Next Article in Journal
Integration of Ultra- and Nanofiltration for Potato Processing Water (PPW) Treatment in a Circular Water Recovery System
Next Article in Special Issue
Using Tannic-Acid-Based Complex to Modify Polyacrylonitrile Hollow Fiber Membrane for Efficient Oil-In-Water Separation
Previous Article in Journal
Dyeable Hydrophilic Surface Modification for PTFE Substrates by Surface Fluorination
Previous Article in Special Issue
Environmental Friendly Fabrication of Porous Cement Membranes via Reusable Camphene-Based Freeze-Casting Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances on the Fabrication of Antifouling Phase-Inversion Membranes by Physical Blending Modification Method

by
Tesfaye Abebe Geleta
,
Irish Valerie Maggay
,
Yung Chang
* and
Antoine Venault
*
R&D Center for Membrane Technology, Department of Chemical Engineering, Chung Yuan Christian University, Chung-Li 32023, Taiwan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Membranes 2023, 13(1), 58; https://doi.org/10.3390/membranes13010058
Submission received: 4 October 2022 / Revised: 16 December 2022 / Accepted: 19 December 2022 / Published: 2 January 2023
(This article belongs to the Special Issue Advances in Porous and Dense Membranes: Fabrication and Applications)

Abstract

:
Membrane technology is an essential tool for water treatment and biomedical applications. Despite their extensive use in these fields, polymeric-based membranes still face several challenges, including instability, low mechanical strength, and propensity to fouling. The latter point has attracted the attention of numerous teams worldwide developing antifouling materials for membranes and interfaces. A convenient method to prepare antifouling membranes is via physical blending (or simply blending), which is a one-step method that consists of mixing the main matrix polymer and the antifouling material prior to casting and film formation by a phase inversion process. This review focuses on the recent development (past 10 years) of antifouling membranes via this method and uses different phase-inversion processes including liquid-induced phase separation, vapor induced phase separation, and thermally induced phase separation. Antifouling materials used in these recent studies including polymers, metals, ceramics, and carbon-based and porous nanomaterials are also surveyed. Furthermore, the assessment of antifouling properties and performances are extensively summarized. Finally, we conclude this review with a list of technical and scientific challenges that still need to be overcome to improve the functional properties and widen the range of applications of antifouling membranes prepared by blending modification.

1. Introduction

1.1. Membrane Technology

The global water shortage has resulted from population growth and industrialization and is a key challenge that humanity faces in the 21st century. Continuous progress on wastewater treatment technologies has been made to improve water reuse and reduce water shortages. Among them, the development of thin-layer membranes, working as selective agents between two phases and regulating the permeability of substances, has made tremendous progress.
Pressure-driven membranes are commonly categorized into four groups depending on their pore size and the type of filtration involved: microfiltration (MF), ultrafiltration (UF), nanofiltration (NF), and reverse osmosis (RO) membranes [1,2,3,4]. Owing to their ease of fabrication and operation, high selectivity rates, energy savings, and high-level modularity, membranes have made significant contributions to address the challenges associated with resource and energy scarcity and have been increasingly used in applications such as desalination, water treatment, and food and pharmaceutical manufacturing [5,6,7]. Myriads of membranes used in wastewater treatment are formed from either organic or inorganic membranes [8,9]. The prior is predominantly based on a variety of polymers such as polyamide (PA), poly(vinylidene fluoride) (PVDF), polysulfone (PS), polyethersulfone (PES), polypropylene (PP), polyimide (PI), poly(ether imide) (PEI), poly(tetrafluoroethylene) (PTFE), polycarbonate (PC), polyetherketone (PEEK), polypiperazine (PPZ), polyacrylonitrile (PAN), cellulose acetate (CA), and so on. Polymeric membranes offer facile and straightforward fabrication methods for a wide range of pore sizes, and simple modification techniques that allow for cheap and easy industry scalability [10,11]. However, these materials are prone to fouling and susceptible to extreme operating conditions such as pH, temperature, pressure, etc. On the other hand, inorganic membranes are tailored from ceramics or metals which are characterized by having stronger resistance against harsh operating conditions, although their widespread applications are greatly hampered by the difficulty of their processability, and fabrication techniques, and expensive cost [12]. Furthermore, inorganic membranes are mostly limited to MF and UF applications to date [9]. Hence, the polymeric membrane is still considered the primary choice for the vast majority of the studies on membrane technology.

1.2. Preparation of Porous Membranes via the Phase-Inversion Processes

Phase-inversion processes are well-established techniques for fabricating porous polymeric membranes. In general, phase-inversion is a demixing process that converts a homogeneous polymer casting liquid solution to a solid film in a controlled manner (Figure 1) [13]. The final morphology of the membranes can be determined by the demixing process which is greatly influenced by the thermodynamics, viscoelasticity of the casting solution, and kinetics of the solvent/non-solvent upon precipitation.
Phase-inversion processes can be induced by the utilization of a non-solvent (Non-solvent Induced Phase Separation, NIPS) which can be liquid (wet-immersion or Liquid-Induced Phase Separation, LIPS) or vapor (Vapor-Induced Phase Separation, VIPS). It can also be triggered by a rapid change in temperature (Temperature-Induced Phase Separation process, TIPS).
In LIPS, after casting a homogeneous solution on an adequate substrate, it is immersed in a non-solvent coagulation bath for a desired time. As a result of the exchange of the solvent and the non-solvent, precipitation occurs, and a membrane film is produced. The excess of solvent is then removed by extensive washing of the membrane in water [14,15,16]. UF membranes are usually prepared by this method. In the VIPS process which leads to MF membranes, the homogeneous solution is exposed to vapors of the non-solvent under controlled conditions of humidity, temperature, and time. It is sometimes combined with LIPS to control the morphology of the top surface and increase its the porosity [17]. In TIPS, a rapid change in temperature of the polymeric system caused by immersion of a hot polymer/diluent system in a cold/cool immersion bath induces phase-inversion. In NIPS processes (LIPS and VIPS), the nature of the solvent/non-solvent system and the concentration and composition of the polymer solution are key aspects that impact the phase-inversion approach for membrane production [1,18,19]. In VIPS, one has to consider the impact of the process parameters (example: relative humidity, exposure time to vapors). This process has been extensively reviewed recently and one is invited to refer to the study of Ismail et al. for relevant information [20]. As for the TIPS process, key parameters to consider are the temperatures of the dope and of the coagulation bath, and, if hollow fibers are produced, the dope flow rate and the air gap between the spinneret and the coagulation bath [21,22,23,24].
Phase-inversion methods require relatively simple equipment and can be readily implemented. DI water is frequently used as a non-solvent medium (in LIPS, VIPS) or a coagulation medium (in TIPS), which makes these processes relatively cost-effective. Furthermore, they are applicable to a large variety of materials. The major current drawback is the toxicity of the solvents used. The addition of organic/inorganic additives for antifouling purpose may also cause solubility issues, resulting in the production of inhomogeneous dispersions in the matrix membrane which in turn diminishes the stability of the membranes and ultimately their antifouling performances.

1.3. Membrane Fouling and Fouling Mitigation Techniques

Fouling, the “Achilles’ heel” of membrane technology, is a pervasive issue that if not prevented or acted upon, could leave many water treatments in futility. Membrane fouling is caused by the deposition or adsorption of pollutants or foulants on the surface of the membranes and/or penetration into the pores [25], which then leads to a series of operational and performance issues such as flux decline, decreased permeate purity, and overall poor separation performance. In general, foulants are categorized into inorganic foulants, organic foulants, biofoulants, and composite foulants. Inorganic foulants include inorganic scales, colloidal inorganic substances, etc. Organic foulants encompass oils, organic solvents, organic dyes, natural organic matter (NOM), and so on. Biofoulants, on the other hand, consist of one or more types of biomolecules or cells. Lastly, composite foulants are the combination of the inorganic, organic and biofouling foulants [26]. Although inorganic scaling/fouling poses serious threat to membrane performance, it is inhibited through pre-treatment of the feed or optimization of operating conditions [25]. Therefore, studies on the development of antifouling membranes are geared towards mitigation of organic foulants and biofoulants. Jiang and colleagues published in 2015 an interesting review that tackled the fouling mechanisms of various foulants [25], which helped inspire various antifouling studies.
Proteins, commonly used biofoulants for many studies, exist as amphiphilic compounds due to the existence of both hydrophilic polar groups and hydrophobic non-polar groups. Protein fouling on the membrane surface is generated by protein–membrane interactions prompted by hydrophobic interaction by the protein and membrane. In an aqueous environment, the hydrophilic groups of the proteins can bind with the water molecules loosely adhered on the surface of the membrane via hydrogen bonding and electrostatic interactions. These loosely adhered water molecules could be easily displaced, increasing the chances of the hydrophobic portions of the proteins to form hydrophobic–hydrophobic interactions with the membrane. In addition, if the pH of the solution is equal to the iso-electric point of the proteins, it becomes strongly hydrophobic, in turn facilitating the hydrophobic interactions of protein–membrane. Moreover, proteins tend to aggregate at their iso-electric point, enhancing the protein–protein hydrophobic interaction. Hence, this cascades into enhanced fouling as the aggregated proteins form a hydrophobic interaction with the initially adhered proteins on the membrane surface [26,27].
Natural organic matter (NOM) such as humic acid (HA) and sodium alginate (SA) consist of hydrophobic aliphatic and aromatic moieties, allowing these types of foulants to be adsorbed on the membrane surface through hydrophobic interactions. Organic dyes (e.g., Congo red, methyl blue, rhodamine B, malachite green, methyl orange, rose Bengal dye, acid black 210, and so on [28,29,30,31]) are interesting foulants due to their complex structure, and electrical charge. In addition, dye effluents have high salinity due to the addition of salts (common practice by textile industries to enhance dye uptake) [28,32], which makes the fouling mechanism by organic dyes complex. Distinct from proteins, dye molecules aggregate through ionic association, hydrophobic interaction, hydrogen bonding to form hydrated ions and spherical or flake aggregates in an aqueous solution. Moreover, mutual association and aggregation between adjacent dye molecules on the membrane surface could be induced through hydrophobic interactions. Due to the charged nature of the dye molecules, it can induce concentration polarization, affecting the flux. Additionally, the adsorbed dye can form strong chemical bonds with the groups on the membranes surface (or within the membrane pores) [29]. Oil foulants can exist as free oil, emulsified oil and/or dissolved oil which could form fouling on the membrane surface through hydrophobic–hydrophobic interactions. As a result of the high tendency of the oil droplets to coalesce, this can form an oil layer on the surface of the membrane which makes the penetration of water in the succeeding cycles more challenging. Also, in surfactant stabilized emulsifiers, the surfactants used are amphiphilic which could result in penetration and clogging of the membrane pores [12].
Hence, polymeric membranes must be conferred with antifouling properties that are tailored to alleviate various interactions between the foulant and the membrane surface, and the interaction that transpires between the foulant and adsorbed foulant on the membrane surface. These could be achieved through increasing surface hydrophilicity, enhancing the surface charge, and decreasing the surface roughness of the membranes [33].

1.4. Modification Techniques for Antifouling

In order to mitigate fouling and maintain high permeability/separation performance, porous membranes can be modified via several techniques including grafting polymerization, layer-by-layer deposition, dip-coating, spray coating, spin coating, etc. [5,19,34,35,36]. These techniques modify the top surface of the membrane and involve a post-treatment step [37]. On the other hand, physical blending or simply “blending” involve the process of incorporating the matrix polymer with inorganic nanofillers or hydrophilic and/or amphiphilic polymers in the casting solution before applying a phase-inversion process (LIPS, VIPS or TIPS). This modification technique is a one-step process since the membrane is formed/modified all at once. Many recent reports have shown the suitability of this technique to improve the hydrophilic properties of membranes, resulting in fouling mitigation [34,38,39,40]. Researchers have worked on the modification of polymeric membranes using various additives including polymer-based materials such hydrophilic or amphiphilic polymers, biopolymers and nature-derived materials [41,42,43,44,45,46], metal-based nanomaterials [47,48,49,50,51,52,53,54,55,56,57,58], ceramic-based nanomaterials [59,60,61,62,63,64,65,66], carbon allotropes (graphene-based materials, carbon nanotubes, etc.), layered and porous nanomaterials (e.g., zeolites, metal-organic frameworks (MOFs), etc.) [67,68,69,70,71,72,73], and hybrid nanomaterials (which includes the combination of two or more materials) [74,75,76,77,78,79] in order to develop the hydrophilic characteristics and antifouling capability of membranes necessary for water treatment or their application in the biomedical field. Distinct from other methods, blending modification does not require any additional steps during the fabrication of polymeric membranes or mixed matrix membranes (MMMs), making it a relatively simple and fast method; hence, considerable attention has been given to this over the past decade as illustrated by the number of research publications in Figure 2. A schematic diagram representing each material groups are shown in Figure 3.

1.5. Objectives of the Article

This review article explores the published recent advances on the theme of antifouling membranes focused on one-step formation/modification of membranes or blending modification by phase-inversion. As a review on biofouling of water treatment membranes was published in this journal 10 years ago [80], we tried to solely focus on the period 2012–2022. We begin with exploring the antifouling materials belonging to the class of polymers/organic, metals, ceramics, carbon allotropes and porous nanomaterials, and hybrid nanomaterials. Then, the effects of antifouling materials on membrane morphology are investigated. In addition, the methods used to assess the antifouling properties of membranes are discussed, as well as the performances of the modified membranes, in comparison with the pristine membranes. Finally, future perspectives are proposed.

2. Antifouling Materials for Polymeric Membranes

The key problem in wastewater treatment utilizing polymer-based membranes is developing low-cost and high-performance antifouling membranes. Membrane fouling lowers the lifespan of membranes [1,81,82,83]. Real-world membrane use requires the creation of cost-effective wettable polymeric membranes. The choice of membrane composition and the control of morphological features both have a substantial effect on antifouling properties as they impact the membrane wettability [63,83].
When weak interactions between the foulant and the membrane take place, reversible fouling occurs, which may be readily eliminated by a physical cleaning of the membrane. However, strong interactions between the foulant and the membrane surface arise in irreversible fouling, resulting in the production of a persistent fouling layer that cannot be cleaned by ordinary procedures. Biofouling is often regarded as irreversible. For example, bacteria that stick onto the surface of the membrane can grow and reproduce, eventually covering the whole membrane surface. How to enhance the membrane’s antifouling capability by reducing interactions between the material and bacteria/cells has become a crucial topic in membranes applied to water treatment and in the biomedical field and requires the use of antifouling materials. The high degree of surface hydrophilicity imparted by a suitable surface chemistry result in efficient fouling mitigation in aqueous media [1,7,15,84,85]. Research teams worldwide have investigated the blending of polymers traditionally used for membrane fabrication with various types of antifouling materials belonging to different classes. What follows is a presentation of these additives, which are categorized into five types according to their intrinsic characteristics: polymer/organic-based, metal-based, ceramic-based, carbon allotropes and porous nanomaterials, and hybrid nanomaterials, as represented in Figure 3. We also present a discussion of their effect on the performance of the membranes.

2.1. Polymer/Organic-Based Additives

The majority of polymer materials used for the large-scale production of porous membranes applied in wastewater treatment are hydrophobic. So, the use of hydrophilic or amphiphilic polymers has been investigated. Recent works published in the last 10 years starring these polymeric materials are summarized in Table 1 and introduced as follows. It is important to note that hydrophilic polymers such as poly(ethylene glycol) or PEG, and poly(vinylpyrrolidone) or PVP are commonly used as porogens or pore formers; in this review, we investigated studies that utilized these polymers as antifouling modifiers not as porogens.

2.1.1. Poly(ethylene glycol) and Its Derivatives

Poly(ethylene glycol) (PEG) exists in a variety of molecular weights (Mws), for example PEG 200, PEG 400, PEG 6000 or PEG 20,000 where the number following “PEG” refers to the average molecular mass. Ma et al. studied the influence of PEG Mw and content on the properties of polysulfone (PSf) membranes [42]. Increasing the PEG Mw significantly enhanced the pure water flux (PWF) while decreasing the rejection performances, as it affected the pore size. From there, they selected PEG400 and showed that increasing its content led to higher permeability still maintaining bovine serum albumin (BSA) rejection, which could be explained by the presence of more pores of similar size, that is, by the pore-forming effect of PEG. Note though that boosting the PEG content also reduced the mechanical properties of the membranes due to the increased porosity. Thus, not only the molecular weight but also the PEG content had to be optimized.
As a result of its hydrophilic nature, PEG may easily leach away during membrane washing or even during filtration, resulting in reduced hydrophilicity and antifouling performance [1,87,89]. So, the emphasis has been put in the past decade on the design of amphiphilic additives composed of both hydrophobic and hydrophilic units. The hydrophobic units ensure stability of the antifouling material in the membrane system, and so, they are permitted to maintain for a longer time period the antifouling property [88]. The hydrophilic segment tends to be oriented towards the membrane surface/pore surface during phase-inversion which enables the increase in the membrane hydrophilicity and its antifouling ability. Copolymers of PEG, poly(ethylene glycol) methyl ether (PEGME), and poly(ethylene glycol) methyl ether methacrylate (PEGMA) have been reported [41,86,88,90,92,93,153,154]. Differing from PEG, PEGME and PEGMA derivatives have a -CH3 terminal which decreases the chance of leaching out [154]. However, this cannot be guaranteed over a long period of time. Hence, strategies have been developed in order to resolve this issue. In some cases, the hydrophilic segment was grafted to the matrix polymer, and the resulting copolymer blended with the pristine polymer. A major advantage of this technique is the compatibility between the materials and the strength of the hydrophobic interactions established between the copolymer and the membrane main polymer. For instance, Wu et al. first grafted PEGMA to PVDF, and then blended the grafted copolymer with pristine PVDF [92,93]. Similarly, PEGMA was grafted to PVC and the resulting material blended with PVC to form membranes by LIPS [91]. The use of copolymers requires an optimization step in order to maximize stability and antifouling. Too many hydrophobic segments lead to high stability but poor antifouling property. Conversely, too many hydrophilic segments lead to excellent but not sustainable fouling mitigation. Gao et al. introduced an amphiphilic random comb copolymer containing PEG and poly(dimethylsiloxane) (PDMS) which blended with PES. PEG-r-PDMS was prepared using the free radical polymerization process [88]. They observed that the increased amount of PDMS in the amphiphilic copolymer PEG-r-PDMS resulted in higher WCA, but stability was achieved. In our previous work, we investigated the effects of tailoring both the composition and molecular weights of PS-r-PEGMA copolymers [90], and we found that longer chains led to better antifouling properties, not because of a higher overall hydration level, but because of a tighter hydration layer.

2.1.2. Poly(vinylpyrrolidone)

PVP has similar properties to PEG, as both are hydrophilic in nature and used as pore-forming agents in polymeric membranes. Although there are stability concerns, PVP still remains an important material to tune the surface properties of membranes prepared by blending modification. For instance, it was recently used by Son et al. to tune the permeation and antifouling properties of PES membranes [43]. They specifically laid the focus on high PVP concentrations but concluded that large additive amounts (15 wt%, 20 wt% in the casting solution) were not desired as they induced the formation of a dense layer, decreasing water permeability. Instead, 10 wt% PVP in the casting solution seemed to lead to the best combination of membrane permeability, rejection, and antifouling property. Kanagaraj et al. investigated the formation of polyetherimide (PEI) with various concentrations of PVP (0–8 wt%) and reported that the larger concentration permitted to maximize the PWF (147.1 L/m2 h) and minimize protein adsorption [94]. However, it was at the expense of the membrane’s tensile strength and of protein rejection due to the pore forming effect of PVP in this concentration range. Similar to PEG, PVP is available over a large range of molecular weights. For a given additive content, it is clear that the choice of a larger Mw will result in larger casting solution viscosity, which should help in decreasing mass transfer rates during phase-inversion and lock the additive in the main matrix polymer. Vatsha et al. [96] investigated the effect of using high concentrations of PVP 40 k as modifier for PES. They reported that at 10 wt% of PVP (for 16 wt% and 18 wt% PES), significantly increased the hydrophilicity, but induced the formation of dense surface structure. These reports show that by using larger amounts of PVP, denser structures are obtained, as the total volume fraction of solid content increases. It suggests that PVP can still be retained in the matrix polymer and can be of interest to increase the surface free energy of the membrane through both a change in surface chemistry and surface structure (less pores), which would benefit fouling mitigation, although it may reduce the membrane permeability. Similarly, Vatsha et al. used similar logic, and finally, PVP has been recently combined with polydiol terephthalate to form an amphiphilic triblock copolymer, that resulted in the formation of stable antifouling PSf membranes [95].

2.1.3. Cellulose Nanocrystals

Cellulose nanocrystals (CNC) additives are biocompatible, renewable, environmentally friendly, and they exhibit excellent mechanical strength. Due to their numerous hydroxyl groups and unique mechanical strength, CNC additives have received increasing attention to improve the surface hydrophilicity, permeability, antifouling properties, and mechanical strength of blended membranes [99,100,101]. The presence of hydroxyl groups facilitates the formation of hydrogen-bonding networks, hence the hydration of films. Zhang et al. incorporated CNC into PES UF membranes [102]. They blended different amounts of CNC (0.1–5.0 wt%). The virgin PES membrane displayed a WCA of 66°, which dropped to 43° for the membrane containing 5.0 wt% CNC. Meanwhile, BSA rejection was increased from 93% (virgin membrane) to 97% due to the formation of smaller pores, and the flux recovery ration (FRR) enhanced (from 51 to 90%). Nevertheless, the blended membranes’ tensile strength and elongation at break both decreased dramatically with the CNC loading, which may result from heterogeneous dispersion of the nanocrystals into the polymeric matrix. Similarly, Zhou et al. reported that the agglomeration of the nanocrystals could occur past a certain loading, leading to the formation of “holes” (large pores) seen on the membrane surface, responsible for a decrease in the mechanical properties [101]. Lv et al. explored the incorporation of CNC into PVDF membranes and reported the significant effect of the CNC loading on the membrane structure [98]. While low loadings (0.7–1.4 wt%) were associated with homogeneous dispersion of the CNC and denser surfaces, higher loadings (1.4–2.8 wt%) seemed to arise in the formation of more surface pores, until large defects were observed for the highest loading (4.2 wt%). Similar reports [102,103] suggest the need for adjusting the CNC content in order to take full advantage of its effect on both surface hydrophilicity and mechanical reinforcement of the matrix. Clearly, agglomeration readily occurs past a concentration threshold, which can severely decrease the membrane mechanical strength while homogeneous dispersion reinforces the structure, while contributing to the mitigation of fouling.
Interestingly, the porosity and pore size reported in these papers have shown to increase with increasing CNC concentrations. It is true that agglomeration of CNC is likely possible, which leads to poor dispersion in the polymeric matrix (especially for hydrophobic polymer such as PVDF), which could then result in decreased surface and bulk porosity and pore size of the prepared membranes, as the total solid content of the casting solution increases. Yet, the opposite has been observed in previous works for CNC. In addition, the substructures of the prepared membranes from these studies (both pristine and modified) display large macrovoids extending throughout the entire cross-section, which from former knowledge indicate weak points in membranes [155]. Although it can be argued that the highly hydrophilic nature of CNC could promote faster demixing rates subsequently forming large macrovoids, however, these claims should be substantially backed by kinetics data alongside thermodynamics and viscosity analyses.
The hydrophilic functional groups of CNC endow polymeric membranes such as PVDF, PES, CA, etc. with enhanced hydrophilicity that helped to mitigate the fouling; however, more work is needed to better understand the effects of CNC on the morphology of polymeric membranes, and how it could affect the long-term antifouling stability of CNC blending modified membranes.

2.1.4. Poly(vinyl alcohol)

PVA is an extremely hydrophilic polymer prepared by partial/full hydrolysis of poly(vinyl acetate). It exhibits outstanding thermal, chemical, and mechanical properties. However, its properties strongly depend on the degree of hydroxylation [105]. Nevertheless, it can be solubilized in a quite broad range of solvents, making it suitable for the blending modification of numerous types of polymer membranes [104,105,106], although compatibility issues arise with highly hydrophobic polymers [107]. Yuan et al. recently modified a PES membrane by blending different weight ratios (0–30 wt%) of PVA [106]. They reported morphological changes as well as an increase in membrane hydrophilicity. The authors showed that increasing the amounts of PVA in the casting solution increased the PWF (from 15 to 131 L/m2 h) while decreasing the BSA rejection (from 81% to 61%) due to the formation of macrovoids in the sublayer and the increased number of pores on the surface. Moreover, the FRR could be increased from 52% to 92.6%, indicating the excellent antifouling capabilities of the additive. Zhang et al. fabricated PVDF/PVA membrane [107], working with relatively low additive content (<0.5 wt%) because of the poor compatibility of PVDF and PVA. In spite of that, by fine-tuning the PVA content in the casting solution (0.1 wt% PVA), they managed to prepare stable membranes with improved antifouling properties.

2.1.5. Poly(acrylic acid)

Poly(acrylic acid) (PAA) enables a significant increase in water trapping in a membrane and is characterized by its biocompatibility. Furthermore, it is an effective crosslinking agent; its large number of carboxyl groups allows for easy adsorption or chelation of heavy metal ions, so it has been used for their removal [108,109]. Although the impact of PAA on the hydrophilicity and antifouling property of blending modified membranes can be demonstrated, the stability of the modifying agent in the matrix membrane can also be an issue. Due to its water solubility, PAA elution is inevitable when mixed directly with another polymer. Similarly to PVP, increasing the molecular weight permits one to address this issue, as it improves the tangling of PAA chains with the main polymer matrix chains. Ounifi et al. examined the effect of PAA on the removal of cadmium by cellulose acetate (CA) membranes [108]. The blending of 15 wt% of PAA into the CA matrix membrane led to a sharp decrease in the WCA from 71.5° to 25.0°, while the porosity increased from 44.6% to 75.6%. The addition of PAA facilitated water diffusion through the membrane by improving hydrogen bonding interactions. Thus, improved solvent/non-solvent exchange rates during membrane formation resulted in the formation of finger-like structures, and, in turn, a larger membrane porosity. The decrease in WCA was readily reasoned by the presence of numerous carboxylic groups. As a result, the modified CA/PAA membranes outperformed the convectional CA membrane in terms of water permeability. Interestingly, Cd rejection and humic acid (HA) rejection (99.9%) were also improved. In summary, PAA not only permits the mitigation of fouling of blending modified membranes, but also the removal heavy metals from water bodies, which is undeniably essential to safe wastewater reuse.

2.1.6. Polydopamine

Polydopamine or PDA is a product of dopamine polymerization using oxidizing agents such tris (tris hydroxymethyl aminomethane), ammonium persulfate, sodium periodate or sodium chlorate, copper sulfate buffer solution, and H2O2 [111,156]. The abundance of polar hydrophilic groups in PDA makes it a highly sought-after hydrophilic modifier for antifouling membranes. However, its high affinity towards water makes it highly vulnerable to leaching, allowing for the formation of large pores and increased porosity of the modified membrane as described in the work of Ang et al. [112].
Aside from endowing hydrophilic properties to the polymeric membranes, PDA can be easily functionalized with different functional groups to modify the surface charge of the membrane. Recently, Kallem et al. [114] functionalized PDA with –SO3H and subsequently incorporated into the PES matrix. Sulfonated PDA enhanced the surface negative charge of PES, increasing the electrostatic repulsion forces between the foulants (such as SA, HA and BSA) and the membranes’ surface. NF PES membrane exhibited increased rejection towards mono and divalent salts through the incorporation of quaternized PDA due to the increased presence of positively charged quaternary ammonium functional groups (-N+(CH3)3) [110]. Alongside being a stand-alone hydrophilic additive, PDA can also be used to functionalize nanomaterials such GO [157], nanoclays [158], metal nanoparticles [159], and so on.

