removal Efficiency for Phenol and Perfluorooctane Sulphonate by a Multifunctional CNT/Al2O3 Membrane with Electrochemical Assistance. J. Nanosci. Nanotechnol. 2020, 20, 5951–5958. [Google Scholar] [CrossRef] [PubMed]
  142. He, Z.; Mahmud, S.; Yang, Y.; Zhu, L.; Zhao, Y.; Zeng, Q.; Xiong, Z.; Zhao, S. Polyvinylidene fluoride membrane functionalized with zero valent iron for highly efficient degradation of organic contaminants. Sep. Purif. Technol. 2020, 250, 117266. [Google Scholar] [CrossRef]
  143. Xu, H.; Cheng, W.; Chen, Z.; Zhai, X.; Ma, J.; Zhang, T. Selective oxidation of water pollutants by surface-complexed peroxymonosulfate during filtration with highly dispersed Co(II)-doped ceramic membrane. Chem. Eng. J. 2022, 448, 137686. [Google Scholar] [CrossRef]
  144. Sun, M.; Zucker, I.; Davenport, D.M.; Zhou, X.C.; Qu, J.H.; Elimelech, M. Reactive, Self-Cleaning Ultrafiltration Membrane Functionalized with Iron Oxychloride Nanocatalysts. Environ. Sci. Technol. 2018, 52, 8674–8683. [Google Scholar] [CrossRef]
  145. Li, C.; Feng, G.Q.; Pan, Z.L.; Sun, M.H.; Fan, X.F.; Song, C.W.; Wang, T.H. Facile morphology-controlled synthesis of ZnO electrocatalysts on coal-based carbon membrane for antibiotics wastewater treatment. J. Membr. Sci. 2021, 639, 12. [Google Scholar] [CrossRef]
  146. Aher, A.; Papp, J.; Colburn, A.; Wan, H.; Hatakeyama, E.; Prakash, P.; Weaver, B.; Bhattacharyya, D. Naphthenic acids removal from high TDS produced water by persulfate mediated iron oxide functionalized catalytic membrane, and by nanofiltration. Chem. Eng. J. 2017, 327, 573–583. [Google Scholar] [CrossRef]
  147. Zhang, J.; Yu, H.; Quan, X.; Chen, S.; Zhang, Y. Ceramic membrane separation coupled with catalytic ozonation for tertiary treatment of dyestuff wastewater in a pilot-scale study. Chem. Eng. J. 2016, 301, 19–26. [Google Scholar] [CrossRef]
  148. Chen, S.; Yu, J.; Wang, H.; Yu, H.; Quan, X. A pilot-scale coupling catalytic ozonation–membrane filtration system for recirculating aquaculture wastewater treatment. Desalination 2015, 363, 37–43. [Google Scholar] [CrossRef]
  149. Hu, J.; Fu, W.; Ni, F.; Zhang, X.; Yang, C.; Sang, J. An integrated process for the advanced treatment of hypersaline petrochemical wastewater: A pilot study. Water Res. 2020, 182, 116019. [Google Scholar] [CrossRef]
  150. Pei, S.; Wang, Y.; You, S.; Li, Z.; Ren, N. Electrochemical Removal of Chlorophenol Pollutants by Reactive Electrode Membranes: Scale-Up Strategy for Engineered Applications. Engineering 2022, 9, 77–84. [Google Scholar] [CrossRef]
  151. Lin, H.; Peng, H.; Feng, X.; Li, X.; Zhao, J.; Yang, K.; Liao, J.; Cheng, D.; Liu, X.; Lv, S.; et al. Energy-efficient for advanced oxidation of bio-treated landfill leachate effluent by reactive electrochemical membranes (REMs): Laboratory and pilot scale studies. Water Res. 2021, 190, 116790. [Google Scholar] [CrossRef] [PubMed]
  152. Plakas, K.V.; Sklari, S.D.; Yiankakis, D.A.; Sideropoulos, G.T.; Zaspalis, V.T.; Karabelas, A.J. Removal of organic micropollutants from drinking water by a novel electro-Fenton filter: Pilot-scale studies. Water Res. 2016, 91, 183–194. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Different experiment modes to conduct AOP with catalytic membranes: dead-end single pass (a), dead-end recirculation (b), cross-flow single pass (c), cross-flow recirculation (d), and batch (e).
Figure 1. Different experiment modes to conduct AOP with catalytic membranes: dead-end single pass (a), dead-end recirculation (b), cross-flow single pass (c), cross-flow recirculation (d), and batch (e).
Sustainability 15 07556 g001
Figure 2. Reaction mechanisms in catalytic membrane-based advanced oxidation process.
Figure 2. Reaction mechanisms in catalytic membrane-based advanced oxidation process.