2.1.7. Amphiphilic Copolymers

As mentioned in the previous section of the review, organic fouling and biofouling could be mediated by hydrophobic–hydrophobic interactions (among others prefaced in Section 1.3) with the membranes’ surface. Logic dictates that blending hydrophilic additives could confer conventional polymeric membranes with antifouling properties. However, these additives are highly soluble in water hence, they leach out in the coagulation bath and during membrane operation, ultimately causing the membranes to suffer from fouling. In order to keep the additives in the membrane matrix and enhance the stability of the blending modification, developments of amphiphilic copolymers containing both hydrophobic and hydrophilic segments were explored [95,125,126,127,128,129,130]. The hydrophobic segments act as anchoring fragments as they form hydrophobic–hydrophobic interactions with the polymeric matrix, thus preventing the additive from leaching out. Meanwhile, the hydrophilic segments endow the membrane with increased surface hydrophilicity as a result of surface aggregation during phase-inversion process. In this, the hydrophilic segments migrate towards the surface and promotes the formation of hydration layer through hydrogen bonding or ionic solvation which prevents the foulants from establishing hydrophobic–hydrophobic interactions with the membranes’ surface [25,160]. With higher concentrations of amphiphilic copolymers, surface segregation becomes more apparent, and the surface roughness of the prepared sample is increased. In alleviating foulant adsorption on the membranes’ surface, it was found that smoother surfaces are able to better repel foulants as they have better steric repulsion. On the other hand, rough surfaces could induce decreased steric repulsion as valleys and groves on the surface of the membranes make the polymer brushes easy to be compressed, which decreases the distance between the foulant and membranes’ surface. This phenomenon leads to a decrease in the Gibbs free energy of the system, and as it continues to decrease, fouling becomes more thermodynamically possible [25,161]. Hence, it is significant to find the balance between having strong hydration layer and good steric repulsion.
In utilizing amphiphilic copolymers, several parameters should be considered: (1) choosing between random and block copolymers; (2) regulating the molar ratios of the hydrophobic and hydrophilic segments; and (3) adjusting the chain lengths of the hydrophobic and hydrophilic fragments. The synthesis of random copolymers requires fewer complex procedures than block copolymers, hence, it can be easily scaled-up. Although random copolymers have wider molecular weight distribution, resulting in less stable assemblies compared to block copolymers, [162] random copolymers undeniably still offer good antifouling properties as exhibited from previous works on PS-r-PEGMA [90,163] PDMS-r-PEG random copolymers [128]. Despite having random assembly, 85–90% of PS-r-PEGMA remained in the PVDF matrix after 24 h immersion at 37 °C with constant stirring (simulating bacteria attachment tests) [90], while around 85–90% of PDMS-r-PEG remained in PVDF after 30-day immersion at 30 °C [128]. Of course, it is undeniable that block copolymers have better long-term antifouling stability and reusability as exhibited in previous works [95,125,126,127,160,164,165]. Zhang and colleagues reported that after long-term filtration, Pluronic F127/PMIA membrane water flux almost remain unchanged [160]. In choosing between random and block copolymers, we need to take into consideration the complexity of the copolymer synthesis, and long-term applications of the prepared membranes. Although random copolymers can be easily synthesized, their potential can be offset by their stability in the membrane matrix for long-term filtration and harsher conditions. On the other hand, block copolymers are ideal for long-term filtration applications. However, it requires complex synthesis which are costly to be scaled-up for industry.
Regulating the molar ratios of the hydrophobic and hydrophilic segments of the amphiphilic copolymers is also a significant aspect that needs to be considered to ensure effective antifouling properties and good stability in the membrane matrix. We recently investigated the effect of changing the hydrophobic/hydrophilic in PS-r-PEGMA [90]. Increasing the hydrophilic segments provided better hydration capacity, however, the stability of the copolymer in the matrix was significantly decreased due to the fewer “anchoring” segments that induces hydrophobic–hydrophobic interactions with the membrane matrix. Additionally, when only PEGMA was added (without polystyrene units), the contact angle of the prepared membrane remained high (~100°) which could be due to the inhomogeneous immobilization of hydrophilic segments in the membrane matrix. On the other hand, when polystyrene moieties are greater than PEGMA, the antifouling ability of the modified membrane was compromised, which led to the increased adsorption of protein on the surface. Aside from this, fine-tuning the chain lengths of the copolymer needs to be considered. Previous reports show that the strength of the hydration layer is highly correlated with chain length as more “water-hydrophilic units” interactions are formed through electrostatic interactions and hydrogen bonding. The water molecules attached to the surface could be categorized as free water, freezable water, and non-freezable water. Since free water is the farthest from the surface, it has the weakest bonding thus, it is very mobile and could be easily displaced [166]. Meanwhile, freezable water has moderate mobility, and could not be as easily displaced as free water. On the other hand, due to the proximity of the non-freezable water to the surface, it forms the strongest interaction to the surface, making it the hardest to remove. It has been investigated that with longer chains, more non-freezable water is formed, strengthening the hydration layer, and providing increased antifouling properties [90,167].

2.1.8. Zwitterionic Additives

Zwitterionic materials have both large dipole moments and charged groups. This results in the formation of a stronger hydration layer, compared to other types of non-charged antifouling moieties, essential to long-term fouling mitigation. These materials are endowed with the same number of cations and anions on their polymer chains, found on the same monomer units. Quaternized ammonium is a common cation, and common zwitterionic moieties are categorized as sulfobetaine (SB), carboxybetaine (CB), or phosphorylcholine (PC) based on the nature of their anions (sulfonates in SB, carboxylates anions in CB and phosphonates anions in PC). As a result of their propensity to associate with a large number of water molecules via hydrogen bonding and electrostatic interactions, zwitterionic materials are extremely hydrophilic, and their overall set of properties make them ideal antifouling materials [115,121,168,169,170]. However, and distinct from the other materials above-mentioned, zwitterionic units need to be combined with hydrophobic segments in order to prepare blending modified membranes. If not, the large polarity difference between the zwitterionic material and the membrane material makes the formation of a homogeneous casting solution extremely challenging if not impossible.
Li et al. developed an amphiphilic zwitterionic copolymer by grafting poly(sulfobetaine methacrylate) (PSBMA) moieties on PVDF chains, and then blended the resulting PVDF-g-PSBMA material with PVDF to form membranes by phase-inversion [118]. As for PVDF-g-PEGMA (Section 2.1.1), grafting the hydrophilic monomer on the same polymer used in the membrane preparation greatly enhances compatibility and stability of the system. When the PVDF-g-PSBMA content in the casting solution was increased to 14 wt%, the PWF was drastically enhanced from 121.9 L/m2 h (virgin membrane) to 239.1 L/m2 h. Moreover, compared to the virgin PVDF, the blended membranes showed a lower flux reduction ratio during filtration of a protein solution. The FRR for the virgin and of the best membrane were 51.5% and 81.2%, respectively, proving the efficiency of the zwitterionic material to mitigate fouling. Zhao et al. designed sulfonated polyaniline (SPANI) [126]. This material is self-doped with both positive and negative charges, resulting in the formation of a zwitterionic material. The blending modification of PVDF membranes with SPANI highly improved their hydrophilicity (WCA dropped from 92.0° to 29.0°) and consequently, their PWF (which increased from 97 to 160 L/m2 h). As a result of the excellent ion hydration, the absorption of contaminants was prevented. Our group modified PVDF membrane with a random zwitterionic copolymer containing styrene, sulfobetaine methacrylate, and ethylene glycol methacrylate units (PS-r-PEGMA-r-PSBMA) [117], and we also changed styrene to methyl methacrylate units [116]. The key aspect of this set of studies was the use of PEGMA as a solubility-enhancing unit, facilitating the blending of the zwitterionic material with PVDF. The zwitterionic MF membranes prepared by the VIPS process were shown to resist fouling caused by a large variety of biofoulants including blood components. The design of blood compatible membranes by blending modification was also investigated, using a copolymer of 2-methacryloyloxyethyl phosphorylcholine and methacryloyloxyethyl butylurethane (MPC-derivative), hence containing PC units popular for their hemocompatibility [122]. Finally, Maggay et al. recently investigated the effect of a zwitterionic copolymer derived from 4-vinylpyrridine (denoted as zP(S-r-4VP)) on the antifouling properties of PVDF membranes prepared by VIPS [119] and of PSf membranes formed by a dual-bath process [120]. Interestingly, this material permits the preservation of the antifouling properties of the membranes even after exposure to hot steam (steam sterilization process). This was also a key feature of a recent study staring a copolymer containing sulfobetaine methacrylamide functions (SBAA) [123]. While SBMA can be hydrolyzed, it appears that SBAA is more stable when exposed to hot steam, making its derivatives potential candidates for the formation of antifouling membranes for biomedical applications.

2.1.9. Nature Derived Biopolymers

Growing environmental concerns have prompted many researchers to delve into nature-derived hydrophilic polymers (listed in Table 1) due to their easy access, reasonable cost, non-toxicity, and intrinsic hydrophilicity [144,171]. These biopolymers have been well-documented to improve the hydrophilicity of polymeric membranes, which could help enhance the flux whilst mitigating foulants by forming hydration layer. Although these biopolymers offer promising antifouling properties with less environmental risks, their widespread applications are still being challenged by their solubility in organic solvents, stability in the membrane matrix, and scalability of their synthesis procedures (extraction of the desired compound from natural raw materials). For example, chitosan or its derivatives are insoluble in common solvents [137,139], which makes the modification of the bulk of the membrane difficult. Moreover, since these biopolymers have many organic linkers [131,142,144] that imparts them with hydrophilic properties and polarity, these biopolymer additives could be easily washed off during immersion in water. Although this is beneficial in promoting a porous structure during phase-inversion, prolonged immersion in water such as in water treatment operation could cause these biopolymers to leach out over time and could eventually lead to decreased antifouling properties. Additionally, since they are derived from nature, the steps needed to achieve the “purified state” of the biopolymer are no easy feat. Moreover, the yield from these procedures should also be taken into consideration. Despite being considered as more sustainable alternative additives, there are still a lot of factors that limit the potential of biopolymers from being fully realized in membrane technology.

2.1.10. Concluding Remarks on the Use of Polymeric Additives

Hydrophilic (both synthetic or natural) or amphiphilic polymers and copolymers have remained popular in the past 10 years to improve the antifouling properties of membranes formed in one step by phase-inversion. Their incorporation to the membrane system does not just result in better mitigation of fouling, but also in drastic changes in membrane structure and arising properties (pore size, porosity) through the effect of the additive on membrane formation mechanisms. The maximum concentration of the antifouling additive has to be carefully adjusted in order to reach the desired fouling mitigation while maintaining the mechanical properties. This aspect has to be highlighted as the LIPS process applied to casting solutions containing hydrophilic/amphiphilic additives often leads to membranes with larger macrovoids than those seen in the pristine matrix. These macrovoids surely facilitate transport during filtration (provided also that they are connected through porous walls), but severely weaken the membranes. Moreover, it is possible to combine zwitterionic materials with common membrane polymers which is an important recent highlight. They do not dominate but progress has been recently made, which opens avenues to more applications in the biomedical field, given the well-established hemocompatibility of some of these materials.

2.2. Metal-Based Additives

Apart from organic-based compounds or nanomaterials, inorganic compounds such as metal-based nanomaterials have been investigated to enhance the fouling resistance of membranes against a variety of organic and biofoulants. Common metallic nanomaterials that are utilized for blending modification are shown in Table 2, including metallic or zerovalent metals, biphasic metals such as metal oxides and metal dichalcogenides, and bimetallic oxides.

2.2.1. Metallic Additives

Recently, metallic nanoparticles have been studied as additives in membranes due to their antimicrobial properties. Aside from being non-toxic and hypoallergenic, silver (Ag) nanoparticles have been widely used as an antimicrobial agent in medicine, and also in wastewater treatment. Ag can interact with the sulfur and phosphorous groups, most commonly in thiol groups (S-H) in proteins, which contributes to the disruption of the bacterial proteins and hinders both respiration and electron transfer [232]. Studies have shown that the use of Ag nanoparticles into the membrane matrix significantly enhanced the permeation flux, antimicrobial, and antifouling properties of membranes as a result of increased hydrophilicity [172,173,174,175,176]. However, this is prone to solubility issues in the matrix and is also susceptible to particle aggregation which altogether affects their stability in the matrix. In addition, the long-term use of membranes impregnated with Ag nanoparticles has been greatly challenged by the leaching of Ag, which results in a decline in the antimicrobial and antifouling properties of the membranes. Ultimately, if the leached Ag gets into our drinking water, this could lead to metal poisoning which could lead to fatal consequences [232]. Esfahani recently reported an interesting study on citrate-stabilized gold (Au) nanoparticles for PSf membranes. They reported that the presence of the nanoparticle degraded the cake formed by humic acid which significantly enhanced the antifouling properties of the membrane (FRR = 95%). The disaggregation of the HA cake on the surface could be attributed to the ligand exchange interaction between the HA molecules and citrate-stabilized Au nanoparticles. Instead of forming hydrophobic–hydrophobic bonds with the membrane, HA formed hydrogen bonds with carboxyl groups of the nanoparticles. These newly formed bonds altered the existing hydrogen bonds between HA molecules, and therefore changed the hydrophobic forces that hold the HA molecules together [177,233,234,235].

2.2.2. Biphasic Metals

Zirconium oxide (ZrO2) is of considerable interest for its outstanding thermal stability, chemical stability, corrosion resistance, and biocompatibility. As such, it can contribute to the formation of more stable and biocompatible membranes applied in harsh environments. The Zr-O bond can be readily hydrolyzed, resulting in the formation of hydroxyl groups, and ultimately increasing the membrane’s hydrophilicity [179,180,181]. Pang et al. fabricated an antifouling MMMs UF membrane by blending ZrO2 with PES [179]. The addition of ZrO2 increased the pore size and porosity, resulting in the enhancement of the PWF. Moreover, the FRR of a modified membrane containing 1.0 wt% ZrO2 was 1.8 times that of the bare PES membrane. However, larger surface pore size for the modified membranes resulted in slightly lower rejection performances. Similarly, Huang et al. prepared a UF membrane by blending ZrO2 with natural bamboo cellulose (BC) [181]. Compared to the unmodified cellulose membrane, the ZrO2/BC membrane was reported to be more stable and showed improved antifouling capabilities. To avoid the aggregation of nanoparticles and also to further increase the surface hydrophilicity, Shen et al. grafted poly(N-acryloylmorpholine) (PACMO) on the nanoparticles [180]. The resulting ZrO2-g-PACMO endowed PVDF membranes with excellent FRR (97%) after the separation of oil/water mixtures. The hydrophilic effect of grafted PACMO chains and the well-distributed ZrO2-g-PACMO in the membrane pore channel allowed for the efficient removal of oil droplets from the membrane pores.
Zinc oxide (ZnO) is one of the low-cost and multifunctional inorganic nanoparticles that is gaining popularity due to its unique features including its catalytic activity and powerful bactericidal effect. Furthermore, because of their high hydrophilicity, ZnO nanoparticles are one of the best materials for improving the hydrophilicity of composite membranes. ZnO nanoparticles can effectively adsorb water molecules and can be readily incorporated into a matrix polymer [82,236,237]. Rajabi et al. blended ZnO of different structures (ZnO nanoparticle, ZnO-NP, and ZnO nanorod, ZnO-NR) with PES to investigate the effects of the nanofiller on the membrane’s properties [82]. The improved hydrophilicity of the modified membranes was confirmed by WCA measurements. ZnO-NP and ZnO-NR showed a WCA of 60° and 54°, respectively, while the virgin PES membrane exhibited a WCA of 77.9°, resulting in improved PWF. The antifouling property, determined by filtering milk powder solution, revealed that ZnO-NR slightly outperformed ZnO-NP. It could be attributed to a better migration of ZnO-NR towards the top surface during membrane formation, hence resulting in more adsorbed water molecules. Rabiee et al. studied the permeability and antifouling properties of PVC UF matrix membranes modified with ZnO nanoparticles [236]. Again, the effect of the nanoparticles on fouling resistance was clearly highlighted, with a sharp increase in FRR (from 69.3% to 91.8%) after fouling test (BSA filtration). As for other types of nanoparticles, aggregation may occur as reported by these authors from a loading of 4 wt%, leading to pore blockage. However, a small quantity of ZnO already ensures significant improvement in the overall antifouling properties of membranes, as seen from the results reported by Purushothaman et al., who modified poly(ether ether sulfone) (PEES) membranes and observed a significant improvement of the fouling resistance to HA using as little as 1 wt% nanoparticles [237]. Hence, the FRR jumped from about 45% for the pristine membrane to over 92% for the modified membrane.
The use of titanium oxide (TiO2) has gained momentum due to its ease of synthesis, stability under extreme conditions, and commercial availability. It is considered to be an outstanding material for the development of composite membranes applied in water treatment as it also shows excellent oleophobicity, which reduces membrane fouling by oily compounds. Furthermore, TiO2 nanoparticles are chemically stable, have a high wettability by water, are antimicrobial and can act as catalysts for the degradation of micropollutants [15,38], although we will only focus here on their antifouling properties under normal conditions (i.e., no irradiation promoting catalytic degradation of foulants). Zhang et al. used a nano-TiO2/PEG composite to modify PVDF membranes [38]. PEG was primarily utilized in the casting solution to increase the stability and dispersion of the TiO2 nanoparticles. Compared to bare PVDF membrane, the pore size and porosity of TiO2-modified membranes were decreased. Furthermore, the PWF dropped gradually as the TiO2 concentration increased above 0.15 wt%, which was attributed to the excessive aggregation of nanoparticles, which leads to pore blockage during phase separation as well as to an increase in surface WCA and overall poor membrane performances. However, provided that the homogeneous dispersion of the nanoparticles is achieved for a concentration below 0.15 wt%, the surface hydrophilicity of the membranes could be improved (WCA decreased). Arif et al. also fabricated PVDF/TiO2 composite membranes [190]. By controlling the TiO2/PVDF weight ratio to 0.01/1, they obtained a membrane exhibiting a high FRR (approximately 88.2%) after fouling tests, demonstrating the beneficial impact of the nanoparticles of fouling resistance. Again, however, higher contents (0.02 g TiO2/g of PVDF) led to the agglomeration of TiO2 particles, causing the surface roughness to increase due to the creation of bumps, resulting in poorer antifouling performances. Moghadam et al. also worked with a similar system (PVDF/TiO2) and prepared a UF membrane for the removal and degradation of pollutants [238]. While the catalytic effect of TiO2 under UV irradiation is beyond our scope, the authors showed a significant increase in PWF and protein flux of the membranes without UV irradiation, hence supporting the improvement of membrane hydrophilicity and reduced membrane fouling resulting from the inclusion of nanoparticles in the matrix.
Silicon oxide (SiO2) can be utilized as inorganic modifier of membranes for its stability, strong hydrophilicity, mild reactivity, and high chemical resistance. However, direct mixing with polymers is challenging as it is difficult to obtain well-dispersed nanoparticles in the membrane, limiting their range of application. Thus, their surface modification prior to dispersion is recommended. Tripathi et al. prepared membranes by mixing silica nanoparticles modified with dopamine (DOPA) and polyacrylonitrile (PAN) [239]. The change in performance, hydrophilicity, and antifouling features of composite membranes containing increased concentrations of SiO2-DOPA were compared to those of pure PAN membranes. The modified membranes demonstrated increased PWF and rejection capabilities in comparison to the unmodified PAN membrane. SiO2 nanoparticles that had been modified with DOPA had greater dispersion in the organic solvent (DMF) during membrane fabrication. The particles were evenly distributed in the solution, and there was no evidence of agglomeration even after extended storage of the solution. As a result, the compatibility of particles with the PAN matrix was improved. The authors also stressed the improvement of FRR, and reported both high protein rejection and dye rejection, suggesting their applicability for wastewater treatment.
Copper oxide (CuO) nanoparticles have also been explored recently. It is non-toxic and highly abundant in nature [217]. Hosseini and colleagues [188] reported a decrease in the WCA of PES membranes ascribed to the inherent hydrophilicity of CuO, and subsequent migration on the surface of PES upon phase-inversion. However, a significant increase in WCA occurred as the concentration CuO was increased. This could be attributed to the agglomeration of nanoparticles in the dope solution, and these agglomerated nanoparticles migrated to the surface which resulted in increased surface roughness. The influence of leaching of CuO on the membrane performance was investigated by Kaju et al. [218] in which they found that after exposing the CuO-modified PES membrane to different cleaning solutions (0.1 M NaCl, 0.1 M HCl, 0.1 M NaOH, 0.5% NaClO, and pure water) resulted in increased porosity, enhanced permeability, and decreased salt rejection. Although comparable results were also obtained on pristine PES, no significant increase in porosity on the virgin membrane was observed. Hence, the evident enhancement in the porosity of the modified membranes after long hours of cleaning could be attributed to the leaching of CuO.
Iron oxides in the forms of Fe2O3 (Ferric oxide) and Fe3O4 (Ferrous oxide) are promising metal oxides for improving the membrane performance against fouling with improved permeability [14,51,205,219,220,221,222,223,224,225,226,227]. In addition, due to the magnetic nature of iron oxide, it is useful to the study of the photocatalytic performance of polymeric membranes [240]. However, iron oxides have poor solubility in organic solvents [51] and are also prone to agglomeration, as shown in the works of Demirel et al. [223,224]. To resolve this, modifications involving the immobilization of reactive ligands on the surface of nanoparticles, surface coating with an adsorptive layer, or a combination of both, have been recently studied to improve the affinity of iron oxides in the polymer matrix [51], and consequently enhance the membrane performance. For instance, Zinadini et al. [227] blended O-carboxymethyl chitosan coated Fe3O4 into the PES matrix, resulting in an increased hydrophilicity, dye rejection, and fouling resistance. Kamari and Shahbazi [205] recently reported the incorporation of NH2 functionalized SiO2 coated Fe3O4 in PES membranes that led to better membrane performance in terms of permeability, salt rejection, heavy metal ion removal, methyl red dye retention, and resistance towards BSA fouling.
Aluminum oxide (Al2O3) has also been considered as an antifouling additive for membranes due to its non-toxicity, and affordability [228]. Mehrnia et al. studied the concentration threshold of Al2O3 in PSf membranes by examining the rheological properties of the dope solution with different Al2O3 concentration. According to their study, when the concentration threshold is exceeded, a drastic increase in the viscosity of the dope solution is observed which could affect the thermodynamic stability, and kinetics of the phase-inversion process. There is a noticeable drop in the porosity of PSf beyond 0.39 wt% of Al2O3; however, high permeability is still maintained due to increased surface wettability owing to the hydrophilicity of the nanoparticle [55]. However, it must be noted that despite the intrinsic hydrophilicity of the nanoparticle, further increase in the concentration would not equate to increased surface wettability of the membrane at some point. In fact, the opposite could be observed, as reported in earlier study [229], due to the increased agglomeration of the nanoparticles resulting in increased surface roughness.

2.2.3. Bimetallic Oxides

Bimetallic oxide nanoparticles combine within one material the advantages of several monometallic nanoparticles but exhibit smaller sizes and both higher surface areas and stability. Considering their properties, bimetallic should lead to even better hydration and fouling resistance than monometallic nanoparticles. From these considerations, Arumugham et al. modified PES membranes prepared by LIPS using nanoparticles of iron(III) oxide and manganese (III) oxide [231]. After adjusting the concentration in nanoparticle, they managed to minimize the surface roughness, surface hydrophilicity, and maximize the porosity. Consequently, the optimized membranes exhibited a permeability which was almost double that of the pristine matrix, and fouling was well-mitigated after five cycles of dynamic tests involving BSA (FRR of 65.8%).

2.3. Ceramic-Based Additives

Ceramic materials are more thermally stable and have greater mechanical strength than polymeric materials. In recent years, the blending of hydrophilic ceramic materials (listed in Table 3) into polymeric matrix membranes has emerged as an efficient method to mitigate fouling.
These nanofillers can change the surface characteristics of membranes and increase their antifouling performance [232], provided that homogeneous dispersion is achieved, as aggregation can lead to structural defects then causing a decline of all performances. Fouling resistance, excellent permeability, good flux recovery, chemical stability, and an extended lifespan are just a few of the benefits of blending ceramic elements into polymeric membranes. Thus, ceramic materials show promising results in water treatment [36,282]. Considerable attention has been given to ceramic nanomaterials such as nanoclays due to their intrinsic hydrophilicity, high surface area, good mechanical strength, low toxicity, and low cost [60,62,255,283,284]. In addition to these, nanoclays could also improve the thermal stability and mechanical strength of the membranes [284]. The nanoclays listed in Table 3 have been shown to endow polymeric membranes with improved surface wettability which evidently enhances the antifouling properties. However, they also suffer from immiscibility in the polymer matrix and agglomeration which consequently decreases the membrane performance and also negatively impacts the mechanical properties of the membrane. In general, inorganic nanomaterials have high surface energy. This, along with high interparticle interactions, is the primary contributor for the agglomeration of nanoparticles. With high concentration or loading, the distance between filler nanoparticles becomes very small and they tend to agglomerate because of the van der Waals forces which prevents the nanoparticles from dispersing uniformly [285]. In order to improve this, these nanomaterials are functionalized with hydrophilic matierials such as taurine (–SO3H) and PDA [253], activated carbon [252], folic acid [247], (3-aminopropyl) triethoxysilane (APTES), silane coupling agent (–NH2), ethoxy (–OC2H5) [250], curcumin (–OH, phenols, and carbonyl groups) [260], N-halamine (–NH2, –C=O) [264], and dextran (–OH) [276] to improve their miscibility and dispersibility in the blend (dope casting solution).

2.4. Carbon Allotropes and Porous Nanomaterials

Over the past decades (not limited to membrane technology), carbon allotropes and porous nanomaterials have been tremendously employed in various studies. Lately, these materials have been introduced as upcoming additives for membranes due to their interesting properties such as complex functional groups, high surface area, and so on. Listed below (Table 4) are the different carbon allotropes and porous nanomaterials that have enhanced the antifouling properties of common polymeric membranes through blending approach.

2.4.1. Carbon Allotropes

Carbon allotropes including graphene oxide (GO) and reduced graphene oxide (rGO), as well as other carbon materials such as multiwalled carbon nanotubes (MWCNTs) or carboxylated nanodiamonds (CNDs), have been used in the past decade to develop membranes by blending modification for water treatment [288,302,304,305,306,307,308,344,345,346,347].
CNDs, prepared via the thermal treatment of nanodiamond particles in order to incorporate oxygen-containing groups and remove sp2 non-diamond carbons, have been reported in the study of Li et al. [288]. The authors showed that adding 1 wt% of CNDs drastically improved the antifouling properties of PVDF membranes. Moreover, they emphasized the role of thermal treatment and showed that membranes modified with raw nanodiamonds still suffered from important fouling. They also pointed to the need for adjusting the concentration in CNDs in the casting solution.
Graphene (GN) is a two-dimensional (2D) material composed of carbon atoms with sp2 hybridization orbitals arranged in a hexagonal lattice in a planar configuration. The wetting of GN is thickness dependent, and monolayer GN is hydrophilic, making it a potential candidate for the mitigating membrane biofouling in aqueous media. It has been utilized to construct nanocomposites for several membrane applications [307,344]. Recently, graphene quantum dots (GQDs) modified with citric monohydrate were used in combination with PVC and PEG [308]. The functionalization of GQDs with hydrophilic functional groups improved the antifouling properties provided that careful adjustment of the nanoparticles content. Indeed, the roughness value increased at high concentration (2.0 wt%), causing fouling to increase. The inclusion of oxygen-containing functional groups to graphene such as hydroxyl, carboxyl, carbonyl, and epoxy groups results in the formation of GO, more hydrophilic than GN. Several teams reported that the blending of carbon-based materials into the polymer matrix such as GO [304,305,306] or rGO [306] increases the membrane’s hydrophilicity, resulting in improved permeability during water treatment. Xia et al. incorporated GO into a PVDF membrane prepared via NIPS [305]. They pointed out that GO contributed to the removal of the dissolved organic carbon (DOC) through its oxygen-containing functional groups, allowing the membranes to maintain high flux throughout the permeation process. A low content (0.1–0.3 wt% as in the study of Mohsenpour et al. [304] or 0.5 wt% GO as in the work of Xia et al. [305]) is enough to achieve the desired properties. High amounts of GO have a detrimental impact on the permeability of the membrane, possibly due to a decrease in porosity and pore size resulting from an increased viscosity of the casting solution used to produce the membrane. To boost the performances of membranes, the aggregation of GO nanosheets at high GO concentrations must be avoided. Moreover, not only the GO loading should be fine-tuned, but also the reduction degree, as examined by Meng et al. who reported that a higher dissolution temperature could lead to a lower degree of reduction which could then prevent aggregation [306]. Note also that the presence of hydroxyl groups of GO impacts the kinetic of mass transfers during membrane formation, hence affecting the membrane structure. More or larger macrovoids are likely to be observed if membranes are formed by wet-immersion, as pointed out by Bala et al. [345] or Mohsenpour et al. [304].
MWCNTs are materials composed of multiple graphene sheets organized in a concentric fashion. In water treatment application, they can be used for their ability to adsorb large quantities of micropollutants, property imparted by their large surface area. Nevertheless, they are generally hydrophobic. So, their functionalization with hydrophilic pendant groups before their inclusion in membrane systems is recommended in view of maintaining high membrane permeation. In addition, the grafting of some particular functional groups onto MWCNTs can enhance the interactions, leading to the trapping of pollutants. Mahdavi et al. [303] and then Haghighat et al. [302] synthesized polypyrrole-modified MWCNTs (PPy-MWCNTs), by an oxidation polymerization process, before their inclusion in membrane systems. The existence of amine groups on the surface of the membrane permitted the improvement of the hydrophilicity of the modified membranes, which then led to fouling mitigation. However, as for other carbon-based materials, particle agglomeration may occur at large contents, and the addition of 0.25 wt% of PPy-MWCNTs in the study of Haghighat et al. [302] led to the best outcome in terms of water permeation and dye removal.
So, oxygenated derivatives of graphene and carbon materials have been used in the past ten years in combination with polymeric materials and have proven to be an efficient way to control surface wettability, enhance rejection performances and mitigate fouling of numerous polymeric membranes prepared by phase-inversion. Nevertheless, the preparation of homogeneous casting solution remains a challenge that need to be addressed, and it is difficult to use optical techniques to assess the particles size or the extent of particles aggregation. Yet, this phenomenon must be prevented as it also weakens the membranes. Furthermore, these materials significantly increase the casting solution viscosity, which then has consequences on the membrane structure and arising performances.

2.4.2. Porous Nanomaterials

Porous nanomaterials such metal-organic framework (MOF), Zeolites/ZIF, and cyclodextrin are some of the emerging nanomaterial additives to help resist the adsorption of non-specific foulants. These nanomaterials have highly porous structure and high surface area which could help improve the porosity and flux performance of the membranes. The organic linkers of these nanomaterials could help induce their uniform dispersion in the casting solution. Aside from providing good antifouling abilities to the host matrix as reported in the studies in Table 4, MOF and zeolites/ZIF have good photocatalytic properties towards emerging pollutants such pharmaceutical foulants [348].

2.4.3. Concluding Remarks on the Use of Inorganic Nanoparticles (Metal-Based, Ceramic-Based, Carbon Allotropes and Porous Nanomaterials)

In general, the nanofiller’s hydrophilicity helps to improve the wettability of the modified membranes, ultimately resulting in improved antifouling property of the composite membranes. Compared to polymer additives, smaller concentrations are enough to significantly improve the antifouling properties of membranes. Furthermore, higher rejections as the additive content increases are commonly reported due to the increase in the casting solution viscosity leading to denser membranes. Moreover, some nanoparticles can reduce fouling because of their intrinsic hydrophilicity, but also, because they act as catalyst (although beyond the scope of this review). However, high concentrations cause the aggregation of nanoparticles, which blocks surface pores, reduces the membrane performances, and even sometimes their antifouling characteristics.
Similar to polymeric additives, leaching and aggregation are some of the major drawbacks in using inorganic additives such as metals. However, it appears to be exacerbated in inorganic additives due to its incompatibility with the polymer matrix. The use of physical blending of these nanomaterials in the polymer casting solution and phase-inversion gave rise to the poor distribution of the nanomaterials in the matrix as a result of the formation of aggregates in the dope casting solution. To overcome these problems, researchers have tried to increase the surface area of the nanomaterials by controlling their particle size [33]. Another method is functionalizing the nanomaterials in order to increase the interaction with the host material [204]. These observations reported by several groups, regardless of the nature of the nanoparticles, suggests the need for optimizing the composition of the casting solution before phase separation. Materials such as PEG, DOPA, etc. can be added to the blend, which main function is to limit aggregation and improve dispersion. In order to achieve the latter, the precursors of the nanomaterials must be soluble in the dope casting solution and can react with the non-solvent in the coagulation bath [34,349].

2.5. Hybrid Nanomaterials

Recently, the development of new antifouling materials has involved complex synthesis of additives. Some studies combine two sets of additives and blend them together into the polymeric matrix [81,289,350,351,352,353,354] to improve both antifouling properties and photocatalytic activities of the prepared membranes. However, this should be carried out with great consideration with regards to the ensuing thermodynamics instability of the system, and how this could also influence the viscosity and kinetics of the phase separation. Moreover, we should be reminded that this could lead to high degree of aggregation in polymeric matrix, instead of improving the antifouling properties, this could eventually lead to the poor performance of the membranes in terms of resisting foulants, and separation performance. Additionally, poorly dispersed additives could aggravate leaching.

3. Blending of Additives into the Polymer Matrix and Its Effect on Membrane Morphology

As mentioned earlier, blending modification refers to the blending of the main polymer with an antifouling material in a common solvent, prior to applying a phase-separation technique. There are three major phase-inversion processes leading to the formation of porous membranes (LIPS (wet-immersion), VIPS, and TIPS) with which blending modified membranes can be prepared. Although the purpose of the modification is to impart the membranes with antifouling properties, it also impacts membrane structure. Below is a discussion on major effects observed based upon recent literature.

3.1. Influence of the Antifouling Additive on the Morphology of Liquid-Induced Phase Separation Membranes

The fundamentals of membrane formation by phase-inversion have been thoroughly described [355,356] and are beyond the scope of this text. Instead, we would like to discuss the effects of the additive on thermodynamic and kinetic based on experimental observations published in the last 10 years.
Thermodynamics is described by the Flory-Huggins theory [355]. In this theory, the binodal line separates a single-phase region from a two-phase region. It can be determined experimentally by cloud points measurements, which consists of adding dropwise non-solvent to a casting solution under constant stirring and at constant temperature and measuring the total volume of non-solvent needed to induce permanent cloudy state. For this composition, phase separation has been reached. It can then be positioned in a ternary phase diagram represented by a triangle, with each vertex representing a pure component of the system: polymer, solvent of non-solvent. Adding an antifouling material to the casting solution complexifies the analysis since the composition of the quaternary system (polymer/antifouling material/solvent/non-solvent) should be plotted in a 3D diagram (triangular prism). However, it is not carried out as the third dimension is already taken by the temperature (a ternary phase diagram is only valid at one given temperature). Instead, a pseudo-ternary phase diagram is determined where the polymer and antifouling material may occupy one same vertex of the triangle. So, in this representation, polymer and antifouling material are seen as one unique material. In doing so, the polymer/antifouling material weight ratio in all of the solutions used to determine the set of cloud points should remain constant. As then observed using this experimental approach and tentatively schemed in Figure 4, the addition of an antifouling material to a casting polymer/solvent solution results in the shift of the binodal line towards the polymer-solvent axis of a ternary phase diagram [86,116,117]. In other words, the demixing gap is reduced such that fewer amount of non-solvent is needed to induce phase-separation. As the system is less thermodynamically stable by the addition of an antifouling material, film formation is facilitated. Logically, if the distance to cross the demixing gap is changed, the time needed to reach the demixing region will be affected as well. Shorter times can be expected although changes in the system viscosity also need to be accounted for.
As for kinetic aspects, instantaneous phase-inversion and delayed demixing are often associated with high porosity membranes decorated with macrovoids and with denser, macrovoid-free membranes, respectively. Phase-inversion in LIPS (wet-immersion) often happens rapidly as DI water is a common non-solvent with high affinity with most solvents used in the industrial scale production of membranes, leading to fast exchanges. The addition of hydrophilic of amphiphilic additives increase the affinity of the dope with water, likely promoting exchange rates [357]. Nevertheless, it also affects the viscosity of the system. If an increase in viscosity is measured (seen if few percent of solvent is replaced by the additive, compared to the formulation of the pristine dope), then it may slow-down the exchange rates. Two phenomena need to be accounted for when rationalizing the kinetic of phase-inversion of blending modified membranes: the affinity of the additive with water which is normally high because hydrophilicity is a criterion for antifouling, and the change in viscosity associated with the addition of water. The latter one has often been mentioned in recent literature to explain the disappearance of typical macrovoids observed in LIPS membranes. For example, Khalil et al. developed PLA-based membranes using PVP as an additive [358]. They reported a gradual disappearance of macrovoids with PVP because the additive increased the dope viscosity. Similarly, Nainar et al. [359] mixed PVDF with CA and added PVP to the dope solution. The pristine PVDF membrane possessed macrovoids, which tend to vanish with the addition of hydrophilic CA, a consequence of the higher solution viscosity. PVP then led to larger macrovoids (which could be a consequence of the large affinity with water) although a rather dense sublayer sitting on top of the large macrovoids decorating the cross-sections was observed due to the high viscosity of the casting solution. As another example, macrovoids became much larger with the addition of PS-b-PEGMA to a PVDF/N-methylpyrrolidone dope in one of our previous studies [357] due to both the decrease in dope viscosity and the affinity of PEGMA units with water (Data comparing a virgin PVDF membrane and a membrane containing 1 wt% copolymer obtained with a similar system are displayed in Figure 5). Yuan et al. also reported accelerated phase-inversion with the addition of polyvinyl alcohol (PVA) to a polyethersulfone (PES) membrane, resulting in larger pore size and porosity [106]. The effect of viscosity on membrane structure following the addition of nanoparticles is also reported. Nano-SiO2 particles have been shown to increase the thickness of the top layer of polyvinyl chloride (PVC) membranes, attributed to a more viscous nano-SiO2/PVC/solvent casting solution, compared to a PVC/solvent casting solution [209]. Xia and Ni reported the interplay between viscosity and affinity with water, in a PVDF/graphene oxide (GO) system [305]. The viscosity of the dope increased with the addition of GO, but only 1 wt% led to a sharp increase. At 0.5 wt% and below, the change in viscosity was small. They observed the formation of larger macrovoids in the membrane containing 0.5 wt% GO (compared to the pristine membrane), which they attributed to the affinity of the additive with water, which led to thermodynamic instability, and promoted faster mass transfers. On the other hand, they also pointed out that the porosity of a membrane formed after adding 1 wt% GO decreased (compared to that of a membrane containing 0.5 wt% GO), which they attributed to a larger viscosity.

3.2. Influence of the Antifouling Additive on the Morphology of Vapor-Induced Phase Separation Membranes

The VIPS method has been used much less than the LIPS process to fabricate antifouling membranes. In the past 10 years, few teams have worked on the design of such membranes [90,116,117,119,122,360,361,362]. VIPS leads to MF membranes, which makes it a complementary method to LIPS which generates UF membranes. However, the equipment needed is more complex to implement since a control atmosphere of non-solvent needs to be created. Furthermore, it competes with TIPS in terms of pore size range. Yet, the slowness of mass-transfers makes it an interesting process to control membrane structuring, and to optimize surface segregation of the antifouling material. Moreover, there is no initial fast solvent outflow that could readily drag the hydrophilic/amphiphilic additives out of the film before phase-inversion is complete as in LIPS. Thus, the majority of additive molecules is expected to remain trapped in the film.
As in LIPS, the additive has a considerable impact on membrane formation. From a thermodynamic perspective, the impact of the additive on phase-inversion is the same in VIPS and in LIPS, in that it shifts the binodal line closer to the polymer/solvent axis. Likewise, one can expect a similar trend regarding the effect of the additive on formation kinetic, although it is tricky to measure in a VIPS chamber. Again, the measured effects of the additive on the kinetic of phase-inversion cannot just be attributed to the hydrophilic groups in the additive, as the change in viscosity must be considered as well. Carretier et al. reported faster transfer in a system containing 20 wt% PVDF and 5 wt% PEGMA-b-PS-b-PEGMA than in a system containing 25 wt% PVDF [86] which was not just due to hydrophilic PEGMA, but also to the lower viscosity of the solution containing the antifouling additive. However, Dizon and Venault also reported faster phase-inversion in a solution containing an amphiphilic additive than in a pristine solution. Yet, the former solution was slightly more viscous [117]. In many of our previous reports on the design of antifouling PVDF membranes by VIPS, we noted that the additive helped in preventing nodule formation or at least at decreasing their size [116,117,357,363] (Figure 6). Nodules arise from a crystallization-gelling process [364], and their presence in fouling-resistant PEGylated PVDF membranes formed by VIPS has also been reported by another group [362]. They tend to weaken the membrane structure, but their size can be controlled by the temperature of dissolution of the casting solution. Nevertheless, for a same level of temperature, the amphiphilic additive prevents or at least reduces the growth of nodules. There are two underlying reasons. Firstly, the acceleration of non-solvent transport resulting from the addition of hydrophilic units favors L/L phase separation over crystallization. Secondly, in the case where the viscosity increases after the addition of antifouling copolymer, viscous forces oppose the growth of crystal nuclei. As a consequence, bi-continuous structures are readily formed. These structures are attractive in MF membranes for their combination of high permeability and mechanical stability.

3.3. Influence of the Antifouling Additive on the Morphology of Thermal Induced Phase Separation Membranes

The TIPS method has undeniable advantages over the two NIPS processes in that it leads to stronger membranes with a uniform pore structure. Nevertheless, the need for high temperature of the dope is detrimental to its usage for the widespread formation of antifouling membranes by blending modification. In particular, antifouling polymers lack thermal stability. There are plenty of studies starring the TIPS method followed by a post-treatment imparting the membrane with antifouling properties, but very few exist, to our knowledge, in which only TIPS is used to induce phase separation of a polymer/antifouling material/diluent blend. In the period of time on which this survey focusses, we shall mention the work performed in Matsuyama’s group who entrapped a PEGylated copolymer in PVDF TIPS membranes in one step (hollow fibers) [365] as well as those of Xu et al. [23] and Li et al. [366], who used multi-wall carbon nanotube and nanoparticles of SiO2 and GO, respectively, to fabricate PVDF membranes by TIPS (flat sheet). Regarding the former, the authors did not notice any significant effect of the copolymer on the membrane structure. All membranes, pristine and PEGylated, exhibited a spherulitic porous structure, with similar roughness, characteristic from the TIPS process applied to a PVDF/diluent blend. The authors justified the absence of significant morphological change with the molecular weight of the copolymer. According to them, it corresponded to an effective diameter too small, in comparison with the surface pore diameter, to affect the membrane structure. However, in the case of MMMs, the authors noticed the role of the nanoparticles on the crystallization of the polymer [23,366]. While the MMMs exhibited a morphology characteristic from L/L phase-inversion (cellular pores/sponge-like morphology), the authors explained that the nanoparticles acted as nuclei promoting the crystallization of PVDF and so the formation of small spherulites [23]. Larger amounts of nanoparticles then led to more crystalline grains but also smaller pores. Eventually, increasing further the nanoparticles content resulted in denser membranes as the dope viscosity increased.

4. Assessments of Antifouling Properties of Modified Membranes through Blending Approach

The development of antifouling membranes is complicated by the fact that various fouling scales, such as the nano-scale for proteins and the micro-scale for bacteria, must be considered. If the formation of a strong hydration layer is accepted as a pre-requisite to prevent fouling of membranes used in aqueous media, the characteristic size of foulants needs to be considered. As wastewater or biological media often contain a complex mixture of proteins and cells as biofoulants, both should be employed to assess the fouling-resistance of a membrane. Some of the major tests run by the authors of the studies relevant to this review are summarized below, and some essential membrane characteristics and performances listed in Table 5. From this table, we could clearly see an increasing trend in the antifouling properties of the modified membranes as their hydrophilic properties were enhanced, which essentially constituted the increased rejections of various foulants, and overall better performance.

4.1. Hydrophilicity Tests

As it is for dense interfaces, hydrophilicity is an important factor in determining the ability of a porous membrane to resist fouling [367]. The water contact angle (WCA), hydration capacity (HC), and strength of the hydration layer of the membranes may be measured to characterize the surface hydrophilicity, its ability to trap water, and the stability of the hydration layer, respectively.
WCA is commonly measured with a contact angle analyzer instrument employing the static sessile drop technique. It is likely the most employed test as it is readily carried out. Moreover, it can be dynamically monitored, that is, the evolution of WCA with time can be tracked, which is useful when wetting is not instantaneous. It has been recently reported that with VIPS PVDF membranes, the WCA could be decreased by 15–20° within 80 s [123]. This is associated with the intrinsic hydrophobicity of the PVDF material enhanced by the particular morphology of PVDF membranes prepared by VIPS, which are highly porous and rough, two properties that contribute to prevent or delay wetting. In general, lowering the surface roughness facilitates wetting, and so, permits the reduction in the adhesion forces between the membrane and foulants [190,368,369]. The WCA of PVDF membranes prepared by VIPS [123] is much larger than that of similar membranes obtained by LIPS [357], clear evidence of differences in surface structuring. Hence, some “induction” time may be required for the water molecules to interact with the antifouling moieties on the surface of such rough membranes. Though the WCA provides intelligence on the surface hydrophilicity, one may question its actual relevance when membranes are to be contacted with aqueous medium for some extended period of time, typically in wastewater treatments. As just discussed, the WCA of VIPS membranes, although modified, remains high. Yet, they are able to trap a significant amount of water after prolonged contact. This suggests that water can penetrate and wet the pores of the membrane, provided that it has been modified. Although surface segregation is known to occur during the preparation of antifouling membranes by phase-inversion methods [370], these processes maintain bulk modifications which means that a significant proportion of the antifouling additive is trapped inside the matrix. These trapped antifouling moieties contribute to water trapping in the bulk and are also likely to reduce fouling inside the membrane. Hydration capacity tests or swelling tests, which are gravimetric tests, are readily performed and can provide relevant related information. These tests do not require any equipment other than a precision balance with which the weight of the samples before and after immersion in water are measured. Then, one can report the amount of water trapped per unit volume of the membrane (hydration capacity) or normalized using the dry weight the difference between the weight of the hydrated membrane and the weight of the dry one (swelling ratio also sometimes referred to as equilibrium water content). This will likely provide a useful assessment of the hydrophilicity of a porous membrane. Finally, it is worth mentioning that the strength of the hydration layer can help in discussing the antifouling properties of a membrane. Following the work of Ishihara’s group [167], we have been evaluating the strength of the hydration layer of some blending modified membranes using differential scanning calorimetry measurements [90,123]. These tests, combined with the results of hydration tests, permit the determination of the proportion of non-freezable water. The higher it is, the stronger the hydration layer. Mapping FT-IR can also be used to assess hydration differences between samples which correlates to the strength of the hydration layer. We used this approach after drying samples, to evaluate whether water could still be detected on the surface of membranes, and to compare several antifouling additives [117]. It is more of a qualitative assessment but still permits the highlighting of clear differences regarding the ability of antifouling materials to retain water, and also proved that zwitterionic materials performed better than PEGylated ones.

4.2. Static Fouling Tests

Static fouling tests may be conducted, as carried out for model dense interfaces through which no permeation occurs. These tests require the incubation of the membrane samples with a solution containing the foulants (typically proteins or cells). They mostly permit the evaluation of the surface antifouling properties. If proteins are used, the advantage over dynamic tests is that they are simpler to implement and require lower volumes of solutions. This also significantly lowers the costs when human plasma proteins such as fibrinogen (FN), human serum albumin or γ-globulin are used to test the antifouling properties of biomedical membranes [41,361]. Furthermore, because cells such as gram-negative bacteria or human blood cells can be easily deformed under a pressure gradient sometimes resulting in cell lysis, carrying static adsorption tests eliminates the influence of pressure, and permits us to solely focus on the material–biofoulants interactions.
Table 5. Some performances of blending modified membranes reported in the past decade.
Table 5. Some performances of blending modified membranes reported in the past decade.
Class of AdditiveMatrix PolymerAntifouling AdditiveProcessWater Contact Angle (o)Pore Size (nm)Porosity (%)Pure Water Permeance * (L/m2·h·bar)Flux Recovery Ratio (%)Foulant Adsorption a or Rejection r (% or (μg/cm2) d)Ref.
VirginModifiedVirginModifiedVirginModifiedVirginModifiedVirginModifiedVirginModified
Polymer/organic-basedCAPAALIPS71.525.012.99.644.675.615.319.088.297.6r HA (95.7)r HA (99.9)[108]
PEIPVPVIPS96.456.128.550.44.213.64.836.869.681.0a BSA (170.0) da BSA (80.0) d[94]
PESCNCLIPS66.043.012.09.0--33.360.951.090.0r BSA (93.0)r BSA (97.0)[100]
PESPANLIPS68.456.8----24.655.657.986.1r BSA (93.8)r BSA (87.1)[371]
PESPAN/PEGLIPS68.4-----24.6202.257.990.7r BSA (93.8)r BSA (81.4)[371]
PESPVALIPS79.063.011.234.954.265.5-131.552.092.6r BSA (81.0)r BSA (61.2)[106]
PESPVPLIPS------c 72.01219.5c 80.898.5-r HA (95.0)[43]
PLAPVPLIPS82.134.188.642.979.257.1185.719.357.093.0r BSA (57.0)r BSA (92.0)[358]
PSfPEG400LIPS87.779.8--18.748.40.8420.0---
-
r BSA (90.0)/ r Pepsin (73.0)[42]
PSfzP(S-r-4VP)LIPS120.0100.06.310.473.274.15.711.740.063.0r BSA (80.8)r BSA (95.6)[120]
PSfPEGMAVIPS63.027.09.310.880.873.6110.0512.067.684.5a BSA (15.1) da BSA (4.0) d[372]
PVCLigninLIPS106.741.519.725.576.084.9111.6347.720.492.3r HA (73.0)r HA (98.9)[142]
38.181.3r Oil (68.1)r Oil (97.4)
PVDFSPANILIPS92.029.0----97.2160.066.299.2r BSA (90.0)/a BSA (30.0) dr BSA (95.0)/a BSA (3.0) d[124]
PVDFCNCLIPS81.374.084.5155.0 9.8206.971.682.5r BSA (83.3)r BSA (88.2)[98]
PVDFMPC-derivativeVIPS137.3113.7108.1155.872.076.3c 10871143c 1742--[122]
PVDFMPC-derivativeVIPS13711430207270c ≈900≈900c 1738--[373]
PVDFPS-b-PEGMALIPS85.0-25.551.864.379.843.358.356.091.0a FN (100.0)/a γ-globulin (100.0)/a HAS (100.0)a FN (19.0)/a γ-globulin (35.0)/a HAS (29.0)[41]
PVDFPS-b-PEGMAVIPS1221182608207279500013,000≈10≈82a FN (100.0)/a BSA (100.0)/a LY (100.0)a FN (≈30)/a BSA (≈5)/a LY (≈15)[153]
PVDFPS-b-PEGMAVIPS1221182608207279500010,000≈25≈76.9r MA (>99.7)r MA (>99.7)[153]
PVDFPVALIPS72.269.025.3100.462.857.638.147.583.086.0r BSA (35.0)r BSA (5.2)[107]
PVDFPVDF-g-PSBMALIPS89.067.0----121.9239.151.581.2a BSA (110.0) da BSA (40.0) d[118]
PVDFPS-r-PEGMAVIPS140.047.0583.0513.075.471.9--16.029.0rE. coli (92.5)rE. coli (99.0)[374]
PVDFPS-r-PEGMATIPS135.770.0≈210≈210--≈2400≈2400≈4574.0a BSA (75.0)a BSA (18.0)[365]
PVDFSM-derivativeVIPS145.067.0510.0430.057.062.0≈9000≈12,00036.090.0a BSA (100)a BSA (35)[363]
PVDFPS-r-PEGMA-r-PSBMAVIPS129.0102.0140.070.066.572.91223.01146.066.091.0a BSA (100)/a FN (100)a BSA (12)/a FN (15)[117]
PVDFPS-r-PEGMA-r-PSBAAVIPS140.0112.0≈299≈806755c 200011005373rE. coli (83.5)/a FN (100)/a blood (100)rE. coli (89.6)/a FN (20)/a blood (20)[123]
PVDFPMAA-r-PEGMA-r-SBMAVIPS139.090.0560.0150.070.062.0c 2500900.0-54.0a FN (100)/a E. coli (100)/a blood (100)a FN (12)/a E. coli (5.6)/a blood (5.1)[116]
PVDFzP(S-r-4VP)VIPS≈143≈132250100≈75≈60c ≈625≈190012.069.0aE. coli (100)/a blood (100)aE. coli (15)/a blood (10)[119]
Metal-basedBCMZrO2LIPS41.933.636.539.377.379.8286.1321.565.090.6r BSA (71.6)r BSA (91.2)[181]
PANSiO2-DOPALIPS68.032.0--51.777.1426.71075.049.075.0r BSA (94.0)r BSA (98.8)[239]
PESFe2O3-Mn2O3LIPS73.067.040.045.558.074.0208.0396.064.077.0r BSA (97.0)r BSA (96.0)[231]
PESZrO2LIPS73.652.3----8.283.65497.2r BSA (97.2)r BSA (92.7)[179]
PESZnO-NPLIPS77.960.0--61.069.07.812.039.074.1--[82]
ZnO-NRLIPS77.954.0--61.071.07.812.539.070.2
PVCZnOLIPS67.554.59.312.167.979.8106.5201.069.391.8r BSA (90.2)r BSA (97.5)[236]
PVDFSiO2@GOTIPS--41.220.1--268.5182.648.095.0r BSA (63.5)r BSA (91.7)[366]
PVDFTiO2/PEGLIPS74.469.087.086.052.948.672.873.1----[38]
PVDFTiO2LIPS85.470.2----158.0350.047.588.2r BSA (57.0)r BSA (95.0)[190]
PVDFZrO2-g-PACMOLIPS93.066.031.617.077.658.436.282.438.097.0r Oil (74.9)r Oil (99.9)[180]
Carbon allotropesPEESGOLIPS96.472.345.672.743.265.630.753.962.583.2--[345]
PES **GOLIPS91.067.03.05.152.158.525.0225.0----[304]
GOLIPS85.072.06.68.629.161.0239.6305.2
PVCPP-MWCNTsLIPS67.761.24.24.377.584.030.037.564.870.6r BSA (96.0)r BSA (98.0)[302]
PVCGQDsLIPS65.073.02.843.0157.255.312.219.168.8>80.0r BSA (>98.0)r BSA (>98.0)[308]
PVDFCNDsLIPS76.865.323.833.147.663.7491713585r BSA (>85.0)r BSA (>95.0)[288]
PVDFGOLIPS71.070.070.0115.047.552.5137.0203.5----[304]
PVDFGOLIPS74.270.2484.01034.059.080.04794--r DOC (8.65)r DOC (11.30)[305]
PVDFGO/TiO2LIPS79.061.048.165.269.683.1158.1487.843.071.1r BSA (80.0)r BSA (92.5)[369]
PVDFO-MWCNTTIPS106.898.0--84.583.8270.7164.5-82.7r BSA (68.5)r BSA (90.8)[23]
a adsorption; c commercial membrane; d μg/cm2; r rejection. * Often determined at 1 bar. However, some researchers made use of different transmembrane pressure and the overall pure water permeances were normalized. The table aims at comparing the performances of the pristine and modified membranes under similar operating conditions. So, one is invited to refer to the cited work for more details regarding the experimental protocols. ** Different solvents were used.

4.2.1. Protein Adsorption Tests

BSA is frequently used for static (and dynamic) tests [238,357,365,372] because it is inexpensive. Moreover, it is stable in water-miscible polar solvents, and it is biodegradable [160,375,376,377]. As it possesses absorption peaks in the ultraviolet region, its concentration in the incubation solution can be tracked with a UV-visible spectrometer. Then, a mass balance permits the evaluation of the amount of protein that interacted with the membrane samples. It is somewhat an indirect evaluation of biofouling. A more direct visualization of protein adsorption in the membrane has been reported by the group of Matsuyama who prepared blending modified TIPS membranes [365]. The method consists of utilizing a fluorescent derivative of BSA (fluorescein isothiocyanate conjugate BSA or FITC-BSA). In doing so, the protein can also be directly visualized with a fluorescent microscope or a confocal microscope, and an image analysis software can help in quantifying the change in fluorescence intensity. In our group, we have used other proteins such as lysozyme (LY) [357,361] or fibrinogen and other plasma proteins [41,361]. The method with LY is similar to that involving BSA. LY was used to evaluate the effect of the protein Mw on biofouling. For FN and plasma proteins, an enzyme-linked immunosorbent assay is performed. These plasma proteins are used when membranes are intended to be applied in biomedical devices, in contact with blood, as the early adhesion of FN can then trigger mechanisms of blood cells adhesion and blood coagulation. Therefore, it is relevant to use these proteins as model biofoulants rather than BSA.

4.2.2. Cell Attachment Tests

Bacteria can be utilized to study the interactions between large biofoulants and the membrane material in static conditions. As some can be potentially harmful, the advantage of using static tests over dynamic filtration tests is that volumes needed are much smaller in the former case. Small membrane samples can be positioned in well-plates and then incubated with <1 mL of bacterial solution. Then, the well-plate can be disposed of. Contrariwise, large volumes of bacterial solution would be needed for filtration tests, which would then require careful cleaning and disinfection of the device. Bacteria of different sizes or shapes and belonging to the class of gram-positive or gram-negative have been used in such tests [373] although Escherichia coli is the most employed microorganism for testing the antifouling/antibacterial properties of membranes [209,357,378]. Solid culture medium containing agar can then be employed to quantify the resistance to bacteria of the modified membrane, relative to the pristine one [209,361]. The extent of bacterial attachment can be visualized by various microscopy techniques including SEM [209,357], but also by fluorescence microscope or confocal microscope [122,373]. Genetically modified bacteria with a fluorescent protein have been used, which facilitates the sampling before observation since staining is unnecessary [122]. Moreover, different incubation times of the samples with the bacterial medium may provide information on the resistance of the membranes to biofilm formation.
Regarding blending modified membranes with potential applications in blood-contacting devices modified with zwitterionic materials [116,117,122,123], static adsorption tests making use of platelet, red blood cell or white blood cell concentrates, but also whole blood, can be implemented. These tests provide useful information associated with the hemocompatibility of the membrane surface. With whole blood or platelet concentrates, lack of blood compatibility of the membrane can be visualized by SEM by the formation of a fibrillar network arising from platelet activation [117]. As red blood cells concentrate alongside the incubation medium, numerous cells of typical biconcave shape can be observed if the membrane surface is not slippery. It can be interesting to use confocal microscope (again providing fixing/staining of the cells with a glutaraldehyde solution and a solution of 4′,6-diamidino-2-phenylindole) and extract 3D images to locate the position of the cells. In one of our early works, we used this technique that revealed that the cells were not all located on the membrane surface, but instead, managed to penetrate in the deeper layers of the membranes despite the lack of a pressure gradient [361]. Although the mechanism is unclear, the large pores of VIPS membranes, associated with the flexibility of the cell wall of the cells certainly contributed to their penetration and detection beneath the surface.