Sustainability 15 07556 g002
Figure 3. Fabrication process of the Co@GAC/ceramic membrane module. Reproduced from Ref. [29] with permission from Elsevier.
Figure 3. Fabrication process of the Co@GAC/ceramic membrane module. Reproduced from Ref. [29] with permission from Elsevier.
Sustainability 15 07556 g003
Figure 4. Janus electrified ceramic membrane for degradation of MB. Reproduced from Ref. [60] with permission from Elsevier.
Figure 4. Janus electrified ceramic membrane for degradation of MB. Reproduced from Ref. [60] with permission from Elsevier.
Sustainability 15 07556 g004
Figure 5. Different procedures for regeneration of catalytic membranes by washing: filtration (a), back washing (b), disassembling and cleaning the membrane (c) and disassembling the membrane, separating, and cleaning the catalyst followed by reloading the catalyst (d).
Figure 5. Different procedures for regeneration of catalytic membranes by washing: filtration (a), back washing (b), disassembling and cleaning the membrane (c) and disassembling the membrane, separating, and cleaning the catalyst followed by reloading the catalyst (d).
Sustainability 15 07556 g005
Figure 6. Regeneration of Fe/PDA/PEI/PVDF membrane by chemical reduction. Reproduced from Ref. [142] with permission from Elsevier.
Figure 6. Regeneration of Fe/PDA/PEI/PVDF membrane by chemical reduction. Reproduced from Ref. [142] with permission from Elsevier.
Sustainability 15 07556 g006
Figure 7. Half violin plots for the number of cycles (a) and operation time (b) employed in the literature for the performance stability evaluation of catalytic membranes.
Figure 7. Half violin plots for the number of cycles (a) and operation time (b) employed in the literature for the performance stability evaluation of catalytic membranes.
Sustainability 15 07556 g007
Figure 8. Pilot-scale membrane modules designed by the scaling up strategy (a,b) and the numbering up strategy (c). Reproduced from Refs. [147,148,151] with permission from Elsevier.
Figure 8. Pilot-scale membrane modules designed by the scaling up strategy (a,b) and the numbering up strategy (c). Reproduced from Refs. [147,148,151] with permission from Elsevier.
Sustainability 15 07556 g008
Table 1. Performance stability of catalytic membranes tested by multi-cycle experiments.
Table 1. Performance stability of catalytic membranes tested by multi-cycle experiments.
AOP TypeSubstrate MembraneCatalystFabrication MethodMembrane TypeFiltration
Mode
Operation ModeTarget
Pollutant
C0Flux
(L/(m2·h))
R1Number of CyclesηRef.
PMS PVDFCoAl-LDHfiltrationflat-sheetdead-endrecirculationRA2.5 mg/L80.394%10~100%[3]
PMSPVDFCo2+/Mxenefiltrationflat-sheetdead-endsingle passTC10 mg/L65.798.2%384%[4]
PMSPVDFFeOCl/MoS2filtrationflat-sheetdead-endsingle passRhB10 mg/L40070%587%[13]
PMSPVDFMnO2/CNTfiltrationflat-sheetdead-endsingle passRhB50 mg/L248.51100%8100%[14]
PMSPVDFFeCoS@N-rGOfiltrationflat-sheetdead-endsingle passSMX30 mg/L632.1299.2%591%[15]
PMSPVDFCo@N-Cphase inversionflat-sheetcross-flowsingle passTC20 mg/L636.099.3%591%[16]
PMSPTFECo3O4@NCNTs/g-CNfiltrationflat-sheetdead-endrecirculationSMX10 mg/L276.8698.9%589%[17]
PMSPTFEFeCo@GCTs/GOfiltrationflat-sheetdead-endrecirculationSMX10 mg/L487.398.4%591%[18]
PDSPTFEN-doped GOfiltrationflat-sheetdead-endrecirculationphenol
oxalic acid
50 mg/L
90 mg/L
112594%
100%
329%
80%
[19]
PMSPANFe-doped LaCoO3phase inversionflat-sheetcross flowsingle passTC10 mg/L220.799%599%[20]
PMSPSFSCCMphase inversionflat-sheetcross-flowsingle passRhB20 mg/L126.90100%5~100%[21]