4.3. Dynamic Fouling Tests

The majority of works on the development of antifouling membranes showcases antifouling performances during filtration. This makes sense since membranes are almost only used to separate species during filtration. Thus, blending modified porous membranes, which were the key focus, are no exception. In Table 1, we listed the FRR of membranes, obtained after one or several water/biofouling solution/water filtration cycles. As also listed in the references, most studies report BSA protein as the model biofoulant. This can be explained by its availability at low-cost, its safety-in-use compared to bacterial solutions or blood solutions, but also by its suitability to be rejected by ultrafiltration membranes. Undeniably, the most employed process to prepare blending modified membranes is LIPS which commonly leads to membranes falling in the UF range. For all of these reasons, using BSA is appropriate. Nevertheless, there is no standard procedure for studying fouling during filtration. For example, Li et al. performed three complete water/BSA filtration cycles using a 0.5 g/L BSA solution [288] while Yuan et al. performed one cycle using a 3 mg/L BSA solution [106]. In other works, Arumugham et al. conducted five cycles with a 0.5 g/L BSA solution [231] while Xu et al. studied fouling of their blending modified UF photocatalytic membrane reactor using a 1 g/L BSA solution and conducting one cycle also followed by an irradiation step (aiming at degrading the adsorbed proteins from the pores) [369]. The lack of a standard procedure makes it difficult to compare the antifouling performances of a membrane during filtration, in light of existing literature. So, researchers tend to use controls which are the pristine (unmodified) membrane or a commercial membrane possibly of the same composition in terms of main polymer (though difficult to check since actual material compositions are not always revealed by manufacturers) and similar pore size. In order to assess the effect of the modification on fouling mitigation, researchers determine the FRR, but also the reversible flux decline ratio (Rr), the irreversible flux decline ratio (Rir), and the total flux decline ratio (Rt). Their expressions are reminded below:
F R R ( % ) = J w , f J w , i × 100
R r ( % ) = ( J w , f J f o u l a n t , f J w , i ) × 100
R i r ( % ) = ( 1 J w , f J w , i ) × 100  
R t ( % ) = R r + R i r = ( J w , i J f o u l a n t , f J w , i ) × 100
where Jw,i and Jw,f are the initial and final (after membrane cleansing at the end of the cyclic filtration test) pure water flux, respectively, while Jfoulant,f represents the flux recorded during the final cycle involving the solution of foulant. If efficient, the modification will have a clear impact on both FRR and R i r , meaning that some appropriate washing procedure will permit the recovery of the membrane initial permeability as it minimizes irreversible interactions. Then again, however, there is no clear standard procedure for washing the membrane in these tests. Rejections of the foulant material (BSA or others) are also sometimes reported, but the aim of the modification remains to minimize fouling while maintaining other properties. We have discussed in previous sections that the antifouling material affects membrane structure. As a result, it influences both the membrane permeability and the rejection, since parameters such as the pore size, the porosity or the tortuosity associated with the morphology play a major role on the membrane performances. To eliminate these effects and visualize more efficiently the effect of the material on antifouling performances during cyclic filtration, the flux can be normalized using the initial pure water permeability of the membrane [238,365]. This approach also facilitates the comparison with control commercial membranes whose permeability is often different from that of the modified membrane fabricated in the lab. Indeed, the normalized permeability varies in the range 0–1 with 0 indicating total pore blockage, with 1 being the maximum flux recovery ratio (no fouling).
Although BSA dominates, other foulants are used. Humic acid has been employed [43,108,122,142], as well as oil [63,180], extracellular polymeric substances (EPS) [238], diluted commercial milk [106], microalgae (MA) [153,373] or bacteria [123,363]. Microalgae and bacterial solution are particularly appropriate to the performance evaluation of MF membranes, typically prepared by VIPS or TIPS, given their characteristic size. Nevertheless, careful cleaning and disinfection of the equipment has to be performed after bacterial filtration test. Interestingly, despite significantly smaller pore size than the characteristic size of bacteria tested (Escherichia coli), VIPS membranes were not permitted to reach 100% rejection, due to the deformability of the cell wall [123]. Thus, one cannot just rely on size-sieving effects in order to achieve safe disinfection of water. Instead, antibacterial agents (positively charged polymers or nanoparticles) would be needed that would kill during filtration such that potential bacteria managing to permeate through the membrane would end up dead.
Different from most studies, some teams have utilized their antifouling membranes to treat more complex feeds, sometimes from natural effluents or wastewater. Notably, Yu et al. treated the supernatant liquor obtained from a sedimentation tank located in a municipal wastewater plant [209]. Xia and Ni used micro-polluted raw water from a river [305], while Yong et al. used oily wastewater from a sewage plant belonging to an oil corporation [142]. Moghadam et al. used EPS, composed of multiple organic substances triggering fouling in membrane bioreactors [238]. The complex nature of the feed made of numerous potential foulants truly challenges the membranes and provides actual evidence of their true potential in real life applications. Nevertheless, this approach remains rare as most tests conducted involve a synthetic feed made of one type of foulant in an aqueous solvent.
Finally, a hybrid static/dynamic test was employed to study biofouling by blood components [41,373]. It consisted of recording the water permeability followed by a long-incubation of the membranes with platelet-poor-plasma (PPP). Then, the permeability was recorded immediately after and before cleaning the membranes. A final water permeability cycle ended the test. The incubation of the membrane with PPP was carried out in a similar way as static adsorption tests, and thus, involved low volume of blood-derivative. Given its composition, PPP has a strong “fouling power”. In addition, because incubation was directly followed by a water permeability test without prior cleaning, this hybrid fouling test designed for biomedical membranes severely challenged the polymer matrix.

4.4. Oil Fouling Tests

Oil fouling poses a serious threat to the long-term performance of the membranes. Distinct from biofoulants, oils tend to coalesce and spread across the surface. Since oil repels water, making the surface of the membrane superhydrophilic to prevent oil fouling seems to be the best option. However, this is no easy feat because dissimilar from biofoulants, oils tend to coalesce and spread across the surface. Although it will not form interactions with the hydrated surface of the membrane, the oils would sit on top of it and form an oil layer which hinders the passageways of the water.
Previous studies have placed much emphasis on the wettability tests of the surface which includes, static contact angles of [379]: water in air, water under oil, oil underwater, and sliding contact angle tests in air [380]. For dynamic filtration of oil/water separation, Guo et al. prepared a series of 500 mL feed with 99:1 ratio of water and soybean with 10 mg of sodium dodecyl sulfate (SDS) surfactant [381]. They then performed cyclical filtration of emulsion and water for 18 cycles (9 cycles for oil/water emulsion and 9 cycles of pure water). Amid et al. [263] also evaluated the oil antifouling properties of their modified membranes using three different of olive oil (100 ppm, 150 ppm, and 200 ppm) in the SDS stabilized water-rich emulsions. Other studies have also reported the use of surfactant-stabilized emulsions to test the antifouling properties of membranes [382,383,384]. By using more complex synthetic emulsion feeds, this precedes the development of more effectively robust and highly foulant resistant membranes with wider applicability (e.g., cosmetics and pharmaceutical industries).

5. Conclusions and Future Perspectives

This paper intended to review the recent progress made on the development of antifouling membranes by blending modification, that is, obtained after blending the antifouling additive with a polymer/solvent system before inducing phase separation. We briefly discussed the different phase-inversion processes used to design such membranes, as well as classifying the type of additives (polymer/organic-based, metal-based, ceramic-based, carbon allotrope and porous nanomaterials, and hybrid nanomaterials), examined their effect on fouling and finally introduced common tests carried out by researchers to assess fouling on membranes for wastewater treatment or biomedical-related applications. We saw that the addition of antifouling material to the casting solution in view of improving the antifouling properties is almost always accompanied by an effect on the structure of the membrane due to the change in thermodynamic stability and kinetic of phase-inversion. In the case of LIPS membranes, it often results in the formation of larger macrovoids. In the case of VIPS membrane applied to PVDF (the most used polymer), the additive enables to reduce the size of nodules, while in the case of TIPS, it may affect the dominating phase separation mechanism (i.e., L/L vs. S/L phase separation) especially when nanoparticles are added. Defects and pore blockage are also a commo observation when nanoparticles are added. Thus, changes in permeability and rejection, compared to the pristine (unmodified) membrane is very frequently reported.
Surely, significant progress has been made in the last decade. Yet, after surveying the recent literature, it is clear that efforts still need to be made concerning the following aspects:
Mitigating fouling while maintaining rejection: Thanks to the work of multiple teams worldwide, we have access to a library of materials that efficiently reduce fouling. Nevertheless, the rejection as measured with the pristine membrane is rarely maintained. The blending modified membrane can be seen as a system with distinct retention properties. As such, we may question its domain of application. Can it be used for the same separation or the same treatment as the initial pristine membrane? We believe that researchers should now not just look at the antifouling properties, but instead try to adjust the pore size of the antifouling membrane to the same value as that of the pristine membrane. This means that membrane formation mechanisms should be better controlled (if the antifouling materials accelerates mass transfers during phase-inversion, then one could adjust the non-solvent power of the non-solvent bath by using solvent), and care should be put on fine-tuning the viscosity of the casting solution containing the antifouling material.
The preparation of blending modified membranes in the MF range: LIPS clearly dominates. This has also been the case in the reports published in the past 10 years. This is understandable considering its ease of implementation (in the lab, no expensive equipment is needed except a casting knife to control the initial clearance). However, it leads to UF membranes. Complete wastewater treatment with membrane technology can be achieved providing also the use of smaller (NF) and larger (MF) pore size membranes. Surely, it is very difficult to prepare NF membranes in one step. Moreover, as we have seen, the antifouling material often acted as a pore-former, and it becomes trickier to prepare blending modified membrane for antifouling by phase-inversion. However, how about MF membranes? Both VIPS and TIPS can enable their preparation. The development of blending modified membranes by VIPS has been reported by very few teams including ours. Yet this technique is appropriate to form bi-continuous structures that maintain high flux/mechanical stability and avoid the formation of typical macrovoids (seen in LIPS membranes that weaken the structure). As for the TIPS process, the limitation might be the high temperature of the initial dope that may damage antifouling polymers. However, some nanoparticles have outstanding thermal stability, such that they could be potential candidates for the formation of MF antifouling membranes by blending modification. Too little has been carried out on these aspects and there is room for improvements [23].
The development of blending modified membranes with zwitterionic materials: this class of materials is seen as the latest generation of antifouling materials, providing better stability than PEGylated materials in complex media, and stronger hydration. In recent few years, they have often been used for the surface modification of membranes and dense interfaces to be applied in biomedical devices in contact with biological fluids [385,386,387]. Nevertheless, their use in direct combination with polymers remains limited. The large polarity of the zwitterionic moieties makes it difficult to solubilize them in polymer/solvent systems and prepare homogeneous casting solutions. Yet, addressing solubility issues could further extend the span of applications of blending modified membranes in the biomedical fields (blending membranes for leukodepletion, membranes used as wound-dressings, etc.).
The use of green solvent in the casting solution to form green blending modified membranes: nowadays, the search for green solvents that could be applied to the preparation of phase-inversion membranes is an important direction explored by numerous teams. Although it was beyond the scope of this survey, most papers reviewed here make use of toxic solvents. Thus, not only should we find alternatives to toxic solvents for phase-inversion membranes, but also these greener options should permit the preparation of antifouling membranes in one step. Dimethyl sulfoxide is a greener option and has recently been explored for the making of antifouling membranes by blending modification [304,374], as well as cyrene has been reported in a PES/GO blend [304], but unquestionably a lot more must be achieved to make antifouling membrane production green.
Improving the miscibility of the additives in the polymer matrix and preventing agglomeration: Although physical blending provides a simple approach to modifying both the surface and bulk of the membranes, it is limited by considerable drawbacks such as miscibility issues of some polymeric additives, and agglomeration of inorganic nanoparticles. These problems result in poor dispersion of the additives in the matrix which results in leaching thus, decreasing the antifouling properties of the membranes. In addition, the pores the “leached additives” generate in the polymer matrix albeit, could increase permeability, creates pores on the surface which lowers the rejection, and also induces the formation of large macrovoids along the cross-section which decreases the mechanical strength of the membranes. Moreover, the agglomeration of inorganic additives is found to have induced increased surface roughness which significantly affects the antifouling properties of the membrane as it decreases steric repulsion against foulants. In order to mitigate these problems, recent works have utilized in situ polymerization [372,388,389,390,391,392,393,394,395,396,397,398,399,400,401,402,403], in situ self-assembly of polymeric nanoparticles [404], in situ growth or in situ synthesis [399,405], and in situ generation of nanomaterials [391,406] in the casting solution as strategies to help counter the agglomeration of nanoparticles or nanomaterials in the polymer matrix thus, enhancing the polymer–additive interaction [33]. In order to achieve the latter, the precursors of the nanomaterials must be soluble in the dope casting solution and can react with the non-solvent in the coagulation bath [34,349].
Testing the leachability of inorganic additives: Although tremendous studies have shown the effectiveness of using inorganic nanoparticles in mitigating foulants, it is undeniable that they could still pose massive threats as they could be easily leached out. When this happens, millions of lives could be in danger of metal poisoning. According to previous works [399,405], growth of these nanoparticles in the casting solution could help improve the compatibility of the nanoparticles in the host matrix. However, more studies involving long-term stability are needed in order to test this approach. In addition, leaching tests of these additives should be carried out using sophisticated techniques such as Inductively Coupled Plasma (ICP) Spectroscopy to ensure precise measurements [399].
The need for standard procedures to assess the antifouling properties of membranes: This does not just concern blending modified membranes, but also antifouling membranes prepared by other processes (coating, grafting, etc.). We emphasized some different testing procedures that exist using BSA and reported in the works analyzed for this review. The concentration of the protein, the number of cycles, the duration of these cycles varies from one study to another, or the transmembrane pressure applied such that it is extremely challenging to assess the actual antifouling properties of a membrane. We should be more precise and define some standard testing procedures enabling a fair comparison of the membrane performances. Although this could be a bit tricky given that most of filtration systems in these studies are “in-house”, many parameters would have to depend on the filtration set-up. However, there are still few parameters that could be fixed, and could be established as “protocols” for the antifouling tests such as the concentration of the foulants, and the number of cycles for the “foulant/water” filtration. A study on the domestic wastewater of Shanghai showed that the concentration of total chemical oxygen demand (COD) ranges from 204–440 mg/L (population of 12,000) [407]. Meanwhile, in Indonesia, the COD of greywater wastewater consist of 232–780 mg/L [408]. Considering the COD values of domestic wastewater from these studies, for lab-scale experiments, it would be more logical to set the feed concentration closer to the higher COD recorded to simulate the extreme conditions of wastewater, that is, setting the foulant concentration to 1000 mg/L. In doing so, we could assure that the antifouling properties of the modified membranes could withstand “real-world” conditions. Additionally, in terms of cyclical filtration tests, 3–5 cycles should be carried out (with three repetitions for each condition) so that the stability of the antifouling membranes could be well investigated.
Testing the membranes with more foulants and complex feeds: This again does not just concern blending modified membranes. While surveying the recent literature, we noticed the dominance of one type of feed solution (BSA solution). Testing the antifouling properties with various foulants each time a system is presented would provide supporting evidence of its efficiency to mitigate fouling in different situations. Possibly, natural feeds that combine multiple foulants should be used as well. A few studies made use of complex natural feeds, but more should be carried out in this direction in view of justifying the need for upscaling the production of antifouling membranes at an industrial scale, clearly highlighting their potential range of applications, and showcasing how they can benefit people.

Author Contributions

Conceptualization, A.V.; writing—original draft preparation, T.A.G., I.V.M. and A.V.; writing—review and editing, A.V., T.A.G. and I.V.M.; funding acquisition. A.V.; formal analysis and conceptualization, A.V. and Y.C.; supervision A.V. and Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors express their sincere gratitude to the National Science and Technology Council for their financial support through the grants 108-2221-E-033-006-MY3 and 109-2628-E-033-001-MY3.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

BCMBamboo cellulose membrane
BSABovine serum albumin
CACellulose acetate
CBCarboxybetaine
CNCCellulose nanocrystal
CNDCarboxylated nanodiamond
DMSODimethyl sulfoxide
DOCDissolved organic carbon
DOPADopamine
EPSExtracellular polymeric substances
FNFibrinogen
FRRFlux recovery ratio
GNGraphene
GOGraphene oxide
GQDsGraphene quantum dots
HAHumic acid
HCHydration capacity
HSAHuman serum albumin
LYLysozyme
MAMicroalgae
MFMicrofiltration
MMMMixed matrix membrane
MPC2-methacryloyloxyethyl phosphorylcholine
MWCNTMultiwalled carbon nanotubes
NFNanofiltration
OMWCNTOxidized multi-wall carbon nanotube
PAAPoly (acrylic acid)
PACMOPoly(N-acryloylmorpholine)
PANPolyacrylonitrile
PANIPolyaniline
PCPhosphorylcholine
PEPolyethylene
PEESPoly(ether ether sulfone)
PEGPoly(ethylene glycol)
PEGMAPoly(ethylene Glycol) Methyl Ether Methacrylate
PEIPolyetherimide
PESPoly(ether sulfone)
PLAPolylactic acid
PPPolypropylene
PSPolystyrene
PSBMAPoly(sulfobetaine methacrylate)
PSfPolysulfone
PTFEPolytetrafluoroethylene
PVAPoly(vinyl alcohol)
PVCPoly(vinyl chloride)
PVDFPolyvinylidene fluoride
PVPPolyvinylpyrrolidone
PWFPure water flux
rGOReduced graphene oxide
RirIrreversible flux ratio
RrReversible flux ratio
RtTotal flux ratio
ROReverse osmosis
SBSulfobetaine
SBAASulfobetaine methacrylamide
SBMASulfobetaine methacrylate
SEMScanning electron microscopy
SMStyrene maleic anhydride
SPANISulfonated polyaniline
UFUltrafiltration
WCAWater contact angle
ZnO-NPZinc oxide nanoparticles
ZnO-NRZinc oxide nanorods
zP(S-r-4VP)Zwitterionic poly(styrene-random-4-vinylpyrridine)