PMSnylonFe2O3@CNTfiltrationflat-sheetdead-endsingle passTC0.04 mmol/L16.396.1%591%[22]
PDSnylonNG/rGO/CNTfiltrationflat-sheetdead-endrecirculationSMX0.5 mg/L46.1594.8%585%[23]
PMSceramic membraneCuOphase-inversion and liquid-phase sinteringhollow fibercross-flowsingle passRhB20 mg/L50081.5%593%[24]
PMSceramic membraneCuOimpregnation and calcinationhollow fibercross-flowsingle passSDZ1 mg/L7091.7%593%[25]
PMSceramic membraneCuOimpregnation and calcinationhollow fibercross-flowsingle passBPA10 mg/L7091.4%5~100%[26]
PMSceramic membraneCo3O4@CNTfiltrationflat-sheetdead-endsingle passMB15 mg/L22484.4%683%[27]
PMSceramic membraneMn2O3spraying and calcinatingflat-sheetdead-endsingle passacetaminophen1 mg/L6099%10~100%[28]
PMSceramic membraneCo@GACfilling the channelsflat-sheetdead-endsingle passBPA10 mg/L3592.8%697%[29]
H2O2FEPFe2O3thermoforming methodflat-sheetcross-flowsingle passMB40 mg/L1599.69%373%[30]
H2O2PANgoethite/maleate ferroxanephase inversionflat-sheetdead-endsingle passamoxicillin105 mg/L23.286.3%/92.3%497%/98%[31]
O3ceramic membraneCuOfiltration and
calcination
tubulardead-endrecirculation1,4-dioxane200 mg/L363.8146%498%[32]
O3ceramic membraneMnCe oxidefiltration, in situ reaction and calcinationflat-sheetdead-endrecirculationatrazine1 mg/L6090%594%[33]
O3ceramic membraneCuMn2O4impregnation and calcinationtubularcross-flowsingle passBP-32 mg/L63.15775.3%3091%[34]
O3ceramic membraneN-rGOpneumatic methodtubularcross-flowsingle passbenzotriazole0.084 mol/L660.86100%1892%[35]
PhotoPVDF
PTFE
TiO2impregnation and hydrothermal treatmentflat-sheetdead-endrecirculationdiclofenac/
ethinylestradiol
1.01 µmol/L1144/133192%, 94%;
89%, 92%
3~100%, 96%;
67%, 83%
[36]
Photoceramic membraneTiO2impregnation and calcinationflat-sheetcross-flowrecirculationMB0.083 mmol/L12580%4~100%[37]
Photoceramic membraneGO–TiO2filtration and
calcination
flat-sheetdead-end
cross-flow
recirculation
recirculation
naproxen
diclofenac
CBZ
50 ppb92.6100%, 100%, 90%;
100%, 100%, 30%
7
11
[38]
Photocellulose acetateGO–TiO2filtrationflat-sheetdead-endsingle passCR50 mg/L483.597%398%[39]
PhotoPSFGO–N-TiO2filtrationflat-sheetdead-endsingle passMB50 mg/L7050%866%[40]
PhotoNone:
free standing
rGO@TiO2electro-spinning and calcinationflat-sheetcross-flowrecirculationpropranolol2 mg/L342.8676.1%10~100%[41]
PhotoPVDFZnIn2S4filtrationflat-sheetdead-endrecirculationfluvastatin10 mg/L27097.19%694%[42]
Photo + H2O2ceramic membraneα-Fe2O3spraying and calcinatingflat-sheetdead-endrecirculationTC20 mg/L158.7382%588%[43]
Photo + H2O2PANβ-FeOOHimpregnation and in situ mineralizationflat-sheetcross-flowsingle pasmethyl blue20 mg/L11–1399.8%597%[44]
Photo + H2O2air-laid paperβ-FeOOHimpregnation and in situ mineralizationflat-sheetdead-endsingle passTC40 mg/L60099.5%598%[45]
Photo + H2O2PSFferrocenelayer-by-layer interfacial polymerization and ion-exchangeflat-sheetdead-endsingle passBPS20 mg/L4.299.5%3[46]
Photo + PMSPPZIF-67in situ synthesisflat-sheetcross-flowsingle passMB
MO
20 mg/L
20 mg/L
216.8TOC removal:
74.2%
83.5%
599%
98%
[47]
Photo + PMSPTFEα-Fe2O3/bacterial cellulosefiltrationflat-sheetdead-endsingle passRhB20 mg/L4.93100%597%[48]
Photo + PMSPVDF CoFe2O4/carbon nanofiber phase inversionflat-sheetdead-endrecirculationberberine10 mg/L36893.2%497%[49]
Photo + PDSPVDFflat-sheetdead-endsingle passofloxacin5 mg/L54%1026%[50]
Electroporous titanium plate (anode)
titanium mesh (cathode)
Pd (anode)
–(cathode)
spray coating and calcinationflat-sheetdead-endsingle passSMX100 μg/L500096.3%10~100%[51]
ElectroRuO2 plated titanic mesh (anode)
RuO2 plated titanic mesh (cathode)