References

  1. Zahid, M.; Rashid, A.; Akram, S.; Rehan, Z.A.; Razzaq, W. A comprehensive review on polymeric nano-composite membranes for water treatment. J. Membr. Sci. Technol. 2018, 8, 1–20. [Google Scholar] [CrossRef]
  2. Wang, D.-M.; Lai, J.-Y. Recent advances in preparation and morphology control of polymeric membranes formed by nonsolvent induced phase separation. Curr. Opin. Chem. Eng. 2013, 2, 229–237. [Google Scholar] [CrossRef]
  3. Ang, W.L.; Lau, W.J.; Tan, Y.H.; Mahmoudi, E.; Mohammad, A.W. An Overview on Development of Membranes Incorporating Branched Macromolecules for Water Treatment. Sep. Purif. Rev. 2021, 52, 1–23. [Google Scholar] [CrossRef]
  4. Sajid, M.; Jillani, S.M.S.; Baig, N.; Alhooshani, K. Layered double hydroxide-modified membranes for water treatment: Recent advances and prospects. Chemosphere 2022, 287, 132140. [Google Scholar] [CrossRef] [PubMed]
  5. Dong, X.; Lu, D.; Harris, T.A.L.; Escobar, I.C. Polymers and solvents used in membrane fabrication: A review focusing on sustainable membrane development. Membranes 2021, 11, 309. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, Z.; Wang, W.; Xu, X.; Liu, X.; Li, Y.; Zhang, P. Enhanced morphology and hydrophilicity of PVDF flat membrane with modified CaCO3@SMA additive via thermally induced phase separation method. J. Ind. Eng. Chem. 2021, 107, 444–455. [Google Scholar] [CrossRef]
  7. Khan, B.; Zhan, W.; Lina, C. Cellulose acetate (CA) hybrid membrane prepared by phase inversion method combined with chemical reaction with enhanced permeability and good anti-fouling property. J. Appl. Polym. Sci. 2020, 137, 49556. [Google Scholar] [CrossRef]
  8. Obotey Ezugbe, E.; Rathilal, S. Membrane technologies in wastewater treatment: A review. Membranes 2020, 10, 89. [Google Scholar] [CrossRef]
  9. Cui, Z.; Jiang, Y.; Field, R. Fundamentals of pressure-driven membrane separation processes. In Membrane Technology; Elsevier: Amsterdam, The Netherlands, 2010; pp. 1–18. [Google Scholar]
  10. Lee, A.; Elam, J.W.; Darling, S.B. Membrane materials for water purification: Design, development, and application. Environ. Sci. Water Res. Technol. 2016, 2, 17–42. [Google Scholar] [CrossRef]
  11. Van der Bruggen, B.; Vandecasteele, C.; Van Gestel, T.; Doyen, W.; Leysen, R. A review of pressure-driven membrane processes in wastewater treatment and drinking water production. Environ. Prog. 2003, 22, 46–56. [Google Scholar] [CrossRef]
  12. Maggay, I.V.; Chang, Y.; Venault, A.; Dizon, G.V.; Wu, C.-J. Functionalized porous filtration media for gravity-driven filtration: Reviewing a new emerging approach for oil and water emulsions separation. Sep. Purif. Technol. 2021, 259, 117983. [Google Scholar] [CrossRef]
  13. Díez, B.; Rosal, R. A critical review of membrane modification techniques for fouling and biofouling control in pressure-driven membrane processes. Nanotechnol. Environ. Eng. 2020, 5, 15. [Google Scholar] [CrossRef]
  14. Koulivand, H.; Shahbazi, A.; Vatanpour, V. Fabrication and characterization of a high-flux and antifouling polyethersulfone membrane for dye removal by embedding Fe3O4-MDA nanoparticles. Chem. Eng. Res. Des. 2019, 145, 64–75. [Google Scholar] [CrossRef]
  15. Awad, E.S.; Sabirova, T.M.; Tretyakova, N.A.; Alsalhy, Q.F.; Figoli, A.; Salih, I.K. A mini-review of enhancing ultrafiltration membranes (UF) for wastewater treatment: Performance and stability. Chem. Eng. 2021, 5, 34. [Google Scholar] [CrossRef]
  16. Ji, Y.-L.; Yin, M.-J.; An, Q.-F.; Gao, C.-J. Recent developments in polymeric nano-based separation membranes. Fundam. Res. 2021, 2, 254–267. [Google Scholar] [CrossRef]
  17. Kumar, S.; Guria, C.; Mandal, A. Synthesis, characterization and performance studies of polysulfone/bentonite nanoparticles mixed-matrix ultra-filtration membranes using oil field produced water. Sep. Purif. Technol. 2015, 150, 145–158. [Google Scholar] [CrossRef]
  18. Tang, Y.; Lin, Y.; Ma, W.; Wang, X. A review on microporous polyvinylidene fluoride membranes fabricated via thermally induced phase separation for MF/UF application. J. Membr. Sci. 2021, 639, 119759. [Google Scholar] [CrossRef]
  19. Remanan, S.; Sharma, M.; Bose, S.; Das, N.C. Recent advances in preparation of porous polymeric membranes by unique techniques and mitigation of fouling through surface modification. Chem. Sel. 2018, 3, 609–633. [Google Scholar] [CrossRef]
  20. Ismail, N.; Venault, A.; Mikkola, J.-P.; Bouyer, D.; Drioli, E.; Kiadeh, N.T.H. Investigating the potential of membranes formed by the vapor induced phase separation process. J. Memb. Sci 2020, 597, 117601. [Google Scholar] [CrossRef]
  21. Cui, Z.; Hassankiadeh, N.T.; Lee, S.Y.; Lee, J.M.; Woo, K.T.; Sanguineti, A.; Arcella, V.; Lee, Y.M.; Drioli, E. Poly(vinylidene fluoride) membrane preparation with an environmental diluent via thermally induced phase separation. J. Membr. Sci. 2013, 444, 223–236. [Google Scholar] [CrossRef]
  22. Ichi Sawada, S.; Ursino, C.; Galiano, F.; Simone, S.; Drioli, E.; Figoli, A. Effect of citrate-based non-toxic solvents on Poly(vinylidene fluoride) membrane preparation via thermally induced phase separation. J. Membr. Sci. 2015, 493, 232. [Google Scholar] [CrossRef]
  23. Xu, H.-P.; Lang, W.-Z.; Yan, X.; Zhang, X.; Guo, Y.-J. Preparation and characterizations of Poly(vinylidene fluoride)/oxidized multi-wall carbon nanotube membranes with bi-continuous structure by thermally induced phase separation method. J. Membr. Sci. 2014, 467, 142–152. [Google Scholar] [CrossRef]
  24. Cui, A.; Liu, Z.; Xiao, C.; Zhang, Y. Effect of micro-sized SiO2-particle on the performance of PVDF blend membranes via TIPS. J. Membr. Sci. 2010, 360, 259–264. [Google Scholar] [CrossRef]
  25. Zhang, R.; Liu, Y.; He, M.; Su, Y.; Zhao, X.; Elimelech, M.; Jiang, Z. Antifouling membranes for sustainable water purification: Strategies and mechanisms. Chem. Soc. Rev. 2016, 45, 5888–5924. [Google Scholar] [CrossRef] [PubMed]
  26. He, Z.; Lan, X.; Hu, Q.; Li, H.; Li, L.; Mao, J. Antifouling strategies based on super-phobic polymer materials. Prog. Org. Coat. 2021, 157, 106285. [Google Scholar] [CrossRef]
  27. Huisman, I.H.; Prádanos, P.; Hernández, A. The effect of protein–protein and protein–membrane interactions on membrane fouling in ultrafiltration. J. Membr. Sci. 2000, 179, 79–90. [Google Scholar] [CrossRef]
  28. Sunil, K.; Sherugar, P.; Rao, S.; Lavanya, C.; Balakrishna, G.R.; Arthanareeswaran, G.; Padaki, M. Prolific approach for the removal of dyes by an effective interaction with polymer matrix using ultrafiltration membrane. J. Environ. Chem. Eng. 2021, 9, 106328. [Google Scholar] [CrossRef]
  29. Wei, Q.; Tian, Z.; Wang, H.; Qin, S.; Qin, Q.; Zhang, J.; Cui, Z. Fabrication of ultrafiltration membrane surface with synergistic anti-fouling effect of “dispersion-impedance” and anti-fouling mechanism of dye. J. Membr. Sci. 2022, 663, 121028. [Google Scholar] [CrossRef]
  30. Hairom, N.H.H.; Mohammad, A.W.; Kadhum, A.A.H. Nanofiltration of hazardous Congo red dye: Performance and flux decline analysis. J. Water Process Eng. 2014, 4, 99–106. [Google Scholar] [CrossRef] [Green Version]
  31. Kadhim, R.J.; Al-Ani, F.H.; Al-Shaeli, M.; Alsalhy, Q.F.; Figoli, A. Removal of dyes using graphene oxide (GO) mixed matrix membranes. Membranes 2020, 10, 366. [Google Scholar] [CrossRef]
  32. Huang, L.; Huang, S.; Venna, S.R.; Lin, H. Rightsizing nanochannels in reduced graphene oxide membranes by solvating for dye desalination. Environ. Sci. Technol. 2018, 52, 12649–12655. [Google Scholar] [CrossRef] [PubMed]
  33. Li, X.; Sotto, A.; Li, J.; Van der Bruggen, B. Progress and perspectives for synthesis of sustainable antifouling composite membranes containing in situ generated nanoparticles. J. Membr. Sci. 2017, 524, 502–528. [Google Scholar] [CrossRef]
  34. Nasir, A.; Masood, F.; Yasin, T.; Hameed, A. Progress in polymeric nanocomposite membranes for wastewater treatment: Preparation, properties and applications. J. Ind. Eng. Chem. 2019, 79, 29–40. [Google Scholar] [CrossRef]
  35. Lau, S.K.; Yong, W.F. Recent Progress of Zwitterionic Materials as Antifouling Membranes for Ultrafiltration, Nanofiltration, and Reverse Osmosis. ACS Appl. Polym. Mater. 2021, 3, 4390–4412. [Google Scholar] [CrossRef]
  36. Sonawane, S.; Thakur, P.; Sonawane, S.H.; Bhanvase, B.A. Nanomaterials for membrane synthesis: Introduction, mechanism, and challenges for wastewater treatment. In Handbook of Nanomaterials for Wastewater Treatment; Elsevier: Amsterdam, The Netherlands, 2021; pp. 537–553. [Google Scholar]
  37. Sile-Yuksel, M.; Tas, B.; Koseoglu-Imer, D.Y.; Koyuncu, I. Effect of silver nanoparticle (AgNP) location in nanocomposite membrane matrix fabricated with different polymer type on antibacterial mechanism. Desalination 2014, 347, 120–130. [Google Scholar] [CrossRef]
  38. Zhang, J.; Wang, Z.; Zhang, X.; Zheng, X.; Wu, Z. Enhanced antifouling behaviours of polyvinylidene fluoride membrane modified through blending with nano-TiO2/polyethylene glycol mixture. Appl. Surf. Sci. 2015, 345, 418–427. [Google Scholar] [CrossRef]
  39. Lowry, G.V.; Gregory, K.B.; Apte, S.C.; Lead, J.R. Transformations of nanomaterials in the environment. Environ. Sci. Technol. 2012, 46, 6893–6899. [Google Scholar] [CrossRef]
  40. Rezakazemi, M.; Amooghin, A.E.; Montazer-Rahmati, M.M.; Ismail, A.F.; Matsuura, T. State-of-the-art membrane based CO2 separation using mixed matrix membranes (MMMs): An overview on current status and future directions. Prog. Polym. Sci. 2014, 39, 817–861. [Google Scholar] [CrossRef]
  41. Venault, A.; Ballad, M.R.B.; Liu, Y.-H.; Aimar, P.; Chang, Y. Hemocompatibility of PVDF/PS-b-PEGMA membranes prepared by LIPS process. J. Membr. Sci. 2015, 477, 101–114. [Google Scholar] [CrossRef]
  42. Ma, Y.; Shi, F.; Ma, J.; Wu, M.; Zhang, J.; Gao, C. Effect of PEG additive on the morphology and performance of polysulfone ultrafiltration membranes. Desalination 2011, 272, 51–58. [Google Scholar] [CrossRef]
  43. Son, M.; Kim, H.; Jung, J.; Jo, S.; Choi, H. Influence of extreme concentrations of hydrophilic pore-former on reinforced polyethersulfone ultrafiltration membranes for reduction of humic acid fouling. Chemosphere 2017, 179, 194–201. [Google Scholar] [CrossRef] [PubMed]
  44. Khodadadi, F.; Mansourianfar, M.; Bozorg, A. Application of Dextran to Manipulate Formation Mechanism, Morphology, and Performance of Ultrafiltration Membranes. Chem. Eng. Res. Des. 2022, 183, 452–465. [Google Scholar] [CrossRef]
  45. Mulyati, S.; Aprilia, S.; Muchtar, S.; Syamsuddin, Y.; Rosnelly, C.M.; Bilad, M.R.; Samsuri, S.; Ismail, N.M. Fabrication of Polyvinylidene Difluoride Membrane with Enhanced Pore and Filtration Properties by Using Tannic Acid as an Additive. Polymers 2022, 14, 186. [Google Scholar] [CrossRef]
  46. Kong, L.; Zhang, D.; Shao, Z.; Han, B.; Lv, Y.; Gao, K.; Peng, X. Superior effect of TEMPO-oxidized cellulose nanofibrils (TOCNs) on the performance of cellulose triacetate (CTA) ultrafiltration membrane. Desalination 2014, 332, 117–125. [Google Scholar] [CrossRef]
  47. Huang, J.; Wang, H.; Zhang, K. Modification of PES membrane with Ag–SiO2: Reduction of biofouling and improvement of filtration performance. Desalination 2014, 336, 8–17. [Google Scholar] [CrossRef]
  48. Dadari, S.; Rahimi, M.; Zinadini, S. Removal of heavy metal from aqueous medium using novel high-performance, antifouling, and antibacterial nanofiltration polyethersulfone membrane modified with green synthesized Ni-doped Al2O3. Korean J. Chem. Eng. 2022, 39, 2424–2443. [Google Scholar] [CrossRef]
  49. Vani, B.; Shivakumar, M.; Kalyani, S.; Sridhar, S. TiO2 nanoparticles incorporated high-performance polyphenyl sulfone mixed matrix membranes for ultrafiltration of domestic greywater. Iran. Polym. J. 2021, 30, 917–934. [Google Scholar] [CrossRef]
  50. Ganjali, M.R.; Badiei, A.; Mouradzadegun, A.; Vatanpour, V.; Khadem, S.S.M.; Munir, M.T.; Habibzadeh, S.; Saeb, M.R.; Koyuncu, I. Erbium (III) molybdate as a new nanofiller for fabrication of antifouling polyethersulfone membranes. Mater. Today Commun. 2020, 25, 101379. [Google Scholar] [CrossRef]
  51. Hosseini, S.; Afshari, M.; Fazlali, A.; Farahani, S.K.; Bandehali, S.; Van der Bruggen, B.; Bagheripour, E. Mixed matrix PES-based nanofiltration membrane decorated by (Fe3O4–polyvinylpyrrolidone) composite nanoparticles with intensified antifouling and separation characteristics. Chem. Eng. Res. Des. 2019, 147, 390–398. [Google Scholar] [CrossRef]
  52. Wang, Y.; Hu, T.-T.; Han, X.-L.; Wang, Y.-Q.; Li, J.-D. Fabrication of Cu(OH)2 nanowires blended Poly(vinylidene fluoride) ultrafiltration membranes for oil-water separation. Chin. J. Polym. Sci. 2018, 36, 612–619. [Google Scholar] [CrossRef]
  53. Hosseini, S.M.; Bagheripour, E.; Ansari, M. Adapting the performance and physico-chemical properties of PES nanofiltration membrane by using of magnesium oxide nanoparticles. Korean J. Chem. Eng. 2017, 34, 1774–1780. [Google Scholar] [CrossRef]
  54. Younas, H.; Shao, J.; Bai, H.; Liu, L.; He, Y. Inherent porous structure modified by titanium dioxide nanoparticle incorporation and effect on the fouling behavior of hybrid Poly(vinylidene fluoride) membranes. J. Appl. Polym. Sci. 2016, 133. [Google Scholar] [CrossRef]
  55. Mehrnia, M.R.; Mojtahedi, Y.M.; Homayoonfal, M. What is the concentration threshold of nanoparticles within the membrane structure? A case study of Al2O3/PSf nanocomposite membrane. Desalination 2015, 372, 75–88. [Google Scholar] [CrossRef]
  56. Rabiee, H.; Farahani, M.H.D.A.; Vatanpour, V. Preparation and characterization of emulsion poly (vinyl chloride)(EPVC)/TiO2 nanocomposite ultrafiltration membrane. J. Membr. Sci. 2014, 472, 185–193. [Google Scholar] [CrossRef]
  57. Zhang, Y.; Yuan, M.; Lin, R.; Yue, X. Preparation and properties of Fe3+/PVDF-PMMA catalytic membrane. Desalin. Water Treat. 2013, 51, 3909–3913. [Google Scholar] [CrossRef]
  58. Vatanpour, V.; Madaeni, S.S.; Khataee, A.R.; Salehi, E.; Zinadini, S.; Monfared, H.A. TiO2 embedded mixed matrix PES nanocomposite membranes: Influence of different sizes and types of nanoparticles on antifouling and performance. Desalination 2012, 292, 19–29. [Google Scholar] [CrossRef]
  59. Keskin, B.; Mehrabani, S.A.N.; Arefi-Oskoui, S.; Vatanpour, V.; Teber, O.O.; Khataee, A.; Orooji, Y.; Koyuncu, I. Development of Ti2AlN MAX phase/cellulose acetate nanocomposite membrane for removal of dye, protein and lead ions. Carbohydr. Polym. 2022, 296, 119913. [Google Scholar] [CrossRef]
  60. Dehghankar, M.; Mohammadi, T.; Tavakolmoghadam, M.; Tofighy, M.A. Polyvinylidene Fluoride/Nanoclays (Cloisite 30B and Palygorskite) Mixed Matrix Membranes with Improved Performance and Antifouling Properties. Ind. Eng. Chem. Res. 2021, 60, 12078–12091. [Google Scholar] [CrossRef]
  61. Alghamdi, M.M.; El-Zahhar, A.A. Novel cellulose acetate propionate-halloysite composite membranes with improved permeation flux, salt rejection, and antifouling properties. Polym. Adv. Technol. 2020, 31, 2526–2534. [Google Scholar] [CrossRef]
  62. Wei, D.; Zhou, S.; Li, M.; Xue, A.; Zhang, Y.; Zhao, Y.; Zhong, J.; Yang, D. PVDF/palygorskite composite ultrafiltration membranes: Effects of nano-clay particles on membrane structure and properties. Appl. Clay Sci. 2019, 181, 105171. [Google Scholar] [CrossRef]
  63. Ahmad, T.; Guria, C.; Mandal, A. Synthesis, characterization and performance studies of mixed-matrix poly (vinyl chloride)-bentonite ultrafiltration membrane for the treatment of saline oily wastewater. Process Saf. Environ. Prot. 2018, 116, 703–717. [Google Scholar] [CrossRef]
  64. Arefi-Oskoui, S.; Khataee, A.; Vatanpour, V. Effect of solvent type on the physicochemical properties and performance of NLDH/PVDF nanocomposite ultrafiltration membranes. Sep. Purif. Technol. 2017, 184, 97–118. [Google Scholar] [CrossRef]
  65. Zhao, Y.; Li, N.; Yuan, F.; Zhang, H.; Xia, S. Preparation and characterization of hydrophilic and antifouling poly (ether sulfone) ultrafiltration membranes modified with Zn–Al layered double hydroxides. J. Appl. Polym. Sci. 2016, 133. [Google Scholar] [CrossRef]
  66. Gohari, R.J.; Korminouri, F.; Lau, W.; Ismail, A.; Matsuura, T.; Chowdhury, M.; Halakoo, E.; Gohari, M.J. A novel super-hydrophilic PSf/HAO nanocomposite ultrafiltration membrane for efficient separation of oil/water emulsion. Sep. Purif. Technol. 2015, 150, 13–20. [Google Scholar] [CrossRef]
  67. Asadi, A.; Gholami, F.; Nazari, S.; Dolatshah, M. Preparation of antifouling and antibacterial polyvinylidene fluoride membrane by incorporating functionalized multiwalled carbon nanotubes. J. Water Process Eng. 2022, 49, 103042. [Google Scholar] [CrossRef]
  68. Cheng, L.; Li, L.; Pei, X.; Ma, Y.; Liu, F.; Li, J. PVDF/MOFs mixed matrix ultrafiltration membrane for efficient water treatment. Front. Chem. 2022, 10, 985750. [Google Scholar] [CrossRef]
  69. Nemade, P.R.; Ganjare, A.V.; Ramesh, K.; Rakte, D.M.; Vaishnavi, P.; Thapa, G. Low fouling sulphonated carbon soot-polysulphone membranes for rapid dehydration of stabilized oil-water emulsions. J. Water Process Eng. 2020, 38, 101590. [Google Scholar] [CrossRef]
  70. Anjum, T.; Tamime, R.; Khan, A.L. Mixed-matrix membranes comprising of polysulfone and porous UiO-66, Zeolite 4A, and their combination: Preparation, removal of humic acid, and antifouling properties. Membranes 2020, 10, 393. [Google Scholar] [CrossRef]
  71. Shen, Z.; Chen, W.; Xu, H.; Yang, W.; Kong, Q.; Wang, A.; Ding, M.; Shang, J. Fabrication of a novel antifouling polysulfone membrane with in situ embedment of mxene nanosheets. Int. J. Environ. Res. Public Health 2019, 16, 4659. [Google Scholar] [CrossRef] [Green Version]
  72. Javadi, M.; Jafarzadeh, Y.; Yegani, R.; Kazemi, S. PVDF membranes embedded with PVP functionalized nanodiamond for pharmaceutical wastewater treatment. Chem. Eng. Res. Des. 2018, 140, 241–250. [Google Scholar] [CrossRef]
  73. Vatanpour, V.; Yekavalangi, M.E.; Safarpour, M. Preparation and characterization of nanocomposite PVDF ultrafiltration membrane embedded with nanoporous SAPO-34 to improve permeability and antifouling performance. Sep. Purif. Technol. 2016, 163, 300–309. [Google Scholar] [CrossRef] [Green Version]
  74. Ma, W.; Pan, J.; Ren, W.; Chen, L.; Huang, L.; Xu, S.; Jiang, Z. Fabrication of antibacterial and self-cleaning CuxP@g-C3N4/PVDF-CTFE mixed matrix membranes with enhanced properties for efficient ultrafiltration. J. Membr. Sci. 2022, 659, 120792. [Google Scholar] [CrossRef]
  75. Vatanpour, V.; Keskin, B.; Mehrabani, S.A.N.; Karimi, H.; Arabi, N.; Behroozi, A.H.; Shokrollahi-far, A.; Gul, B.Y.; Koyuncu, I. Investigation of boron nitride/silver/graphene oxide nanocomposite on separation and antibacterial improvement of polyethersulfone membranes in wastewater treatment. J. Environ. Chem. Eng. 2022, 10, 107035. [Google Scholar] [CrossRef]
  76. Bandehali, S.; Moghadassi, A.; Parvizian, F.; Zhang, Y.; Hosseini, S.M.; Shen, J. New mixed matrix PEI nanofiltration membrane decorated by glycidyl-POSS functionalized graphene oxide nanoplates with enhanced separation and antifouling behaviour: Heavy metal ions removal. Sep. Purif. Technol. 2020, 242, 116745. [Google Scholar] [CrossRef]
  77. Ghalamchi, L.; Aber, S.; Vatanpour, V.; Kian, M. Development of an antibacterial and visible photocatalytic nanocomposite microfiltration membrane incorporated by Ag3PO4/CuZnAl NLDH. Sep. Purif. Technol. 2019, 226, 218–231. [Google Scholar] [CrossRef]
  78. Wang, H.; Wang, Z.-M.; Yan, X.; Chen, J.; Lang, W.-Z.; Guo, Y.-J. Novel organic-inorganic hybrid polyvinylidene fluoride ultrafiltration membranes with antifouling and antibacterial properties by embedding N-halamine functionalized silica nanospheres. J. Ind. Eng. Chem. 2017, 52, 295–304. [Google Scholar] [CrossRef]
  79. Martin, A.; Arsuaga, J.M.; Roldan, N.; Martinez, A.; Sotto, A. Effect of amine functionalization of SBA-15 used as filler on the morphology and permeation properties of polyethersulfone-doped ultrafiltration membranes. J. Membr. Sci. 2016, 520, 8–18. [Google Scholar] [CrossRef]
  80. Nguyen, T.; Roddick, F.A.; Fan, L. Biofouling of water treatment membranes: A review of the underlying causes, monitoring techniques and control measures. Membranes 2012, 2, 804–840. [Google Scholar] [CrossRef] [Green Version]
  81. Vatanpour, V.; Jouyandeh, M.; Khadem, S.S.M.; Paziresh, S.; Dehqan, A.; Ganjali, M.R.; Moradi, H.; Mirsadeghi, S.; Badiei, A.; Munir, M.T. Highly antifouling polymer-nanoparticle-nanoparticle/polymer hybrid membranes. Sci. Total Environ. 2022, 810, 152228. [Google Scholar] [CrossRef]
  82. Rajabi, H.; Ghaemi, N.; Madaeni, S.S.; Daraei, P.; Astinchap, B.; Zinadini, S.; Razavizadeh, S.H. Nano-ZnO embedded mixed matrix polyethersulfone (PES) membrane: Influence of nanofiller shape on characterization and fouling resistance. Appl. Surf. Sci. 2015, 349, 66–77. [Google Scholar] [CrossRef]
  83. Ahmad, T.; Guria, C. Progress in the modification of polyvinyl chloride (PVC) membranes: A performance review for wastewater treatment. J. Water Process Eng. 2022, 45, 102466. [Google Scholar] [CrossRef]
  84. Alpatova, A.; Kim, E.-S.; Sun, X.; Hwang, G.; Liu, Y.; El-Din, M.G. Fabrication of porous polymeric nanocomposite membranes with enhanced anti-fouling properties: Effect of casting composition. J. Membr. Sci. 2013, 444, 449–460. [Google Scholar] [CrossRef]
  85. Sherazi, T.A.; Azam, S.; Shah, S.H.; Hussain, S.; Naqvi, S.A.R.; Li, S. Zwitterionic analog structured ultrafiltration membranes for high permeate flux and improved anti-fouling performance. J. Membr. Sci. 2022, 643, 120060. [Google Scholar] [CrossRef]
  86. Carretier, S.; Chen, L.-A.; Venault, A.; Yang, Z.-R.; Aimar, P.; Chang, Y. Design of PVDF/PEGMA-b-PS-b-PEGMA membranes by VIPS for improved biofouling mitigation. J. Membr. Sci. 2016, 510, 355–369. [Google Scholar] [CrossRef] [Green Version]
  87. Chen, Y.; Wei, M.; Wang, Y. Upgrading polysulfone ultrafiltration membranes by blending with amphiphilic block copolymers: Beyond surface segregation. J. Membr. Sci. 2016, 505, 53–60. [Google Scholar] [CrossRef]
  88. Gao, F.; Zhang, G.; Zhang, Q.; Zhan, X.; Chen, F. Improved antifouling properties of poly (ether sulfone) membrane by incorporating the amphiphilic comb copolymer with mixed poly (ethylene glycol) and poly (dimethylsiloxane) brushes. Ind. Eng. Chem. Res. 2015, 54, 8789–8800. [Google Scholar] [CrossRef]
  89. Gao, H.; Sun, X.; Gao, C. Antifouling polysulfone ultrafiltration membranes with sulfobetaine polyimides as novel additive for the enhancement of both water flux and protein rejection. J. Membr. Sci. 2017, 542, 81–90. [Google Scholar] [CrossRef]
  90. Maggay, I.V.; Yeh, T.-H.; Venault, A.; Hsu, C.-H.; Dizon, G.V.; Chang, Y. Tuning the molecular design of random copolymers for enhancing the biofouling mitigation of membrane materials. J. Membr. Sci. 2019, 588, 117217. [Google Scholar] [CrossRef]
  91. Wu, H.; Li, T.; Liu, B.; Chen, C.; Wang, S.; Crittenden, J.C. Blended PVC/PVC-g-PEGMA ultrafiltration membranes with enhanced performance and antifouling properties. Appl. Surf. Sci. 2018, 455, 987–996. [Google Scholar] [CrossRef]
  92. Wu, Q.; Tiraferri, A.; Li, T.; Xie, W.; Chang, H.; Bai, Y.; Liu, B. Superwettable PVDF/PVDF-g-PEGMA Ultrafiltration Membranes. ACS Omega 2020, 5, 23450–23459. [Google Scholar] [CrossRef]
  93. Wu, Q.; Tiraferri, A.; Wu, H.; Xie, W.; Liu, B. Improving the Performance of PVDF/PVDF-g-PEGMA Ultrafiltration Membranes by Partial Solvent Substitution with Green Solvent Dimethyl Sulfoxide during Fabrication. ACS Omega 2019, 4, 19799–19807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kanagaraj, P.; Nagendran, A.; Rana, D.; Matsuura, T.; Neelakandan, S.; Malarvizhi, K. Effects of polyvinylpyrrolidone on the permeation and fouling-resistance properties of polyetherimide ultrafiltration membranes. Ind. Eng. Chem. Res. 2015, 54, 4832–4838. [Google Scholar] [CrossRef]
  95. Zhu, M.; Li, D.; Sun, X.; Gao, C. Antifouling polysulfone membranes with an amphiphilic triblock additive. Mater. Chem. Phys. 2022, 285, 126108. [Google Scholar] [CrossRef]
  96. Vatsha, B.; Ngila, J.C.; Moutloali, R.M. Preparation of antifouling polyvinylpyrrolidone (PVP 40K) modified polyethersulfone (PES) ultrafiltration (UF) membrane for water purification. Phys. Chem. Earth Parts A/B/C 2014, 67, 125–131. [Google Scholar] [CrossRef]
  97. Xu, C.; Huang, W.; Lu, X.; Yan, D.; Chen, S.; Huang, H. Preparation of PVDF porous membranes by using PVDF-g-PVP powder as an additive and their antifouling property. Radiat. Phys. Chem. 2012, 81, 1763–1769. [Google Scholar] [CrossRef]
  98. Lv, J.; Zhang, G.; Zhang, H.; Zhao, C.; Yang, F. Improvement of antifouling performances for modified PVDF ultrafiltration membrane with hydrophilic cellulose nanocrystal. Appl. Surf. Sci. 2018, 440, 1091–1100. [Google Scholar] [CrossRef]
  99. Zaman, M.; Liu, H.; Xiao, H.; Chibante, F.; Ni, Y. Hydrophilic modification of polyester fabric by applying nanocrystalline cellulose containing surface finish. Carbohydr. Polym. 2013, 91, 560–567. [Google Scholar] [CrossRef]
  100. Zhang, D.; Karkooti, A.; Liu, L.; Sadrzadeh, M.; Thundat, T.; Liu, Y.; Narain, R. Fabrication of antifouling and antibacterial polyethersulfone (PES)/cellulose nanocrystals (CNC) nanocomposite membranes. J. Membr. Sci. 2018, 549, 350–356. [Google Scholar] [CrossRef]
  101. Zhou, J.; Chen, J.; He, M.; Yao, J. Cellulose acetate ultrafiltration membranes reinforced by cellulose nanocrystals: Preparation and characterization. J. Appl. Polym. Sci. 2016, 133. [Google Scholar] [CrossRef]
  102. Bai, H.; Wang, X.; Zhou, Y.; Zhang, L. Preparation and characterization of Poly(vinylidene fluoride) composite membranes blended with nano-crystalline cellulose. Prog. Nat. Sci. Mater. Int. 2012, 22, 250–257. [Google Scholar] [CrossRef]
  103. Lv, J.; Zhang, G.; Zhang, H.; Yang, F. Exploration of permeability and antifouling performance on modified cellulose acetate ultrafiltration membrane with cellulose nanocrystals. Carbohydr. Polym. 2017, 174, 190–199. [Google Scholar] [CrossRef] [PubMed]
  104. Mansor, E.S.; Abdallah, H.; Shaban, A.M. Fabrication of high selectivity blend membranes based on poly vinyl alcohol for crystal violet dye removal. J. Environ. Chem. Eng. 2020, 8, 103706. [Google Scholar] [CrossRef]
  105. Razmgar, K.; Nasiraee, M. Polyvinyl alcohol-based membranes for filtration of aqueous solutions: A comprehensive review. Polym. Eng. Sci. 2022, 62, 25–43. [Google Scholar] [CrossRef]
  106. Yuan, H.; Wang, Y.; Cheng, L.; Liu, W.; Ren, J.; Meng, L. Improved antifouling property of poly (ether sulfone) ultrafiltration membrane through blending with poly (vinyl alcohol). Ind. Eng. Chem. Res. 2014, 53, 18549–18557. [Google Scholar] [CrossRef]
  107. Zhang, J.; Wang, Z.W.; Wang, Q.Y.; Ma, J.X.; Cao, J.; Hu, W.J.; Wu, Z.C. Relationship between polymers compatibility and casting solution stability in fabricating PVDF/PVA membranes. J. Membr. Sci. 2017, 537, 263–271. [Google Scholar] [CrossRef]
  108. Ounifi, I.; Guesmi, Y.; Ursino, C.; Santoro, S.; Mahfoudhi, S.; Figoli, A.; Ferjanie, E.; Hafiane, A. Antifouling membranes based on cellulose acetate (CA) blended with poly (acrylic acid) for heavy metal remediation. Appl. Sci. 2021, 11, 4354. [Google Scholar] [CrossRef]
  109. Yu, H.; Cao, Y.; Kang, G.; Liu, J.; Li, M. Tethering methoxy polyethylene glycols to improve the antifouling property of PSF/PAA-blended membranes. J. Appl. Polym. Sci. 2012, 124, E123–E133. [Google Scholar] [CrossRef]
  110. Ismail, R.A.; Kumar, M.; Khanzada, N.K.; Thomas, N.; Sreedhar, N.; An, A.K.; Arafat, H.A. Hybrid NF and UF membranes tailored using quaternized polydopamine for enhanced removal of salts and organic pollutants from water. Desalination 2022, 539, 115954. [Google Scholar] [CrossRef]
  111. Mulyati, S.; Muchtar, S.; Arahman, N.; Meirisa, F.; Syamsuddin, Y.; Zuhra, Z.; Rosnelly, C.M.; Shamsuddin, N.; Mat Nawi, N.I.; Wirzal, M.D.H. One-Pot polymerization of dopamine as an additive to enhance permeability and antifouling properties of polyethersulfone membrane. Polymers 2020, 12, 1807. [Google Scholar] [CrossRef]
  112. Ang, M.B.M.Y.; Macni, C.R.M.; Caparanga, A.R.; Huang, S.-H.; Tsai, H.-A.; Lee, K.-R.; Lai, J.-Y. Mitigating the fouling of mixed-matrix cellulose acetate membranes for oil–water separation through modification with polydopamine particles. Chem. Eng. Res. Des. 2020, 159, 195–204. [Google Scholar] [CrossRef]
  113. Jiang, J.-H.; Zhu, L.-P.; Zhang, H.-T.; Zhu, B.-K.; Xu, Y.-Y. Improved hydrodynamic permeability and antifouling properties of Poly(vinylidene fluoride) membranes using polydopamine nanoparticles as additives. J. Membr. Sci. 2014, 457, 73–81. [Google Scholar] [CrossRef]
  114. Kallem, P.; Ibrahim, Y.; Hasan, S.W.; Show, P.L.; Banat, F. Fabrication of novel polyethersulfone (PES) hybrid ultrafiltration membranes with superior permeability and antifouling properties using environmentally friendly sulfonated functionalized polydopamine nanofillers. Sep. Purif. Technol. 2021, 261, 118311. [Google Scholar] [CrossRef]
  115. Bui, V.T.; Abdelrasoul, A.; McMartin, D.W. Influence of zwitterionic structure design on mixed matrix membrane stability, hydrophilicity, and fouling resistance: A computational study. J. Mol. Graph. Modell. 2022, 114, 108187. [Google Scholar] [CrossRef] [PubMed]
  116. Dizon, G.V.; Lee, Y.-S.; Venault, A.; Maggay, I.V.; Chang, Y. Zwitterionic PMMA-r-PEGMA-r-PSBMA copolymers for the formation of anti-biofouling bicontinuous membranes by the VIPS process. J. Membr. Sci. 2021, 618, 118753. [Google Scholar] [CrossRef]
  117. Dizon, G.V.; Venault, A. Direct in-situ modification of PVDF membranes with a zwitterionic copolymer to form bi-continuous and fouling resistant membranes. J. Membr. Sci. 2018, 550, 45–58. [Google Scholar] [CrossRef]
  118. Li, J.-H.; Li, M.-Z.; Miao, J.; Wang, J.-B.; Shao, X.-S.; Zhang, Q.-Q. Improved surface property of PVDF membrane with amphiphilic zwitterionic copolymer as membrane additive. Appl. Surf. Sci. 2012, 258, 6398–6405. [Google Scholar] [CrossRef]
  119. Maggay, I.V.; Suba, M.C.A.M.; Aini, H.N.; Wu, C.-J.; Tang, S.-H.; Aquino, R.B.; Chang, Y.; Venault, A. Thermostable antifouling zwitterionic vapor-induced phase separation membranes. J. Membr. Sci. 2021, 627, 119227. [Google Scholar] [CrossRef]
  120. Maggay, I.V.B.; Aini, H.N.; Lagman, M.M.G.; Tang, S.-H.; Aquino, R.R.; Chang, Y.; Venault, A. A Biofouling Resistant Zwitterionic Polysulfone Membrane Prepared by a Dual-Bath Procedure. Membranes 2022, 12, 69. [Google Scholar] [CrossRef]
  121. Nagandran, S.; Goh, P.S.; Ismail, A.F.; Wong, T.-W.; Binti Wan Dagang, W.R.Z. The recent progress in modification of polymeric membranes using organic macromolecules for water treatment. Symmetry 2020, 12, 239. [Google Scholar] [CrossRef] [Green Version]
  122. Venault, A.; Hsu, C.-H.; Ishihara, K.; Chang, Y. Zwitterionic bi-continuous membranes from a phosphobetaine copolymer/Poly(vinylidene fluoride) blend via VIPS for biofouling mitigation. J. Membr. Sci. 2018, 550, 377–388. [Google Scholar] [CrossRef]
  123. Venault, A.; Zhou, R.-J.; Galeta, T.A.; Chang, Y. Engineering sterilization-resistant and fouling-resistant porous membranes by the vapor-induced phase separation process using a sulfobetaine methacrylamide amphiphilic derivative. J. Membr. Sci. 2022, 658, 120760. [Google Scholar] [CrossRef]
  124. Zhao, X.; He, C. Efficient preparation of super antifouling PVDF ultrafiltration membrane with one step fabricated zwitterionic surface. ACS Appl. Mater. Interfaces 2015, 7, 17947–17953. [Google Scholar] [CrossRef] [PubMed]
  125. Almaie, S.; Vatanpour, V.; Rasoulifard, M.H.; Dorraji, M.S.S. Novel negatively-charged amphiphilic copolymers of PVDF-g-PAMPS and PVDF-g-PAA to improve permeability and fouling resistance of PVDF UF membrane. React. Funct. Polym. 2022, 179, 105386. [Google Scholar] [CrossRef]
  126. Wang, X.; Zeng, B.; Chen, T.; Liu, X.; Wu, T.; Shen, H.; Luo, W.; Yuan, C.; Xu, Y.; Chen, G. Polyethersulfone microfiltration membrane modified by an amphiphilic dithiolane-containing copolymer for improving anti-protein-fouling performance and rejection of nanoparticles. Polym. Adv. Technol. 2020, 31, 2816–2826. [Google Scholar] [CrossRef]
  127. Zhao, J.; Wang, Q.; Yang, J.; Li, Y.; Liu, Z.; Zhang, L.; Zhao, Y.; Zhang, S.; Chen, L. Comb-shaped amphiphilic triblock copolymers blend PVDF membranes overcome the permeability-selectivity trade-off for protein separation. Sep. Purif. Technol. 2020, 239, 116596. [Google Scholar] [CrossRef]
  128. Roy, S.; Bhalani, D.V.; Jewrajka, S.K. Surface segregation of segmented amphiphilic copolymer of poly (dimethylsiloxane) and poly (ethylene glycol) on Poly(vinylidene fluoride) blend membrane for oil–water emulsion separation. Sep. Purif. Technol. 2020, 232, 115940. [Google Scholar] [CrossRef]
  129. Han, N.; Zhang, W.; Wang, W.; Yang, C.; Tan, L.; Cui, Z.; Li, W.; Zhang, X. Amphiphilic cellulose for enhancing the antifouling and separation performances of poly (acrylonitrile-co-methyl acrylate) ultrafiltration membrane. J. Membr. Sci. 2019, 591, 117276. [Google Scholar] [CrossRef]
  130. Liu, L.; Huang, L.; Shi, M.; Li, W.; Xing, W. Amphiphilic PVDF-g-PDMAPMA ultrafiltration membrane with enhanced hydrophilicity and antifouling properties. J. Appl. Polym. Sci. 2019, 136, 48049. [Google Scholar] [CrossRef]
  131. Yadav, S.; Ibrar, I.; Altaee, A.; Déon, S.; Zhou, J. Preparation of novel high permeability and antifouling polysulfone-vanillin membrane. Desalination 2020, 496, 114759. [Google Scholar] [CrossRef]
  132. Arthanareeswaran, G.; Ismail, A. Enhancement of permeability and antibiofouling properties of polyethersulfone (PES) membrane through incorporation of quorum sensing inhibition (QSI) compound. J. Taiwan Inst. Chem. Eng. 2017, 72, 200–212. [Google Scholar]
  133. Yadav, S.; Ibrar, I.; Samal, A.K.; Altaee, A.; Déon, S.; Zhou, J.; Ghaffour, N. Preparation of fouling resistant and highly perm-selective novel PSf/GO-vanillin nanofiltration membrane for efficient water purification. J. Hazard. Mater. 2022, 421, 126744. [Google Scholar] [CrossRef] [PubMed]
  134. Khoerunnisa, F.; Sihombing, M.; Nurhayati, M.; Dara, F.; Triadi, H.A.; Nasir, M.; Hendrawan, H.; Pratiwi, A.; Ng, E.-P.; Opaprakasit, P. Poly (ether sulfone)-based ultrafiltration membranes using chitosan/ammonium chloride to enhance permeability and antifouling properties. Polym. J. 2022, 54, 525–537. [Google Scholar] [CrossRef]
  135. Zareei, F.; Bandehali, S.; Hosseini, S.M. Enhancing the separation and antifouling properties of PES nanofiltration membrane by use of chitosan functionalized magnetic nanoparticles. Korean J. Chem. Eng. 2021, 38, 1014–1022. [Google Scholar] [CrossRef]
  136. Kumar, R.; Isloor, A.M.; Ismail, A.; Matsuura, T. Synthesis and characterization of novel water soluble derivative of chitosan as an additive for polysulfone ultrafiltration membrane. J. Membr. Sci. 2013, 440, 140–147. [Google Scholar] [CrossRef]
  137. Kumar, R.; Isloor, A.M.; Ismail, A.; Matsuura, T. Performance improvement of polysulfone ultrafiltration membrane using N-succinyl chitosan as additive. Desalination 2013, 318, 1–8. [Google Scholar] [CrossRef]
  138. Kumar, R.; Isloor, A.M.; Ismail, A.F.; Rashid, S.A.; Matsuura, T. Polysulfone–Chitosan blend ultrafiltration membranes: Preparation, characterization, permeation and antifouling properties. Rsc Adv. 2013, 3, 7855–7861. [Google Scholar] [CrossRef]
  139. Qin, A.; Li, X.; Zhao, X.; Liu, D.; He, C. Preparation and characterization of nano-chitin whisker reinforced PVDF membrane with excellent antifouling property. J. Membr. Sci. 2015, 480, 1–10. [Google Scholar] [CrossRef]
  140. Xie, M.; Huan, G.; Xia, W.; Feng, X.; Chen, L.; Zhao, Y. Preparation and performance optimization of PVDF anti-fouling membrane modified by chitin. J. Polym. Eng. 2018, 38, 179–186. [Google Scholar] [CrossRef]
  141. Colburn, A.; Vogler, R.J.; Patel, A.; Bezold, M.; Craven, J.; Liu, C.; Bhattacharyya, D. Composite membranes derived from cellulose and lignin sulfonate for selective separations and antifouling aspects. Nanomaterials 2019, 9, 867. [Google Scholar] [CrossRef] [Green Version]
  142. Yong, M.; Zhang, Y.; Sun, S.; Liu, W. Properties of polyvinyl chloride (PVC) ultrafiltration membrane improved by lignin: Hydrophilicity and antifouling. J. Membr. Sci. 2019, 575, 50–59. [Google Scholar] [CrossRef]
  143. Lavanya, C.; Balakrishna, R.G. Naturally derived polysaccharides-modified PSF membranes: A potency in enriching the antifouling nature of membranes. Sep. Purif. Technol. 2020, 230, 115887. [Google Scholar] [CrossRef]
  144. Lavanya, C.; Soontarapa, K.; Jyothi, M.; Balakrishna, R.G. Environmental friendly and cost effective caramel for congo red removal, high flux, and fouling resistance of polysulfone membranes. Sep. Purif. Technol. 2019, 211, 348–358. [Google Scholar] [CrossRef]
  145. Zhang, L.; Shan, C.; Jiang, X.; Li, X.; Yu, L. High hydrophilic antifouling membrane modified with capsaicin-mimic moieties via microwave assistance (MWA) for efficient water purification. Chem. Eng. J. 2018, 338, 688–699. [Google Scholar] [CrossRef]
  146. Xu, J.; Feng, X.; Hou, J.; Wang, X.; Shan, B.; Yu, L.; Gao, C. Preparation and characterization of a novel polysulfone UF membrane using a copolymer with capsaicin-mimic moieties for improved anti-fouling properties. J. Membr. Sci. 2013, 446, 171–180. [Google Scholar] [CrossRef]
  147. Manawi, Y.; Kochkodan, V.; Mahmoudi, E.; Johnson, D.J.; Mohammad, A.W.; Atieh, M.A. Characterization and separation performance of a novel polyethersulfone membrane blended with acacia gum. Sci. Rep. 2017, 7, 15831. [Google Scholar] [CrossRef] [Green Version]
  148. Liu, X.; Wu, H.; Wu, P. Synchronous Engineering for Biomimetic Murray Porous Membranes Using Isocyanate. Nano Lett. 2022, 22, 3077–3086. [Google Scholar] [CrossRef]
  149. Zhao, H.; Wu, L.; Zhou, Z.; Zhang, L.; Chen, H. Improving the antifouling property of polysulfone ultrafiltration membrane by incorporation of isocyanate-treated graphene oxide. Phys. Chem. Chem. Phys. 2013, 15, 9084–9092. [Google Scholar] [CrossRef]
  150. Hebbar, R.S.; Isloor, A.M.; Ismail, A.; Shilton, S.J.; Obaid, A.; Fun, H.-K. Probing the morphology and anti-organic fouling behaviour of a polyetherimide membrane modified with hydrophilic organic acids as additives. New J. Chem. 2015, 39, 6141–6150. [Google Scholar] [CrossRef] [Green Version]
  151. Saniei, N.; Ghasemi, N.; Zinatizadeh, A.; Zinadini, S.; Ramezani, M.; Derakhshan, A. Preparation and characterization of a novel antifouling nano filtration poly ethersulfone (PES) membrane by embedding goethite-tannic acid nanoparticles. Sep. Purif. Technol. 2020, 241, 116646. [Google Scholar] [CrossRef]
  152. Peng, G.; Yaoqin, W.; Changmei, S.; Chunnuan, J.; Ying, Z.; Rongjun, Q.; Ying, W. Preparation and properties of PVC-based ultrafiltration membrane reinforced by in-situ synthesized p-aramid nanoparticles. J. Membr. Sci. 2022, 642, 119993. [Google Scholar] [CrossRef]
  153. Venault, A.; Ballad, M.R.B.; Huang, Y.-T.; Liu, Y.-H.; Kao, C.-H.; Chang, Y. Antifouling PVDF membrane prepared by VIPS for microalgae harvesting. Chem. Eng. Sci. 2016, 142, 97–111. [Google Scholar] [CrossRef]
  154. Saini, B.; Sinha, M.K. Effect of hydrophilic poly (ethylene glycol) methyl ether additive on the structure, morphology, and performance of polysulfone flat sheet ultrafiltration membrane. J. Appl. Polym. Sci. 2019, 136, 47163. [Google Scholar] [CrossRef]
  155. Kosma, V.A.; Beltsios, K.G. Macrovoids in solution-cast membranes: Direct probing of systems exhibiting horizontal macrovoid growth. J. Membr. Sci. 2012, 407, 93–107. [Google Scholar] [CrossRef]
  156. Wei, Q.; Zhang, F.; Li, J.; Li, B.; Zhao, C. Oxidant-induced dopamine polymerization for multifunctional coatings. Polym. Chem. 2010, 1, 1430–1433. [Google Scholar] [CrossRef]
  157. Kumar, M.; Sreedhar, N.; Thomas, N.; Mavukkandy, M.; Ismail, R.A.; Aminabhavi, T.M.; Arafat, H.A. Polydopamine-coated graphene oxide nanosheets embedded in sulfonated poly (ether sulfone) hybrid UF membranes with superior antifouling properties for water treatment. Chem. Eng. J. 2022, 433, 133526. [Google Scholar] [CrossRef]
  158. Guo, X.; Fan, S.; Hu, Y.; Fu, X.; Shao, H.; Zhou, Q. A novel membrane biofouling mitigation strategy of D-amino acid supported by polydopamine and halloysite nanotube. J. Membr. Sci. 2019, 579, 131–140. [Google Scholar] [CrossRef]
  159. Wu, H.; Liu, Y.; Huang, J.; Mao, L.; Chen, J.; Li, M. Preparation and characterization of antifouling and antibacterial polysulfone ultrafiltration membranes incorporated with a silver–polydopamine nanohybrid. J. Appl. Polym. Sci. 2018, 135, 46430. [Google Scholar] [CrossRef]
  160. Zhang, L.; Shi, Y.; Wang, T.; Li, S.; Zheng, X.; Zhao, Z.; Feng, Y.; Zhao, Z. Fabrication of novel anti-fouling poly (m-phenylene isophthalamide) ultrafiltration membrane modified with Pluronic F127 via coupling phase inversion and surface segregation. Sep. Purif. Technol. 2022, 282, 120106. [Google Scholar] [CrossRef]
  161. Norde, W. Protein adsorption at solid surfaces: A thermodynamic approach. Pure Appl. Chem. 1994, 66, 491–496. [Google Scholar] [CrossRef] [Green Version]
  162. Atwood, J.L. Comprehensive Supramolecular Chemistry II; Elsevier: Amsterdam, The Netherlands, 2017. [Google Scholar]
  163. Liu, Y.; Su, Y.; Zhao, X.; Zhang, R.; Ma, T.; He, M.; Jiang, Z. Enhanced membrane antifouling and separation performance by manipulating phase separation and surface segregation behaviors through incorporating versatile modifier. J. Membr. Sci. 2016, 499, 406–417. [Google Scholar] [CrossRef]
  164. Wang, J.; Liu, Y.; Liu, T.; Xu, X.; Hu, Y. Improving the perm-selectivity and anti-fouling property of UF membrane through the micro-phase separation of PSf-b-PEG block copolymers. J. Membr. Sci. 2020, 599, 117851. [Google Scholar] [CrossRef]
  165. Lin, H.-T.; Venault, A.; Huang, H.Q.; Lee, K.-R.; Chang, Y. Introducing a PEGylated diblock copolymer into PVDF hollow-fibers for reducing their fouling propensity. J. Taiwan Inst. Chem. Eng. 2018, 87, 252–263. [Google Scholar] [CrossRef]
  166. Kardela, J.H.; Millichamp, I.S.; Ferguson, J.; Parry, A.L.; Reynolds, K.J.; Aldred, N.; Clare, A.S. Nonfreezable water and polymer swelling control the marine antifouling performance of polymers with limited hydrophilic content. ACS Appl. Mater. Interfaces 2019, 11, 29477–29489. [Google Scholar] [CrossRef] [PubMed]
  167. Morisaku, T.; Watanabe, J.; Konno, T.; Takai, M.; Ishihara, K. Hydration of phosphorylcholine groups attached to highly swollen polymer hydrogels studied by thermal analysis. Polymer 2008, 49, 4652–4657. [Google Scholar] [CrossRef]
  168. Zheng, L.; Sundaram, H.S.; Wei, Z.; Li, C.; Yuan, Z. Applications of zwitterionic polymers. React. Funct. Polym. 2017, 118, 51–61. [Google Scholar] [CrossRef]
  169. Laschewsky, A. Structures and synthesis of zwitterionic polymers. Polymers 2014, 6, 1544–1601. [Google Scholar] [CrossRef]
  170. Li, M.; Zhuang, B.; Yu, J. Functional Zwitterionic Polymers on Surface: Structures and Applications. Chem.-Asian J. 2020, 15, 2060–2075. [Google Scholar] [CrossRef]
  171. Zhang, X.; Ma, J.; Zheng, J.; Dai, R.; Wang, X.; Wang, Z. Recent advances in nature-inspired antifouling membranes for water purification. Chem. Eng. J. 2021, 432, 134425. [Google Scholar] [CrossRef]
  172. Al-Bulushia, M.; Al-Hinaic, M.; Al-Obaidanid, S.; Dobrestove, S.; Nxumalog, E.; Al-Abria, M. Antifouling enhancement of polyethersulfone membranes incorporated with silver nanoparticles for bovine serum albumin and humic acid removal. Desalin. Water Treat. 2021, 243, 123–142. [Google Scholar] [CrossRef]
  173. Azhar, F.H.; Harun, Z.; Yusof, K.N.; Alias, S.S.; Hashim, N.; Sazali, E.S. A study of different concentrations of bio-silver nanoparticles in polysulfone mixed matrix membranes in water separation performance. J. Water Process Eng. 2020, 38, 101575. [Google Scholar] [CrossRef]
  174. Rana, S.; Nazar, U.; Ali, J.; Ali, Q.u.A.; Ahmad, N.M.; Sarwar, F.; Waseem, H.; Jamil, S.U.U. Improved antifouling potential of polyether sulfone polymeric membrane containing silver nanoparticles: Self-cleaning membranes. Environ. Technol. 2018, 39, 1413–1421. [Google Scholar] [CrossRef] [PubMed]
  175. Ahmad Rehan, Z.; Gzara, L.; Bahadar Khan, S.; A Alamry, K.; El-Shahawi, M.; H Albeirutty, M.; Figoli, A.; Drioli, E.; M Asiri, A. Synthesis and characterization of silver nanoparticles-filled polyethersulfone membranes for antibacterial and anti-biofouling application. Recent Pat. Nanotechnol. 2016, 10, 231–251. [Google Scholar] [CrossRef] [PubMed]
  176. Zhang, M.; Field, R.W.; Zhang, K. Biogenic silver nanocomposite polyethersulfone UF membranes with antifouling properties. J. Membr. Sci. 2014, 471, 274–284. [Google Scholar] [CrossRef]
  177. Esfahani, M.R.; Koutahzadeh, N.; Esfahani, A.R.; Firouzjaei, M.D.; Anderson, B.; Peck, L. A novel gold nanocomposite membrane with enhanced permeation, rejection and self-cleaning ability. J. Membr. Sci. 2019, 573, 309–319. [Google Scholar] [CrossRef]
  178. Goswami, R.; Gogoi, M.; Borah, H.J.; Ingole, P.G.; Hazarika, S. Biogenic synthesized Pd-nanoparticle incorporated antifouling polymeric membrane for removal of crystal violet dye. J. Environ. Chem. Eng. 2018, 6, 6139–6146. [Google Scholar] [CrossRef]
  179. Pang, R.; Li, X.; Li, J.; Lu, Z.; Sun, X.; Wang, L. Preparation and characterization of ZrO2/PES hybrid ultrafiltration membrane with uniform ZrO2 nanoparticles. Desalination 2014, 332, 60–66. [Google Scholar] [CrossRef]
  180. Shen, X.; Xie, T.; Wang, J.; Liu, P.; Wang, F. An anti-fouling Poly(vinylidene fluoride) hybrid membrane blended with functionalized ZrO2 nanoparticles for efficient oil/water separation. RSC Adv. 2017, 7, 5262–5271. [Google Scholar] [CrossRef] [Green Version]
  181. Huang, X.; Tian, F.; Chen, G.; Wang, F.; Weng, R.; Xi, B. Preparation and Characterization of Regenerated Cellulose Membrane Blended with ZrO2 Nanoparticles. Membranes 2021, 12, 42. [Google Scholar] [CrossRef]
  182. Pang, R.-Z.; Li, X.; Li, J.-S.; Lu, Z.-Y.; Huang, C.; Sun, X.-Y.; Wang, L.-J. In situ preparation and antifouling performance of ZrO2/PVDF hybrid membrane. Acta Phys.-Chim. Sin. 2013, 29, 2592–2598. [Google Scholar]
  183. Shukla, A.K.; Alam, J.; Ali, F.A.A.; Alhoshan, M. Efficient soluble anionic dye removal and antimicrobial properties of ZnO embedded-Polyphenylsulfone membrane. Water Environ. J. 2021, 35, 670–684. [Google Scholar] [CrossRef]
  184. Ahmad, A.L.; Sugumaran, J.; Shoparwe, N.F. Antifouling properties of PES membranes by blending with ZnO nanoparticles and NMP–acetone mixture as solvent. Membranes 2018, 8, 131. [Google Scholar] [CrossRef] [Green Version]
  185. Sarihan, A.; Eren, E. Novel high performanced and fouling resistant PSf/ZnO membranes for water treatment. Membr. Water Treat 2017, 8, 563–574. [Google Scholar]
  186. Nasrollahi, N.; Aber, S.; Vatanpour, V.; Mahmoodi, N.M. The effect of amine functionalization of CuO and ZnO nanoparticles used as additives on the morphology and the permeation properties of polyethersulfone ultrafiltration nanocomposite membranes. Compos. Part B Eng. 2018, 154, 388–409. [Google Scholar] [CrossRef]
  187. Nasrollahi, N.; Vatanpour, V.; Aber, S. Improving the permeability and antifouling property of PES ultrafiltration membranes using the drying method and incorporating the CuO-ZnO nanocomposite. J. Water Process Eng. 2019, 31, 100891. [Google Scholar] [CrossRef]
  188. Hosseini, S.M.; Karami, F.; Farahani, S.K.; Bandehali, S.; Shen, J.; Bagheripour, E.; Seidypoor, A. Tailoring the separation performance and antifouling property of polyethersulfone based NF membrane by incorporating hydrophilic CuO nanoparticles. Korean J. Chem. Eng. 2020, 37, 866–874. [Google Scholar] [CrossRef]
  189. Dzinun, H.; Othman, M.H.D.; Ismail, A.F.; Puteh, M.H.; Rahman, M.A.; Jaafar, J.; Adrus, N.; Hashim, N.A. Antifouling behavior and separation performance of immobilized TiO2 in dual layer hollow fiber membranes. Polym. Eng. Sci. 2018, 58, 1636–1643. [Google Scholar] [CrossRef]
  190. Arif, Z.; Sethy, N.K.; Kumari, L.; Mishra, P.K.; Verma, B. Antifouling behaviour of PVDF/TiO2 composite membrane: A quantitative and qualitative assessment. Iran. Polym. J. 2019, 28, 301–312. [Google Scholar] [CrossRef]
  191. Alsohaimi, I.H.; Kumar, M.; Algamdi, M.S.; Khan, M.A.; Nolan, K.; Lawler, J. Antifouling hybrid ultrafiltration membranes with high selectivity fabricated from polysulfone and sulfonic acid functionalized TiO2 nanotubes. Chem. Eng. J. 2017, 316, 573–583. [Google Scholar] [CrossRef]
  192. Ahmad, A.L.; Che Lah, N.F.; Norzli, N.A.; Pang, W.Y. A Contrastive Study of Self-Assembly and Physical Blending Mechanism of TiO2 Blended Polyethersulfone Membranes for Enhanced Humic Acid Removal and Alleviation of Membrane Fouling. Membranes 2022, 12, 162. [Google Scholar] [CrossRef]
  193. Hosseini, S.S.; Fakharian Torbati, S.; Alaei Shahmirzadi, M.A.; Tavangar, T. Fabrication, characterization, and performance evaluation of polyethersulfone/TiO2 nanocomposite ultrafiltration membranes for produced water treatment. Polym. Adv. Technol. 2018, 29, 2619–2631. [Google Scholar] [CrossRef]
  194. Teli, S.B.; Molina, S.; Sotto, A.; Calvo, E.G.A.; Abajob, J.D. Fouling resistant polysulfone–PANI/TiO2 ultrafiltration nanocomposite membranes. Ind. Eng. Chem. Res. 2013, 52, 9470–9479. [Google Scholar] [CrossRef]
  195. Enayatzadeh, M.; Mohammadi, T.; Fallah, N. Influence of TiO2 nanoparticles loading on permeability and antifouling properties of nanocomposite polymeric membranes: Experimental and statistical analysis. J. Polym. Res. 2019, 26, 240. [Google Scholar] [CrossRef]
  196. Vatanpour, V.; Hazrati, M.; Sheydaei, M.; Dehqan, A. Investigation of using UV/H2O2 pre-treatment process on filterability and fouling reduction of PVDF/TiO2 nanocomposite ultrafiltration membrane. Chem. Eng. Process.-Process Intensif. 2022, 170, 108677. [Google Scholar] [CrossRef]
  197. Sun, T.; Liu, Y.; Shen, L.; Xu, Y.; Li, R.; Huang, L.; Lin, H. Magnetic field assisted arrangement of photocatalytic TiO2 particles on membrane surface to enhance membrane antifouling performance for water treatment. J. Colloid Interface Sci. 2020, 570, 273–285. [Google Scholar] [CrossRef]
  198. Nair, A.K.; Shalin, P.; JagadeeshBabu, P. Performance enhancement of polysulfone ultrafiltration membrane using TiO2 nanofibers. Desalin. Water Treat. 2016, 57, 10506–10514. [Google Scholar] [CrossRef]
  199. Bidsorkhi, H.C.; Riazi, H.; Emadzadeh, D.; Ghanbari, M.; Matsuura, T.; Lau, W.; Ismail, A. Preparation and characterization of a novel highly hydrophilic and antifouling polysulfone/nanoporous TiO2 nanocomposite membrane. Nanotechnology 2016, 27, 415706. [Google Scholar] [CrossRef]
  200. Zhang, H.; Guo, Y.; Zhang, X.; Hu, X.; Wang, C.; Yang, Y. Preparation and characterization of PSF-TiO2 hybrid hollow fiber UF membrane by sol–gel method. J. Polym. Res. 2020, 27, 376. [Google Scholar] [CrossRef]
  201. Behboudi, A.; Jafarzadeh, Y.; Yegani, R. Preparation and characterization of TiO2 embedded PVC ultrafiltration membranes. Chem. Eng. Res. Des. 2016, 114, 96–107. [Google Scholar] [CrossRef]
  202. Li, X.; Fang, X.; Pang, R.; Li, J.; Sun, X.; Shen, J.; Han, W.; Wang, L. Self-assembly of TiO2 nanoparticles around the pores of PES ultrafiltration membrane for mitigating organic fouling. J. Membr. Sci. 2014, 467, 226–235. [Google Scholar] [CrossRef]
  203. Li, X.; Li, J.; Fang, X.; Bakzhan, K.; Wang, L.; Van der Bruggen, B. A synergetic analysis method for antifouling behavior investigation on PES ultrafiltration membrane with self-assembled TiO2 nanoparticles. J. Colloid Interface Sci. 2016, 469, 164–176. [Google Scholar] [CrossRef]
  204. Akbari, A.; Yegani, R.; Pourabbas, B.; Behboudi, A. Analysis of antifouling behavior of high dispersible hydrophilic poly (ethylene glycol)/vinyl functionalized SiO2 nanoparticles embedded polyethylene membrane. Desalin. Water Treat. 2017, 76, 83–97. [Google Scholar] [CrossRef] [Green Version]
  205. Kamari, S.; Shahbazi, A. Biocompatible Fe3O4@SiO2-NH2 nanocomposite as a green nanofiller embedded in PES–nanofiltration membrane matrix for salts, heavy metal ion and dye removal: Long–term operation and reusability tests. Chemosphere 2020, 243, 125282. [Google Scholar] [CrossRef] [PubMed]