graphene/SnO2/carbon nanofibers (anode)
–(cathode)
electrospinning and carbonizationflat-sheetdead-endrecirculationSMX15 mg/L216.485%694%[52]
Electroporous Ti plate (anode)
steel mesh (cathode)
blue TiO2 (anode)
–(cathode)
electrooxidation, calcination and electroreductionflat-sheetdead-endrecirculationtriclosan10 mg/L358298.5%4~100%[53]
Electrostainless steel network/PTFE (anode)
stainless steel network (cathode)
CNT (anode)
–(cathode)
filtrationflat-sheetdead-endsingle passSMX
ciprofloxacin
amoxicillin
50 mg/L127.4890%
76%
98%
472%
79%
92%
[54]
Electrotitanium ring/PTFE (anode)
stainless steel (cathode)
single cobalt atom and nitrogen atom co-doped graphene (anode)
–(cathode)
filtrationflat-sheetdead-endsingle passMB5 mg/L15.2999.5%590%[55]
Electrostainless-steel mesh/PVDF (anode)
titanium foil (cathode)
nano porous carbon/CuFeO2 (anode)
–(cathode)
phase inversionflat-sheetdead-endsingle passMO10 mg/L12598.72%469%[56]
Electroouter CNT layer (anode)
inner CNT layer (cathode)
–(anode)
–(cathode)
layer-by-layer coatinghollow fiberdead-endsingle passMC-LR0.5 mg/L50099.8%3~100%[57]
Electroceramic membrane (anode)/Ti mesh
titanium mesh (cathode)
TiO2/SnO2-Sb (anode)
–(cathode)
adhesive assembleflat-sheetdead-endsingle pass2,4-D1 mg/L27862.4%8~100%[58]
Electro + PMStitanium mesh (anode)
PTFE (cathode)
–(anode)
MnFe2O4-rGO (cathode)
filtrationflat-sheetdead-endrecirculationOTC10 mg/L553.688.7%586.1%[59]
Electro + PMSceramic
membrane (anode and cathode)
Pd (anode and cathode)confocal magnetron co−sputteringflat-sheetcross flowsingle passMB10 μmol/L461.3494.5%895%[60]
Electro + Photo + PMSstainless steel (anode)
titanium plate (cathode)
TiO2-ZnO (anode)
–(cathode)
atomic layer depositionflat-sheetcross flowsingle passatrazine
MB
5 mg/L
5 mg/L
59.9100%590%
85%
[61]
Abbreviations: BPA: Bisphenol A; BPS: Bisphenol S; BP-3: Benzophenone-3; CBZ: Carbamazepine; CR: Congo red; CNT: Carbon nanotube; OTC: Oxytetracycline; TOC: Total organic carbon; GAC: granular activated carbon; LDH: layered double hydroxide; MB: Methylene blue; MC-LR: Microcystin-LR; MO: Methyl orange; N-rGO: Nitrogen-doped reduced graphene oxide; PAN: Polyacrylonitrile; PDS: Peroxydisulfate; PMS: Peroxymonosulfate; PP: Polypropylene; PSF: Polysulfone; PTFE: Poly(tetrafuoroethylene-co-hexafuoropropylene); PTFE: Polytetrafluoroethylene; PVDF: Polyvinylidene fluoride; RA: Ranitidine; RhB: Rhodamine B; SDZ: Sulfadiazine; SMX: Sulfamethoxazole; TC: Tetracycline; ZIF-67: Zeolitic imidazole framework-67; 2,4-D: 2,4-Dichlorophenoxyacetic acid.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shi, Y.; Zhang, T.; Chang, Q.; Ma, C.; Yang, Y.; Wang, S.; Pan, Z.; Sun, Y.; Ding, G. Performance Stability and Regeneration Property of Catalytic Membranes Coupled with Advanced Oxidation Process: A Comprehensive Review. Sustainability 2023, 15, 7556. https://doi.org/10.3390/su15097556

AMA Style

Shi Y, Zhang T, Chang Q, Ma C, Yang Y, Wang S, Pan Z, Sun Y, Ding G. Performance Stability and Regeneration Property of Catalytic Membranes Coupled with Advanced Oxidation Process: A Comprehensive Review. Sustainability. 2023; 15(9):7556. https://doi.org/10.3390/su15097556

Chicago/Turabian Style

Shi, Yawei, Tongwen Zhang, Qian Chang, Chang Ma, Yao Yang, Songbo Wang, Zonglin Pan, Ya Sun, and Guanghui Ding. 2023. "Performance Stability and Regeneration Property of Catalytic Membranes Coupled with Advanced Oxidation Process: A Comprehensive Review" Sustainability 15, no. 9: 7556. https://doi.org/10.3390/su15097556

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