  206. Yu, H.; Zhang, X.; Zhang, Y.; Liu, J.; Zhang, H. Development of a hydrophilic PES ultrafiltration membrane containing SiO2@N-Halamine nanoparticles with both organic antifouling and antibacterial properties. Desalination 2013, 326, 69–76. [Google Scholar] [CrossRef]
  207. Liu, Y.; Wei, R.; Lin, O.; Zhang, W.; Du, Y.; Wang, C.; Zhang, C. Enhanced hydrophilic and antipollution properties of PES membrane by anchoring SiO2/HPAN nanomaterial. ACS Sustain. Chem. Eng. 2017, 5, 7812–7823. [Google Scholar] [CrossRef]
  208. Qin, A.; Wu, X.; Ma, B.; Zhao, X.; He, C. Enhancing the antifouling property of Poly(vinylidene fluoride)/SiO2 hybrid membrane through TIPS method. J. Mater. Sci. 2014, 49, 7797–7808. [Google Scholar] [CrossRef]
  209. Yu, Z.; Liu, X.; Zhao, F.; Liang, X.; Tian, Y. Fabrication of a low-cost nano-SiO2/PVC composite ultrafiltration membrane and its antifouling performance. J. Appl. Polym. Sci. 2015, 132. [Google Scholar] [CrossRef]
  210. Zinadini, S.; Zinatizadeh, A.; Rahimi, M.; Vatanpour, V. Magnetic field-augmented coagulation bath during phase inversion for preparation of ZnFe2O4/SiO2/PES nanofiltration membrane: A novel method for flux enhancement and fouling resistance. J. Ind. Eng. Chem. 2017, 46, 9–18. [Google Scholar] [CrossRef]
  211. Ghandashtani, M.B.; Ashtiani, F.Z.; Karimi, M.; Fouladitajar, A. A novel approach to fabricate high performance nano-SiO2 embedded PES membranes for microfiltration of oil-in-water emulsion. Appl. Surf. Sci. 2015, 349, 393–402. [Google Scholar] [CrossRef]
  212. Yin, J.; Zhou, J. Novel polyethersulfone hybrid ultrafiltration membrane prepared with SiO2-g-(PDMAEMA-co-PDMAPS) and its antifouling performances in oil-in-water emulsion application. Desalination 2015, 365, 46–56. [Google Scholar] [CrossRef]
  213. Khodadousti, S.; Zokaee Ashtiani, F.; Karimi, M.; Fouladitajar, A. Preparation and characterization of novel PES-(SiO2-g-PMAA) membranes with antifouling and hydrophilic properties for separation of oil-in-water emulsions. Polym. Adv. Technol. 2019, 30, 2221–2232. [Google Scholar] [CrossRef]
  214. Muhamad, M.S.; Salim, M.R.; Lau, W.-J. Preparation and characterization of PES/SiO2 composite ultrafiltration membrane for advanced water treatment. Korean J. Chem. Eng. 2015, 32, 2319–2329. [Google Scholar] [CrossRef]
  215. Hu, Y.; Lü, Z.; Wei, C.; Yu, S.; Liu, M.; Gao, C. Separation and antifouling properties of hydrolyzed PAN hybrid membranes prepared via in-situ sol-gel SiO2 nanoparticles growth. J. Membr. Sci. 2018, 545, 250–258. [Google Scholar] [CrossRef]
  216. Zhu, J.; Zhao, X.; He, C. Zwitterionic SiO2 nanoparticles as novel additives to improve the antifouling properties of PVDF membranes. RSC Adv. 2015, 5, 53653–53659. [Google Scholar] [CrossRef]
  217. Nasrollahi, N.; Aber, S.; Vatanpour, V.; Mahmoodi, N.M. Development of hydrophilic microporous PES ultrafiltration membrane containing CuO nanoparticles with improved antifouling and separation performance. Mater. Chem. Phys. 2019, 222, 338–350. [Google Scholar] [CrossRef]
  218. Kajau, A.; Motsa, M.; Mamba, B.B.; Mahlangu, O. Leaching of CuO Nanoparticles from PES Ultrafiltration Membranes. ACS Omega 2021, 6, 31797–31809. [Google Scholar] [CrossRef] [PubMed]
  219. Abdel-Karim, A.; Ismail, S.H.; Bayoumy, A.M.; Ibrahim, M.; Mohamed, G.G. Antifouling PES/Cu@Fe3O4 mixed matrix membranes: Quantitative structure–activity relationship (QSAR) modeling and wastewater treatment potentiality. Chem. Eng. J. 2021, 407, 126501. [Google Scholar] [CrossRef]
  220. Sivasankari, S.; Kalaivizhi, R.; Gowriboy, N. Cellulose Acetate (CA) Membrane Tailored with Fe3O4@ZnO Core Shell Nanoparticles: Fabrication, Structural analysis and Its Adsorption Analysis. ChemistrySelect 2021, 6, 2350–2359. [Google Scholar] [CrossRef]
  221. Nawi, N.S.M.; Lau, W.J.; Yusof, N.; Said, N.; Ismail, A.F. Enhancing water flux and antifouling properties of PES hollow fiber membranes via incorporation of surface-functionalized Fe3O4 nanoparticles. J. Chem. Technol. Biotechnol. 2022, 97, 1006–1020. [Google Scholar] [CrossRef]
  222. Hebbar, R.S.; Isloor, A.M.; Ananda, K.; Abdullah, M.S.; Ismail, A. Fabrication of a novel hollow fiber membrane decorated with functionalized Fe2O3 nanoparticles: Towards sustainable water treatment and biofouling control. New J. Chem. 2017, 41, 4197–4211. [Google Scholar] [CrossRef]
  223. Demirel, E.; Zhang, B.; Papakyriakou, M.; Xia, S.; Chen, Y. Fe2O3 nanocomposite PVC membrane with enhanced properties and separation performance. J. Membr. Sci. 2017, 529, 170–184. [Google Scholar] [CrossRef] [Green Version]
  224. Liu, Q.; Demirel, E.; Chen, Y.; Gong, T.; Zhang, X.; Chen, Y. Improving antifouling performance for the harvesting of Scenedesmus acuminatus using Fe2O3 nanoparticles incorporated PVC nanocomposite membranes. J. Appl. Polym. Sci. 2019, 136, 47685. [Google Scholar] [CrossRef]
  225. Cui, Y.; Zheng, J.; Wang, Z.; Li, B.; Yan, Y.; Meng, M. Magnetic induced fabrication of core-shell structure Fe3O4@TiO2 photocatalytic membrane: Enhancing photocatalytic degradation of tetracycline and antifouling performance. J. Environ. Chem. Eng. 2021, 9, 106666. [Google Scholar] [CrossRef]
  226. Vatanpour, V.; Shahsavarifar, S.; Khorshidi, S.; Masteri-Farahani, M. A novel antifouling ultrafiltration membranes prepared from percarboxylic acid functionalized SiO2 bound Fe3O4 nanoparticle (SCMNP-COOOH)/polyethersulfone nanocomposite for BSA separation and dye removal. J. Chem. Technol. Biotechnol. 2019, 94, 1341–1353. [Google Scholar] [CrossRef]
  227. Zinadini, S.; Zinatizadeh, A.; Rahimi, M.; Vatanpour, V.; Zangeneh, H.; Beygzadeh, M. Novel high flux antifouling nanofiltration membranes for dye removal containing carboxymethyl chitosan coated Fe3O4 nanoparticles. Desalination 2014, 349, 145–154. [Google Scholar] [CrossRef]
  228. Garcia-Ivars, J.; Alcaina-Miranda, M.-I.; Iborra-Clar, M.-I.; Mendoza-Roca, J.-A.; Pastor-Alcañiz, L. Enhancement in hydrophilicity of different polymer phase-inversion ultrafiltration membranes by introducing PEG/Al2O3 nanoparticles. Sep. Purif. Technol. 2014, 128, 45–57. [Google Scholar] [CrossRef]
  229. Ghazanfari, D.; Bastani, D.; Mousavi, S.A. Preparation and characterization of poly (vinyl chloride)(PVC) based membrane for wastewater treatment. J. Water Process Eng. 2017, 16, 98–107. [Google Scholar] [CrossRef]
  230. Huang, Z.; Liu, J.; Liu, Y.; Xu, Y.; Li, R.; Hong, H.; Shen, L.; Lin, H.; Liao, B.-Q. Enhanced permeability and antifouling performance of polyether sulfone (PES) membrane via elevating magnetic Ni@MXene nanoparticles to upper layer in phase inversion process. J. Membr. Sci. 2021, 623, 119080. [Google Scholar] [CrossRef]
  231. Arumugham, T.; Ouda, M.; Krishnamoorthy, R.; Hai, A.; Gnanasundaram, N.; Hasan, S.W.; Banat, F. Surface-engineered polyethersulfone membranes with inherent Fe–Mn bimetallic oxides for improved permeability and antifouling capability. Environ. Res. 2022, 204, 112390. [Google Scholar] [CrossRef]
  232. Ng, L.Y.; Mohammad, A.W.; Leo, C.P.; Hilal, N. Polymeric membranes incorporated with metal/metal oxide nanoparticles: A comprehensive review. Desalination 2013, 308, 15–33. [Google Scholar] [CrossRef]
  233. Brigante, M.; Zanini, G.; Avena, M. On the dissolution kinetics of humic acid particles: Effects of pH, temperature and Ca2+ concentration. Colloids Surf. A Physicochem. Eng. Asp. 2007, 294, 64–70. [Google Scholar] [CrossRef]
  234. Esfahani, M.R.; Pallem, V.L.; Stretz, H.A.; Wells, M.J. Humic acid disaggregation with/of gold nanoparticles: Effects of nanoparticle size and pH. Environ. Nanotechnol. Monit. Manag. 2016, 6, 54–63. [Google Scholar] [CrossRef]
  235. Piccolo, A.; Conte, P.; Cozzolino, A. Effects of mineral and monocarboxylic acids on the molecular association of dissolved humic substances. Eur. J. Soil Sci. 1999, 50, 687–694. [Google Scholar] [CrossRef]
  236. Rabiee, H.; Vatanpour, V.; Farahani, M.H.D.A.; Zarrabi, H. Improvement in flux and antifouling properties of PVC ultrafiltration membranes by incorporation of zinc oxide (ZnO) nanoparticles. Sep. Purif. Technol. 2015, 156, 299–310. [Google Scholar] [CrossRef]
  237. Purushothaman, M.; Arvind, V.; Saikia, K.; Vaidyanathan, V.K. Fabrication of highly permeable and anti-fouling performance of Poly (ether ether sulfone) nanofiltration membranes modified with zinc oxide nanoparticles. Chemosphere 2022, 286, 131616. [Google Scholar] [CrossRef]
  238. Moghadam, M.T.; Lesage, G.; Mohammadi, T.; Mericq, J.P.; Mendret, J.; Heran, M.; Faur, C.; Brosillon, S.; Hemmati, M.; Naeimpoor, F. Improved antifouling properties of TiO2/PVDF nanocomposite membranes in UV-coupled ultrafiltration. J. Appl. Polym. Sci. 2015, 132. [Google Scholar] [CrossRef]
  239. Tripathi, B.P.; Dubey, N.C.; Subair, R.; Choudhury, S.; Stamm, M. Enhanced hydrophilic and antifouling polyacrylonitrile membrane with polydopamine modified silica nanoparticles. RSC Adv. 2016, 6, 4448–4457. [Google Scholar] [CrossRef]
  240. Zhang, E.; Wang, L.; Zhang, B.; Xie, Y.; Sun, C.; Jiang, C.; Zhang, Y.; Wang, G. Modification of polyvinylidene fluoride membrane with different shaped α-Fe2O3 nanocrystals for enhanced photocatalytic oxidation performance. Mater. Chem. Phys. 2018, 214, 41–47. [Google Scholar] [CrossRef]
  241. Farjami, M.; Vatanpour, V.; Moghadassi, A. Influence of the various pore former additives on the performance and characteristics of the bare and EPVC/boehmite nanocomposite ultrafiltration membranes. Mater. Today Commun. 2019, 21, 100663. [Google Scholar] [CrossRef]
  242. Farjami, M.; Vatanpour, V.; Moghadassi, A. Effect of nanoboehmite/poly (ethylene glycol) on the performance and physiochemical attributes EPVC nano-composite membranes in protein separation. Chem. Eng. Res. Des. 2020, 156, 371–383. [Google Scholar] [CrossRef]
  243. Vatanpour, V.; Madaeni, S.S.; Rajabi, L.; Zinadini, S.; Derakhshan, A.A. Boehmite nanoparticles as a new nanofiller for preparation of antifouling mixed matrix membranes. J. Membr. Sci. 2012, 401, 132–143. [Google Scholar] [CrossRef]
  244. Zhu, J.; Tian, M.; Zhang, Y.; Zhang, H.; Liu, J. Fabrication of a novel “loose” nanofiltration membrane by facile blending with Chitosan–Montmorillonite nanosheets for dyes purification. Chem. Eng. J. 2015, 265, 184–193. [Google Scholar] [CrossRef]
  245. Ang, M.B.M.Y.; Devanadera, K.P.O.; Duena, A.N.R.; Luo, Z.-Y.; Chiao, Y.-H.; Millare, J.C.; Aquino, R.R.; Huang, S.-H.; Lee, K.-R. Modifying cellulose acetate mixed-matrix membranes for improved oil–water separation: Comparison between sodium and organo-montmorillonite as particle additives. Membranes 2021, 11, 80. [Google Scholar] [CrossRef] [PubMed]
  246. El-Zahhar, A.A.; Alghamdi, M.M.; Asiri, B.M. Poly (vinyl chloride)-MMT composite membranes with enhanced properties and separation performance. Desalin. Water Treat 2019, 155, 381–389. [Google Scholar] [CrossRef]
  247. Shokri, E.; Shahed, E.; Hermani, M.; Etemadi, H. Towards enhanced fouling resistance of PVC ultrafiltration membrane using modified montmorillonite with folic acid. Appl. Clay Sci. 2021, 200, 105906. [Google Scholar] [CrossRef]
  248. Ahmad, T.; Guria, C.; Shekhar, S. Effects of inorganic salts in the casting solution on morphology of poly (vinyl chloride)/bentonite ultrafiltration membranes. Mater. Chem. Phys. 2022, 280, 125805. [Google Scholar] [CrossRef]
  249. Kumar, P.; Dixit, S.; Yadav, V.L. Effect of hydrophilic bentonite nano particle on the performance of polyvinylchloride membrane. Mater. Res. Express 2019, 6, 126415. [Google Scholar] [CrossRef]
  250. Zhang, Y.; Zhao, J.; Chu, H.; Zhou, X.; Wei, Y. Effect of modified attapulgite addition on the performance of a PVDF ultrafiltration membrane. Desalination 2014, 344, 71–78. [Google Scholar] [CrossRef]
  251. Kallem, P.; Bharath, G.; Rambabu, K.; Srinivasakannan, C.; Banat, F. Improved permeability and antifouling performance of polyethersulfone ultrafiltration membranes tailored by hydroxyapatite/boron nitride nanocomposites. Chemosphere 2021, 268, 129306. [Google Scholar] [CrossRef]
  252. Kallem, P.; Ouda, M.; Bharath, G.; Hasan, S.W.; Banat, F. Enhanced water permeability and fouling resistance properties of ultrafiltration membranes incorporated with hydroxyapatite decorated orange-peel-derived activated carbon nanocomposites. Chemosphere 2022, 286, 131799. [Google Scholar] [CrossRef]
  253. Mu, Y.; Zhu, K.; Luan, J.; Zhang, S.; Zhang, C.; Na, R.; Yang, Y.; Zhang, X.; Wang, G. Fabrication of hybrid ultrafiltration membranes with improved water separation properties by incorporating environmentally friendly taurine modified hydroxyapatite nanotubes. J. Membr. Sci. 2019, 577, 274–284. [Google Scholar] [CrossRef]
  254. Zhang, X.; Lang, W.-Z.; Xu, H.-P.; Yan, X.; Guo, Y.-J. The effects of hydroxyapatite nano whiskers and its synergism with polyvinylpyrrolidone on Poly(vinylidene fluoride) hollow fiber ultrafiltration membranes. RSC Adv. 2015, 5, 21532–21543. [Google Scholar] [CrossRef]
  255. Farahani, M.H.D.A.; Vatanpour, V. A comprehensive study on the performance and antifouling enhancement of the PVDF mixed matrix membranes by embedding different nanoparticulates: Clay, functionalized carbon nanotube, SiO2 and TiO2. Sep. Purif. Technol. 2018, 197, 372–381. [Google Scholar] [CrossRef]
  256. Daraei, P.; Ghaemi, N. Synergistic effect of Cloisite 15A and 30B nanofillers on the characteristics of nanocomposite polyethersulfone membrane. Appl. Clay Sci. 2019, 172, 96–105. [Google Scholar] [CrossRef]
  257. Arkaban, M.; Mahdavian, L.; Arkaban, H. Synthesis of Poly(vinylidene Fluoride)/Modified SBA-15 Nanoparticles Composite Membrane for Water Purification. Silicon 2020, 12, 2031–2039. [Google Scholar] [CrossRef]
  258. Bandehali, S.; Parvizian, F.; Ruan, H.; Moghadassi, A.; Shen, J.; Figoli, A.; Adeleye, A.S.; Hilal, N.; Matsuura, T.; Drioli, E. A planned review on designing of high-performance nanocomposite nanofiltration membranes for pollutants removal from water. J. Ind. Eng. Chem. 2021, 101, 78–125. [Google Scholar] [CrossRef]
  259. Guo, J.; Sotto, A.; Martín, A.; Kim, J. Preparation and characterization of polyethersulfone mixed matrix membranes embedded with Ti-or Zr-incorporated SBA-15 materials. J. Ind. Eng. Chem. 2017, 45, 257–265. [Google Scholar] [CrossRef]
  260. Moradi, G.; Zinadini, S.; Rahimi, M.; Shiri, F. Efficient Zn2+, Pb2+, and Ni2+ removal using antifouling mixed matrix nanofiltration membrane with curcumin modified mesoporous Santa Barbara Amorphous-15 (Cur-SBA-15) filler. J. Environ. Chem. Eng. 2022, 10, 107302. [Google Scholar] [CrossRef]
  261. Wang, H.; Lu, X.; Lu, D.; Wang, P.; Ma, J. Development of a high-performance polysulfone hybrid ultrafiltration membrane using hydrophilic polymer-functionalized mesoporous SBA−15 as filler. J. Appl. Polym. Sci. 2019, 136, 47353. [Google Scholar] [CrossRef]
  262. Wang, Y.; Shen, H.; Cui, C.; Hou, L.; Chen, W.; Liu, Q.; Xu, J.; Wang, Z.; Hu, J. Towards to better permeability and antifouling sulfonated poly (aryl ether ketone sulfone) with carboxyl group ultrafiltration membrane blending with amine functionalization of SBA-15. Sep. Purif. Technol. 2021, 265, 118512. [Google Scholar] [CrossRef]
  263. Amid, M.; Nabian, N.; Delavar, M. Performance evaluation and modeling study of PC blended membranes incorporated with SDS-modified and unmodified halloysite nanotubes in the separation of oil from water. J. Environ. Chem. Eng. 2021, 9, 105237. [Google Scholar] [CrossRef]
  264. Duan, L.; Huang, W.; Zhang, Y. High-flux, antibacterial ultrafiltration membranes by facile blending with N-halamine grafted halloysite nanotubes. Rsc Adv. 2015, 5, 6666–6674. [Google Scholar] [CrossRef]
  265. Grylewicz, A.; Szymański, K.; Darowna, D.; Mozia, S. Influence of Polymer Solvents on the Properties of Halloysite-Modified Polyethersulfone Membranes Prepared by Wet Phase Inversion. Molecules 2021, 26, 2768. [Google Scholar] [CrossRef] [PubMed]
  266. Kamal, N.; Ahzi, S.; Kochkodan, V. Polysulfone/halloysite composite membranes with low fouling properties and enhanced compaction resistance. Appl. Clay Sci. 2020, 199, 105873. [Google Scholar] [CrossRef]
  267. Kamal, N.; Kochkodan, V.; Zekri, A.; Ahzi, S. Polysulfone membranes embedded with halloysites nanotubes: Preparation and properties. Membranes 2019, 10, 2. [Google Scholar] [CrossRef] [Green Version]
  268. Liu, Z.; Mi, Z.; Jin, S.; Wang, C.; Wang, D.; Zhao, X.; Zhou, H.; Chen, C. The influence of sulfonated hyperbranched polyethersulfone-modified halloysite nanotubes on the compatibility and water separation performance of polyethersulfone hybrid ultrafiltration membranes. J. Membr. Sci. 2018, 557, 13–23. [Google Scholar] [CrossRef]
  269. Ma, J.; He, Y.; Zeng, G.; Yang, X.; Chen, X.; Zhou, L.; Peng, L.; Sengupta, A. High-flux PVDF membrane incorporated with β-cyclodextrin modified halloysite nanotubes for dye rejection and Cu (II) removal from water. Polym. Adv. Technol. 2018, 29, 2704–2714. [Google Scholar] [CrossRef]
  270. Mishra, G.; Mukhopadhyay, M. Enhanced antifouling performance of halloysite nanotubes (HNTs) blended poly (vinyl chloride)(PVC/HNTs) ultrafiltration membranes: For water treatment. J. Ind. Eng. Chem. 2018, 63, 366–379. [Google Scholar] [CrossRef]
  271. Mozia, S.; Grylewicz, A.; Zgrzebnicki, M.; Darowna, D.; Czyżewski, A. Investigations on the properties and performance of mixed-matrix polyethersulfone membranes modified with halloysite nanotubes. Polymers 2019, 11, 671. [Google Scholar] [CrossRef] [Green Version]
  272. Ouda, M.; Hai, A.; Krishnamoorthy, R.; Govindan, B.; Othman, I.; Kui, C.C.; Choi, M.Y.; Hasan, S.W.; Banat, F. Surface tuned polyethersulfone membrane using an iron oxide functionalized halloysite nanocomposite for enhanced humic acid removal. Environ. Res. 2022, 204, 112113. [Google Scholar] [CrossRef]
  273. Park, S.; Yang, E.; Park, H.; Choi, H. Fabrication of functionalized halloysite nanotube blended ultrafiltration membranes for high flux and fouling resistance. Environ. Eng. Res. 2019, 25, 771–778. [Google Scholar] [CrossRef] [Green Version]
  274. Wan Ikhsan, S.N.; Yusof, N.; Mat Nawi, N.I.; Bilad, M.R.; Shamsuddin, N.; Aziz, F.; Ismail, A.F. Halloysite nanotube-ferrihydrite incorporated polyethersulfone mixed matrix membrane: Effect of nanocomposite loading on the antifouling performance. Polymers 2021, 13, 441. [Google Scholar] [CrossRef] [PubMed]
  275. Wang, B.; He, H.; Li, Y.; Liu, C.; Bai, J.; Zhou, X.; Li, J.; Wang, Y. Fabrication of green poly (L-lactic acid) hybrid membrane through incorporation of functionalized natural halloysite nanotubes. J. Chem. Technol. Biotechnol. 2022, 97, 1676–1683. [Google Scholar] [CrossRef]
  276. Yu, H.; Zhang, Y.; Sun, X.; Liu, J.; Zhang, H. Improving the antifouling property of polyethersulfone ultrafiltration membrane by incorporation of dextran grafted halloysite nanotubes. Chem. Eng. J. 2014, 237, 322–328. [Google Scholar] [CrossRef]
  277. Zeng, G.; He, Y.; Zhan, Y.; Zhang, L.; Pan, Y.; Zhang, C.; Yu, Z. Novel polyvinylidene fluoride nanofiltration membrane blended with functionalized halloysite nanotubes for dye and heavy metal ions removal. J. Hazard. Mater. 2016, 317, 60–72. [Google Scholar] [CrossRef]
  278. Zeng, G.; He, Y.; Zhan, Y.; Zhang, L.; Shi, H.; Yu, Z. Preparation of a novel Poly(vinylidene fluoride) ultrafiltration membrane by incorporation of 3-aminopropyltriethoxysilane-grafted halloysite nanotubes for oil/water separation. Ind. Eng. Chem. Res. 2016, 55, 1760–1767. [Google Scholar] [CrossRef]
  279. Bai, Z.; Wang, L.; Liu, C.; Yang, C.; Lin, G.; Liu, S.; Jia, K.; Liu, X. Interfacial coordination mediated surface segregation of halloysite nanotubes to construct a high-flux antifouling membrane for oil-water emulsion separation. J. Membr. Sci. 2021, 620, 118828. [Google Scholar] [CrossRef]
  280. Arefi-Oskoui, S.; Vatanpour, V.; Khataee, A. Development of a novel high-flux PVDF-based ultrafiltration membrane by embedding Mg-Al nanolayered double hydroxide. J. Ind. Eng. Chem. 2016, 41, 23–32. [Google Scholar] [CrossRef]
  281. Balcik, C.; Ozbey-Unal, B.; Cifcioglu-Gozuacik, B.; Keyikoglu, R.; Karagunduz, A.; Khataee, A. Fabrication of PSf nanocomposite membranes incorporated with ZnFe layered double hydroxide for separation and antifouling aspects. Sep. Purif. Technol. 2022, 285, 120354. [Google Scholar] [CrossRef]
  282. Li, C.; Sun, W.; Lu, Z.; Ao, X.; Li, S. Ceramic nanocomposite membranes and membrane fouling: A review. Water Res. 2020, 175, 115674. [Google Scholar] [CrossRef]
  283. Morihama, A.C.D.; Mierzwa, J.C. Clay nanoparticles effects on performance and morphology of Poly(vinylidene fluoride) membranes. Braz. J. Chem. Eng. 2014, 31, 79–93. [Google Scholar] [CrossRef] [Green Version]
  284. Seshasayee, M.; Yu, Z.; Arthanareeswaran, G.; Das, D. Preparation of nanoclay embedded polymeric membranes for the filtration of natural organic matter (NOM) in a circular crossflow filtration system. J. Water Process Eng. 2020, 37, 101408. [Google Scholar] [CrossRef]
  285. Dlamini, D.S.; Li, J.; Mamba, B.B. Critical review of montmorillonite/polymer mixed-matrix filtration membranes: Possibilities and challenges. Appl. Clay Sci. 2019, 168, 21–30. [Google Scholar] [CrossRef]
  286. Koulivand, H.; Shahbazi, A.; Vatanpour, V.; Rahmandoost, M. Novel antifouling and antibacterial polyethersulfone membrane prepared by embedding nitrogen-doped carbon dots for efficient salt and dye rejection. Mater. Sci. Eng. C 2020, 111, 110787. [Google Scholar] [CrossRef] [PubMed]
  287. Koulivand, H.; Shahbazi, A.; Vatanpour, V.; Rahmandoust, M. Development of carbon dot-modified polyethersulfone membranes for enhancement of nanofiltration, permeation and antifouling performance. Sep. Purif. Technol. 2020, 230, 115895. [Google Scholar] [CrossRef]
  288. Li, Y.; Huang, S.; Zhou, S.; Fane, A.G.; Zhang, Y.; Zhao, S. Enhancing water permeability and fouling resistance of polyvinylidene fluoride membranes with carboxylated nanodiamonds. J. Membr. Sci. 2018, 556, 154–163. [Google Scholar] [CrossRef]
  289. Hosseinpour, S.; Azimian-Kivi, M.; Jafarzadeh, Y.; Yegani, R. Pharmaceutical wastewater treatment using polypropylene membranes incorporated with carboxylated and PEG-grafted nanodiamond in membrane bioreactor (MBR). Water Environ. J. 2021, 35, 1249–1259. [Google Scholar] [CrossRef]
  290. Li, Y.; He, S.; Zhou, Z.; Zhou, S.; Huang, S.; Fane, A.G.; Zheng, C.; Zhang, Y.; Zhao, S. Carboxylated Nanodiamond-enhanced photocatalytic membranes with improved antifouling and self-cleaning properties. Ind. Eng. Chem. Res. 2020, 59, 3538–3549. [Google Scholar] [CrossRef]
  291. Shahida, S.; Siddiqa, A.; Salim, N.; Qaisar, S.; Khan, M.I.; Farooq, U.; Shanableh, A.; Elboughdiri, N.; Kolsi, L.; Bouazzi, Y. Fabrication and characterization of nanocomposite membranes for the rejection of textile dye. Inorg. Nano-Met. Chem. 2022, 97, 1676–1683. [Google Scholar] [CrossRef]
  292. Vatanpour, V.; Naeeni, R.S.E.; Ghadimi, A.; Karami, A.; Sadatnia, B. Effect of detonation nanodiamond on the properties and performance of polyethersulfone nanocomposite membrane. Diam. Relat. Mater. 2018, 90, 244–255. [Google Scholar] [CrossRef]
  293. Ajibadea, T.F.; Xuc, J.; Tiana, H.; Guanc, L.; Zhanga, K. O-MWCNT/PAN/PVDF ultrafiltration membranes with boosted properties for oil and water separation. Desalin. Water Treat. 2021, 224, 122–135. [Google Scholar] [CrossRef]
  294. Hudaib, B.; Gomes, V.; Shi, J.; Zhou, C.; Liu, Z. Poly(vinylidene fluoride)/polyaniline/MWCNT nanocomposite ultrafiltration membrane for natural organic matter removal. Sep. Purif. Technol. 2018, 190, 143–155. [Google Scholar] [CrossRef]
  295. Huhn-Ibarra, M.J.; Loría-Bastarrachea, M.I.; Duarte-Aranda, S.; Montes-Luna, A.d.J.; Ortiz-Espinoza, J.; González-Díaz, M.O.; Aguilar-Vega, M. PPSU dual layer hollow fiber mixed matrix membranes with functionalized MWCNT for enhanced antifouling, salt and dye rejection in water treatment. J. Appl. Polym. Sci. 2022, 139, e53203. [Google Scholar] [CrossRef]
  296. Irfan, M.; Basri, H.; Irfan, M.; Lau, W.-J. An acid functionalized MWCNT/PVP nanocomposite as a new additive for fabrication of an ultrafiltration membrane with improved anti-fouling resistance. RSC Adv. 2015, 5, 95421–95432. [Google Scholar] [CrossRef]
  297. Liu, C.; Wang, W.; Li, Y.; Cui, F.; Xie, C.; Zhu, L.; Shan, B. PMWCNT/PVDF ultrafiltration membranes with enhanced antifouling properties intensified by electric field for efficient blood purification. J. Membr. Sci. 2019, 576, 48–58. [Google Scholar] [CrossRef]
  298. Parsamanesh, M.; Mansourpanah, Y.; Dadkhah Tehrani, A. Improving the efficacy of PES-based mixed matrix membranes incorporated with citric acid–amylose-modified MWCNTs for HA removal from water. Polym. Bull. 2021, 78, 1293–1311. [Google Scholar] [CrossRef]
  299. Peydayesh, M.; Mohammadi, T.; Bakhtiari, O. Water desalination via novel positively charged hybrid nanofiltration membranes filled with hyperbranched polyethyleneimine modified MWCNT. J. Ind. Eng. Chem. 2019, 69, 127–140. [Google Scholar] [CrossRef]
  300. Rahimi, Z.; Zinatizadeh, A.; Zinadini, S. Preparation of high antibiofouling amino functionalized MWCNTs/PES nanocomposite ultrafiltration membrane for application in membrane bioreactor. J. Ind. Eng. Chem. 2015, 29, 366–374. [Google Scholar] [CrossRef]
  301. Zhang, Y.; Wang, C.; Zhang, L.; Shi, J.; Yuan, H.; Lu, J. Doubly modified MWCNTs embedded in polyethersulfone (PES) ultrafiltration membrane and its anti-fouling performance. J. Polym. Eng. 2022, 42, 885–898. [Google Scholar] [CrossRef]
  302. Haghighat, N.; Vatanpour, V. Fouling decline and retention increase of polyvinyl chloride nanofiltration membranes blended by polypyrrole functionalized multiwalled carbon nanotubes. Mater. Today Commun. 2020, 23, 100851. [Google Scholar] [CrossRef]
  303. Mahdavi, M.R.; Delnavaz, M.; Vatanpour, V.; Farahbakhsh, J. Effect of blending polypyrrole coated multiwalled carbon nanotube on desalination performance and antifouling property of thin film nanocomposite nanofiltration membranes. Sep. Purif. Technol. 2017, 184, 119–127. [Google Scholar] [CrossRef]
  304. Mohsenpour, S.; Leaper, S.; Shokri, J.; Alberto, M.; Gorgojo, P. Effect of graphene oxide in the formation of polymeric asymmetric membranes via phase inversion. J. Membr. Sci. 2022, 641, 119924. [Google Scholar] [CrossRef]
  305. Xia, S.; Ni, M. Preparation of Poly(vinylidene fluoride) membranes with graphene oxide addition for natural organic matter removal. J. Membr. Sci. 2015, 473, 54–62. [Google Scholar] [CrossRef]
  306. Meng, N.; Priestley, R.C.E.; Zhang, Y.; Wang, H.; Zhang, X. The effect of reduction degree of GO nanosheets on microstructure and performance of PVDF/GO hybrid membranes. J. Membr. Sci. 2016, 501, 169–178. [Google Scholar] [CrossRef]
  307. Luo, X.; He, Z.; Gong, H.; He, L. Recent advances in oil-water separation materials with special wettability modified by graphene and its derivatives: A review. Chem. Eng. Process.-Process Intensif. 2022, 170, 108678. [Google Scholar] [CrossRef]
  308. Vatanpour, V.; Khadem, S.S.M.; Masteri-Farahani, M.; Mosleh, N.; Ganjali, M.R.; Badiei, A.; Pourbashir, E.; Mashhadzadeh, A.H.; Munir, M.T.; Mahmodi, G. Anti-fouling and permeable polyvinyl chloride nanofiltration membranes embedded by hydrophilic graphene quantum dots for dye wastewater treatment. J. Water Process Eng. 2020, 38, 101652. [Google Scholar] [CrossRef]
  309. Raval, H.D.; Makwana, P.; Sharma, S. Biofouling of polysulfone and polysulfone-graphene oxide nanocomposite membrane and foulant removal. Mater. Res. Express 2018, 5, 065322. [Google Scholar] [CrossRef]
  310. Xue, J.; Wang, S.; Han, X.; Wang, Y.; Hua, X.; Li, J. Chitosan-functionalized graphene oxide for enhanced permeability and antifouling of ultrafiltration membranes. Chem. Eng. Technol. 2018, 41, 270–277. [Google Scholar] [CrossRef] [Green Version]
  311. Yoon, Y.; Kye, H.; Yang, W.S.; Kang, J.-W. Comparing graphene oxide and reduced graphene oxide as blending materials for polysulfone and polyvinylidene difluoride membranes. Appl. Sci. 2020, 10, 2015. [Google Scholar] [CrossRef] [Green Version]
  312. Karkooti, A.; Yazdi, A.Z.; Chen, P.; McGregor, M.; Nazemifard, N.; Sadrzadeh, M. Development of advanced nanocomposite membranes using graphene nanoribbons and nanosheets for water treatment. J. Membr. Sci. 2018, 560, 97–107. [Google Scholar] [CrossRef]
  313. Rashidi, R.; Khakpour, S.; Masoumi, S.; Jafarzadeh, Y. Effects of GO-PEG on the performance and structure of PVC ultrafiltration membranes. Chem. Eng. Res. Des. 2022, 177, 815–825. [Google Scholar] [CrossRef]
  314. Vatanpour, V.; Khadem, S.S.M.; Dehqan, A.; Al-Naqshabandi, M.A.; Ganjali, M.R.; Hassani, S.S.; Rashid, M.R.; Saeb, M.R.; Dizge, N. Efficient removal of dyes and proteins by nitrogen-doped porous graphene blended polyethersulfone nanocomposite membranes. Chemosphere 2021, 263, 127892. [Google Scholar] [CrossRef] [PubMed]
  315. Ma, H.; Xie, Q.; Wu, C.; Shen, L.; Hong, Z.; Zhang, G.; Lu, Y.; Shao, W. A facile approach to enhance performance of PVDF-matrix nanocomposite membrane via manipulating migration behavior of graphene oxide. J. Membr. Sci. 2019, 590, 117268. [Google Scholar] [CrossRef]
  316. Zhang, W.; Zhang, Y.; Wang, Y.; Tian, S.; Han, N.; Li, W.; Wang, W.; Liu, H.; Yan, X.; Zhang, X. Fluffy-like amphiphilic graphene oxide (f-GO) and its effects on improving the antifouling of PAN-based composite membranes. Desalination 2022, 527, 115575. [Google Scholar] [CrossRef]
  317. Karimipour, H.; Shahbazi, A.; Vatanpour, V. Fouling decline and retention increase of polyethersulfone membrane by incorporating melamine-based dendrimer amine functionalized graphene oxide nanosheets (GO/MDA). J. Environ. Chem. Eng. 2021, 9, 104849. [Google Scholar] [CrossRef]
  318. Salim, A.; Abbas, M.A.; Khan, I.A.; Khan, M.Z.; Javaid, F.; Mushtaq, S.; Batool, M.; Yasir, M.; Khan, A.L.; Khan, A.U. Graphene oxide incorporated polyether sulfone nanocomposite antifouling ultrafiltration membranes with enhanced hydrophilicity. Mater. Res. Express 2022, 9, 075503. [Google Scholar] [CrossRef]
  319. Zhao, C.; Xu, X.; Chen, J.; Wang, G.; Yang, F. Highly effective antifouling performance of PVDF/graphene oxide composite membrane in membrane bioreactor (MBR) system. Desalination 2014, 340, 59–66. [Google Scholar] [CrossRef]
  320. Guo, J.; Zhang, Y.; Chen, F.; Chai, Y. A Membrane with Strong Resistance to Organic and Biological Fouling Using Graphene Oxide and D-Tyrosine as Modifiers. Membranes 2022, 12, 486. [Google Scholar] [CrossRef]
  321. Gholami, N.; Mahdavi, H. Nanofiltration composite membranes of polyethersulfone and graphene oxide and sulfonated graphene oxide. Adv. Polym. Technol. 2018, 37, 3529–3541. [Google Scholar] [CrossRef] [Green Version]
  322. Tomczak, E.; Blus, M. Polymer membranes from polyvinylidene fluoride or cellulose acetate improved graphene oxide used in the UF process. Desalin. Water Treat. 2021, 214, 146–154. [Google Scholar] [CrossRef]
  323. Nawaz, H.; Umar, M.; Ullah, A.; Razzaq, H.; Zia, K.M.; Liu, X. Polyvinylidene fluoride nanocomposite super hydrophilic membrane integrated with Polyaniline-Graphene oxide nano fillers for treatment of textile effluents. J. Hazard. Mater. 2021, 403, 123587. [Google Scholar] [CrossRef]
  324. Nguyen, H.T.V.; Ngo, T.H.A.; Do, K.D.; Nguyen, M.N.; Dang, N.T.T.; Nguyen, T.T.H.; Vien, V.; Vu, T.A. Preparation and characterization of a hydrophilic polysulfone membrane using graphene oxide. J. Chem. 2019, 2019, 3164373. [Google Scholar] [CrossRef] [Green Version]
  325. Alkhouzaam, A.; Qiblawey, H. Synergetic effects of dodecylamine-functionalized graphene oxide nanoparticles on antifouling and antibacterial properties of polysulfone ultrafiltration membranes. J. Water Process Eng. 2021, 42, 102120. [Google Scholar] [CrossRef]
  326. Hu, M.; Cui, Z.; Li, J.; Zhang, L.; Mo, Y.; Dlamini, D.S.; Wang, H.; He, B.; Li, J.; Matsuyama, H. Ultra-low graphene oxide loading for water permeability, antifouling and antibacterial improvement of polyethersulfone/sulfonated polysulfone ultrafiltration membranes. J. Colloid Interface Sci. 2019, 552, 319–331. [Google Scholar] [CrossRef] [PubMed]
  327. Rahimi, A.; Mahdavi, H. Zwitterionic-functionalized GO/PVDF nanocomposite membranes with improved anti-fouling properties. J. Water Process Eng. 2019, 32, 100960. [Google Scholar] [CrossRef]
  328. Jin, F.; Lv, W.; Zhang, C.; Li, Z.; Su, R.; Qi, W.; Yang, Q.-H.; He, Z. High-performance ultrafiltration membranes based on polyethersulfone–graphene oxide composites. Rsc Adv. 2013, 3, 21394–21397. [Google Scholar] [CrossRef]
  329. Wang, H.T.; Ao, D.; Lu, M.C.; Chang, N. Alteration of the morphology of polyvinylidene fluoride membrane by incorporating MOF-199 nanomaterials for improving water permeation with antifouling and antibacterial property. J. Chin. Chem. Soc. 2020, 67, 1807–1817. [Google Scholar] [CrossRef]
  330. Xiao, S.; Huo, X.; Fan, S.; Zhao, K.; Yu, S.; Tan, X. Design and synthesis of Al-MOF/PPSU mixed matrix membrane with pollution resistance. Chin. J. Chem. Eng. 2021, 29, 110–120. [Google Scholar] [CrossRef]
  331. Samari, M.; Zinadini, S.; Zinatizadeh, A.A.; Jafarzadeh, M.; Gholami, F. Designing of a novel polyethersulfone (PES) ultrafiltration (UF) membrane with thermal stability and high fouling resistance using melamine-modified zirconium-based metal-organic framework (UiO-66-NH2/MOF). Sep. Purif. Technol. 2020, 251, 117010. [Google Scholar] [CrossRef]
  332. Rameesha, L.; Rana, D.; Kaleekkal, N.J.; Nagendran, A. Efficacy of MOF-199 in improvement of permeation, morphological, antifouling and antibacterial characteristics of polyvinylidene fluoride membranes. New J. Chem. 2022, 46, 7638–7649. [Google Scholar] [CrossRef]
  333. Shukla, A.K.; Alam, J.; Ali, F.A.A.; Alhoshan, M. A highly permeable zinc-based MOF/polyphenylsulfone composite membrane with elevated antifouling properties. Chem. Commun. 2020, 56, 5231–5234. [Google Scholar] [CrossRef]
  334. Wang, H.; Chen, H.; Zeng, Y.; Chen, G.; Cui, M. Hydrophilic modification and anti-fouling properties of PVDF ultrafiltration membrane via blending of nano-particle MIL-101 MOFs. Dig. J. Nanomater. Biostructures (DJNB) 2021, 16, 515–526. [Google Scholar]
  335. Ahmadipouya, S.; Mousavi, S.A.; Shokrgozar, A.; Mousavi, D.V. Improving dye removal and antifouling performance of polysulfone nanofiltration membranes by incorporation of UiO-66 metal-organic framework. J. Environ. Chem. Eng. 2022, 10, 107535. [Google Scholar] [CrossRef]
  336. Wang, Y.; Li, D.; Li, J.; Li, J.; Fan, M.; Han, M.; Liu, Z.; Li, Z.; Kong, F. Metal organic framework UiO-66 incorporated ultrafiltration membranes for simultaneous natural organic matter and heavy metal ions removal. Environ. Res. 2022, 208, 112651. [Google Scholar] [CrossRef] [PubMed]
  337. Mohammadnezhad, F.; Feyzi, M.; Zinadini, S. A novel Ce-MOF/PES mixed matrix membrane; synthesis, characterization and antifouling evaluation. J. Ind. Eng. Chem. 2019, 71, 99–111. [Google Scholar] [CrossRef]
  338. Gnanasekaran, G.; Balaguru, S.; Arthanareeswaran, G.; Das, D.B. Removal of hazardous material from wastewater by using metal organic framework (MOF) embedded polymeric membranes. Sep. Sci. Technol. 2019, 54, 434–446. [Google Scholar] [CrossRef] [Green Version]
  339. Vinothkumar, K.; Jyothi, M.S.; Lavanya, C.; Sakar, M.; Valiyaveettil, S.; Balakrishna, R.G. Strongly co-ordinated MOF-PSF matrix for selective adsorption, separation and photodegradation of dyes. Chem. Eng. J. 2022, 428, 132561. [Google Scholar] [CrossRef]
  340. Gholami, F.; Zinadini, S.; Zinatizadeh, A.; Abbasi, A. TMU-5 metal-organic frameworks (MOFs) as a novel nanofiller for flux increment and fouling mitigation in PES ultrafiltration membrane. Sep. Purif. Technol. 2018, 194, 272–280. [Google Scholar] [CrossRef]
  341. Yin, J.; Tang, H.; Liu, D.; Huang, T.; Zhu, L. Application of ZIF-67 as a crosslinker to prepare sulfonated polysulfone mixed-matrix membranes for enhanced water permeability and separation properties. Water Sci. Technol. 2021, 84, 144–158. [Google Scholar] [CrossRef]
  342. Liu, D.; Yin, J.; Tang, H.; Wang, H.; Liu, S.; Huang, T.; Fang, S.; Zhu, K.; Xie, Z. Fabrication of ZIF-67@PVDF ultrafiltration membrane with improved antifouling and separation performance for dye wastewater treatment via sulfate radical enhancement. Sep. Purif. Technol. 2021, 279, 119755. [Google Scholar] [CrossRef]
  343. Adams, F.; Nxumalo, E.; Krause, R.; Hoek, E.; Mamba, B. Application of polysulfone/cyclodextrin mixed-matrix membranes in the removal of natural organic matter from water. Phys. Chem. Earth Parts A/B/C 2014, 67, 71–78. [Google Scholar] [CrossRef]
  344. Austria, H.F.; Subrahmanya, T.M.; Setiawan, O.; Widakdo, J.; Chiao, Y.-H.; Hung, W.-S.; Wang, C.-F.; Hu, C.-C.; Lee, K.-R.; Lai, J.-Y. A review on the recent advancements in graphene-based membranes and their applications as stimuli-responsive separation materials. J. Mater. Chem. A 2021, 9, 21510–21531. [Google Scholar] [CrossRef]
  345. Bala, S.; Nithya, D.; Doraisamy, M. Exploring the effects of graphene oxide concentration on properties and antifouling performance of PEES/GO ultrafiltration membranes. High Perform. Polym. 2018, 30, 375–383. [Google Scholar] [CrossRef]
  346. Khalid, A.; Al-Juhani, A.A.; Al-Hamouz, O.C.; Laoui, T.; Khan, Z.; Atieh, M.A. Preparation and properties of nanocomposite polysulfone/multi-walled carbon nanotubes membranes for desalination. Desalination 2015, 367, 134–144. [Google Scholar] [CrossRef]
  347. Shah, P.; Murthy, C.N. Studies on the porosity control of MWCNT/polysulfone composite membrane and its effect on metal removal. J. Membr. Sci. 2013, 437, 90–98. [Google Scholar] [CrossRef]
  348. Wu, C.-J.; Maggay, I.V.; Chiang, C.-H.; Chen, W.; Chang, Y.; Hu, C.; Venault, A. Removal of tetracycline by a photocatalytic membrane reactor with MIL-53 (Fe)/PVDF mixed-matrix membrane. Chem. Eng. J. 2023, 451, 138990. [Google Scholar] [CrossRef]
  349. Deng, W.; Li, Y. Novel superhydrophilic antifouling PVDF-BiOCl nanocomposite membranes fabricated via a modified blending-phase inversion method. Sep. Purif. Technol. 2021, 254, 117656. [Google Scholar] [CrossRef]
  350. Mahdavi, H.; Zeinalipour, N.; Kerachian, M.A.; Heidari, A.A. Preparation of high-performance PVDF mixed matrix membranes incorporated with PVDF-g-PMMA copolymer and GO@SiO2 nanoparticles for dye rejection applications. J. Water Process Eng. 2022, 46, 102560. [Google Scholar] [CrossRef]
  351. Sri, K.S.; Nair, A.K.; Babu, P.J. Synthesis and characterization of silver decorated polysulfone/cellulose acetate hybrid ultrafiltration membranes using functionalized TiO2 nanoparticles. Desalin. Water Treat. 2017, 76, 112–120. [Google Scholar] [CrossRef] [Green Version]
  352. Sisay, E.J.; Veréb, G.; Pap, Z.; Gyulavári, T.; Ágoston, Á.; Kopniczky, J.; Hodúr, C.; Arthanareeswaran, G.; Arumugam, G.K.S.; László, Z. Visible-light-driven photocatalytic PVDF-TiO2/CNT/BiVO4 hybrid nanocomposite ultrafiltration membrane for dairy wastewater treatment. Chemosphere 2022, 307, 135589. [Google Scholar] [CrossRef]
  353. Mahdavi, H.; Zeinalipour, N.; Heidari, A.A. Fabrication of PVDF mixed matrix nanofiltration membranes incorporated with TiO2 nanoparticles and an amphiphilic PVDF-g-PMMA copolymer. J. Appl. Polym. Sci. 2022, 139, e52740. [Google Scholar] [CrossRef]
  354. Zhu, J.; Zhou, S.; Li, M.; Xue, A.; Zhao, Y.; Peng, W.; Xing, W. PVDF mixed matrix ultrafiltration membrane incorporated with deformed rebar-like Fe3O4–palygorskite nanocomposites to enhance strength and antifouling properties. J. Membr. Sci. 2020, 612, 118467. [Google Scholar] [CrossRef]
  355. Mulder, M. Preparation of Synthetic Membranes. In Basic Principles of Membrane Technology; Mulder, M., Ed.; Springer: Dordrecht, The Netherlands, 1996; pp. 71–156. [Google Scholar]
  356. Wang, D.-M.; Venault, A.; Lai, J.-Y. Chapter 2—Fundamentals of nonsolvent-induced phase separation. In Hollow Fiber Membranes; Chung, T.-S., Feng, Y., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 13–56. [Google Scholar]
  357. Venault, A.; Liu, Y.-H.; Wu, J.-R.; Yang, H.-S.; Chang, Y.; Lai, J.-Y.; Aimar, P. Low-biofouling membranes prepared by liquid-induced phase separation of the PVDF/polystyrene-b-poly (ethylene glycol) methacrylate blend. J. Membr. Sci. 2014, 450, 340–350. [Google Scholar] [CrossRef] [Green Version]
  358. Khalil, H.; Hegab, H.M.; Nassar, L.; Wadi, V.S.; Naddeo, V.; Yousef, A.F.; Banat, F.; Hasan, S.W. Asymmetrical ultrafiltration membranes based on polylactic acid for the removal of organic substances from wastewater. J. Water Process Eng. 2022, 45, 102510. [Google Scholar] [CrossRef]
  359. Nainar, M.G.; Jayaraman, K.; Meyyappan, H.K.; Miranda, L.R. Antifouling properties of Poly(vinylidene fluoride)-incorporated cellulose acetate composite ultrafiltration membranes. Korean J. Chem. Eng. 2020, 37, 2248–2261. [Google Scholar] [CrossRef]
  360. Venault, A.; Jumao-as-Leyba, A.J.; Yang, Z.-R.; Carretier, S.; Chang, Y. Formation mechanisms of low-biofouling PVDF/F127 membranes prepared by VIPS process. J. Taiwan Inst. Chem. Eng. 2016, 62, 297–306. [Google Scholar] [CrossRef]
  361. Venault, A.; Wu, J.-R.; Chang, Y.; Aimar, P. Fabricating hemocompatible bi-continuous PEGylated PVDF membranes via vapor-induced phase inversion. J. Membr. Sci. 2014, 470, 18–29. [Google Scholar] [CrossRef] [Green Version]
  362. Mat Nawi, N.I.; Chean, H.M.; Shamsuddin, N.; Bilad, M.R.; Narkkun, T.; Faungnawakij, K.; Khan, A.L. Development of hydrophilic PVDF membrane using vapour induced phase separation method for produced water treatment. Membranes 2020, 10, 121. [Google Scholar] [CrossRef]
  363. Hsu, C.-H.; Venault, A.; Chang, Y. Facile zwitterionization of polyvinylidene fluoride microfiltration membranes for biofouling mitigation. J. Membr. Sci. 2022, 648, 120348. [Google Scholar] [CrossRef]
  364. Li, C.-L.; Wang, D.-M.; Deratani, A.; Quémener, D.; Bouyer, D.; Lai, J.-Y. Insight into the preparation of poly(vinylidene fluoride) membranes by vapor-induced phase separation. J. Membr. Sci. 2010, 361, 154–166. [Google Scholar] [CrossRef]
  365. Zhang, P.; Rajabzadeh, S.; Venault, A.; Wang, S.; Shen, Q.; Jia, Y.; Fang, C.; Kato, N.; Chang, Y.; Matsuyama, H. One-step entrapment of a PS-PEGMA amphiphilic copolymer on the outer surface of a hollow fiber membrane via TIPS process using triple-orifice spinneret. J. Membr. Sci. 2021, 638, 119712. [Google Scholar] [CrossRef]
  366. Li, Z.-K.; Lang, W.-Z.; Miao, W.; Yan, X.; Guo, Y.-J. Preparation and properties of PVDF/SiO2@GO nanohybrid membranes via thermally induced phase separation method. J. Membr. Sci. 2016, 511, 151–161. [Google Scholar] [CrossRef]
  367. Chen, S.; Li, L.; Zhao, C.; Zheng, J. Surface hydration: Principles and applications toward low-fouling/nonfouling biomaterials. Polymer 2010, 51, 5283–5293. [Google Scholar] [CrossRef] [Green Version]
  368. Li, S.; Li, C.; Song, X.; Su, B.; Mandal, B.; Prasad, B.; Gao, X.; Gao, C. Graphene quantum dots-doped thin film nanocomposite polyimide membranes with enhanced solvent resistance for solvent-resistant nanofiltration. ACS Appl. Mater. Interfaces 2019, 11, 6527–6540. [Google Scholar] [CrossRef] [PubMed]
  369. Xu, Z.; Wu, T.; Shi, J.; Teng, K.; Wang, W.; Ma, M.; Li, J.; Qian, X.; Li, C.; Fan, J. Photocatalytic antifouling PVDF ultrafiltration membranes based on synergy of graphene oxide and TiO2 for water treatment. J. Membr. Sci. 2016, 520, 281–293. [Google Scholar] [CrossRef]
  370. Hester, J.F.; Banerjee, P.; Mayes, A.M. Preparation of Protein-Resistant Surfaces on Poly(vinylidene fluoride) Membranes via Surface Segregation. Macromolecules 1999, 32, 1643–1650. [Google Scholar] [CrossRef]
  371. Amirilargani, M.; Sabetghadam, A.; Mohammadi, T. Polyethersulfone/polyacrylonitrile blend ultrafiltration membranes with different molecular weight of polyethylene glycol: Preparation, morphology and antifouling properties. Polym. Adv. Technol. 2012, 23, 398–407. [Google Scholar] [CrossRef]
  372. Zhu, L.-J.; Song, H.-M.; Wang, G.; Zeng, Z.-X.; Zhao, C.-T.; Xue, Q.-J.; Guo, X.-P. Microstructures and performances of pegylated polysulfone membranes from an in situ synthesized solution via vapor induced phase separation approach. J. Colloid Interface Sci. 2018, 515, 152–159. [Google Scholar] [CrossRef] [PubMed]
  373. Hsu, C.-H.; Venault, A.; Huang, Y.-T.; Wu, B.-W.; Chou, C.-J.; Ishihara, K.; Chang, Y. Toward Antibiofouling PVDF Membranes. Langmuir 2019, 35, 6782–6792. [Google Scholar] [CrossRef]
  374. Venault, A.; Aini, H.N.; Galeta, T.A.; Chang, Y. Using the Dimethyl Sulfoxide Green Solvent for the Making of Antifouling PEGylated Membranes by the Vapor-Induced Phase Separation Process. J. Membr. Sci. Lett. 2022, 2, 100025. [Google Scholar] [CrossRef]
  375. Tanis-Kanbur, M.B.; Tamilselvam, N.R.; Chew, J.W. Membrane fouling mechanisms by BSA in aqueous-organic solvent mixtures. J. Ind. Eng. Chem. 2022, 108, 389–399. [Google Scholar] [CrossRef]
  376. Aramwit, P. Introduction to biomaterials for wound healing. In Wound Healing Biomaterials; Elsevier: Amsterdam, The Netherlands, 2016; pp. 3–38. [Google Scholar]
  377. Karimi, M.; Bahrami, S.; Ravari, S.B.; Zangabad, P.S.; Mirshekari, H.; Bozorgomid, M.; Shahreza, S.; Sori, M.; Hamblin, M.R. Albumin nanostructures as advanced drug delivery systems. Expert Opin. Drug Deliv. 2016, 13, 1609–1623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  378. Zhao, Y.-F.; Zhu, L.-P.; Jiang, J.-H.; Yi, Z.; Zhu, B.-K.; Xu, Y.-Y. Enhancing the antifouling and antimicrobial properties of poly (ether sulfone) membranes by surface quaternization from a reactive poly (ether sulfone) based copolymer additive. Ind. Eng. Chem. Res. 2014, 53, 13952–13962. [Google Scholar] [CrossRef]
  379. Venault, A.; Chiang, C.-H.; Chang, H.-Y.; Hung, W.-S.; Chang, Y. Graphene oxide/PVDF VIPS membranes for switchable, versatile and gravity-driven separation of oil and water. J. Membr. Sci. 2018, 565, 131–144. [Google Scholar] [CrossRef]
  380. Zhang, C.; Xia, Y.; Zhang, H.; Zacharia, N.S. Surface functionalization for a nontextured liquid-infused surface with enhanced lifetime. ACS Appl. Mater. Interfaces 2018, 10, 5892–5901. [Google Scholar] [CrossRef]
  381. Guo, Y.; Liu, C.; Xu, W.; Liu, G.; Xiao, K.; Zhao, H.-Z. Interpenetrating network nanoarchitectonics of antifouling Poly(vinylidene fluoride) membranes for oil–water separation. RSC Adv. 2021, 11, 31865–31876. [Google Scholar] [CrossRef]
  382. Bose, J.; Dasgupta, J.; Adhikari, U.; Sikder, J. Tuning permeation characteristics of cellulose acetate membrane embedded with raw and amine-functionalized silicon carbide nanoparticle for oil-water separation. J. Water Process Eng. 2021, 41, 102019. [Google Scholar] [CrossRef]
  383. Kazemi, F.; Jafarzadeh, Y.; Masoumi, S.; Rostamizadeh, M. Oil-in-water emulsion separation by PVC membranes embedded with GO-ZnO nanoparticles. J. Environ. Chem. Eng. 2021, 9, 104992. [Google Scholar] [CrossRef]
  384. De Guzman, M.R.; Andra, C.K.A.; Ang, M.B.M.Y.; Dizon, G.V.C.; Caparanga, A.R.; Huang, S.-H.; Lee, K.-R. Increased performance and antifouling of mixed-matrix membranes of cellulose acetate with hydrophilic nanoparticles of polydopamine-sulfobetaine methacrylate for oil-water separation. J. Membr. Sci. 2021, 620, 118881. [Google Scholar] [CrossRef]
  385. Ishihara, K. Blood-Compatible Surfaces with Phosphorylcholine-Based Polymers for Cardiovascular Medical Devices. Langmuir 2019, 35, 1778–1787. [Google Scholar] [CrossRef]
  386. Liu, P.; Chen, Q.; Yuan, B.; Chen, M.; Wu, S.; Lin, S.; Shen, J. Facile surface modification of silicone rubber with zwitterionic polymers for improving blood compatibility. Mater. Sci. Eng. C 2013, 33, 3865–3874. [Google Scholar] [CrossRef]
  387. Xiang, T.; Lu, T.; Xie, Y.; Zhao, W.-F.; Sun, S.-D.; Zhao, C.-S. Zwitterionic polymer functionalization of polysulfone membrane with improved antifouling property and blood compatibility by combination of ATRP and click chemistry. Acta Biomater. 2016, 40, 162–171. [Google Scholar] [CrossRef] [PubMed]
  388. Chen, R.; Mao, L.; Matindi, C.N.; Liu, G.; He, J.; Cui, Z.; Ma, X.; Fang, K.; Wu, B.; Mamba, B.B. Tailoring the micro-structure of PVC/SMA-g-PEG blend ultrafiltration membrane with simultaneously enhanced hydrophilicity and toughness by in situ reaction-controlled phase inversion. J. Membr. Sci. 2022, 653, 120545. [Google Scholar] [CrossRef]
  389. Elakkiya, S.; Arthanareeswaran, G. Evaluation of membrane tailored with biocompatible halloysite–polyaniline nanomaterial for efficient removal of carcinogenic disinfection by–products precursor from water. Environ. Res. 2022, 204, 112408. [Google Scholar] [CrossRef] [PubMed]
  390. Zhao, X.; Lan, Y.; Yang, K.; Wang, R.; Cheng, L.; Gao, C. Antifouling modification of PVDF membranes via in situ mixed-charge copolymerization and TiO2 mineralization. Appl. Surf. Sci. 2020, 525, 146564. [Google Scholar] [CrossRef]
  391. Zhang, L.; Tang, Y.; Jiang, X.; Yu, L.; Wang, C. Highly dual antifouling and antibacterial ultrafiltration membranes modified with silane coupling agent and capsaicin-mimic moieties. Polymers 2020, 12, 412. [Google Scholar] [CrossRef] [Green Version]
  392. Cui, Z.; Tang, X.; Li, W.; Liu, H.; Zhang, J.; Wang, H.; Li, J. EVOH in situ fibrillation and its effect of strengthening, toughening and hydrophilic modification on PVDF hollow fiber microfiltration membrane via TIPS process. J. Mater. Sci. 2019, 54, 5971–5987. [Google Scholar] [CrossRef]
  393. Zhu, L.-J.; Song, H.-M.; Wang, G.; Zeng, Z.-X.; Xue, Q.-J. Dual stimuli-responsive polysulfone membranes with interconnected networks by a vapor-liquid induced phase separation strategy. J. Colloid Interface Sci. 2018, 531, 585–592. [Google Scholar] [CrossRef]
  394. Liu, D.; Zheng, J.; Wang, X.; Lu, X.; Zhu, J.; He, C. Low-fouling PES membranes fabricated via in situ copolymerization mediated surface zwitterionicalization. New J. Chem. 2018, 42, 2248–2259. [Google Scholar] [CrossRef]
  395. Zhang, L.; Xu, J.; Tang, Y.; Hou, J.; Yu, L.; Gao, C. A novel long-lasting antifouling membrane modified with bifunctional capsaicin-mimic moieties via in situ polymerization for efficient water purification. J. Mater. Chem. A 2016, 4, 10352–10362. [Google Scholar] [CrossRef]
  396. Zhu, J.; Zhang, Y.; Tian, M.; Liu, J. Fabrication of a mixed matrix membrane with in situ synthesized quaternized polyethylenimine nanoparticles for dye purification and reuse. ACS Sustain. Chem. Eng. 2015, 3, 690–701. [Google Scholar] [CrossRef]
  397. Wang, J.J.; Wu, M.B.; Xiang, T.; Wang, R.; Sun, S.D.; Zhao, C.S. Antifouling and blood-compatible poly (ether sulfone) membranes modified by zwitterionic copolymers via In situ crosslinked copolymerization. J. Appl. Polym. Sci. 2015, 132. [Google Scholar]
  398. Tao, M.; Liu, F.; Xue, L. Persistently hydrophilic microporous membranes based on in situ cross-linking. J. Membr. Sci. 2015, 474, 224–232. [Google Scholar] [CrossRef]
  399. Li, X.; Pang, R.; Li, J.; Sun, X.; Shen, J.; Han, W.; Wang, L. In situ formation of Ag nanoparticles in PVDF ultrafiltration membrane to mitigate organic and bacterial fouling. Desalination 2013, 324, 48–56. [Google Scholar] [CrossRef]
  400. Zhang, P.-Y.; Xu, Z.-L.; Yang, H.; Wei, Y.-M.; Wu, W.-Z. Fabrication and characterization of PVDF membranes via an in situ free radical polymerization method. Chem. Eng. Sci. 2013, 97, 296–308. [Google Scholar] [CrossRef]
  401. Zhao, X.; Zhang, W.; Chen, S.; Zhang, J.; Wang, X. Hydrophilicity and crystallization behavior of PVDF/PMMA/TiO2 (SiO2) composites prepared by in situ polymerization. J. Polym. Res. 2012, 19, 9862. [Google Scholar] [CrossRef]
  402. Tao, M.; Liu, F.; Xue, L. Hydrophilic Poly(vinylidene fluoride)(PVDF) membrane by in situ polymerisation of 2-hydroxyethyl methacrylate (HEMA) and micro-phase separation. J. Mater. Chem. 2012, 22, 9131–9137. [Google Scholar] [CrossRef]
  403. Li, H.; Shi, W.; Zhang, Y.; Zhou, R.; Zhang, H. Preparation of hydrophilic PVDF/PPTA blend membranes by in situ polycondensation and its application in the treatment of landfill leachate. Appl. Surf. Sci. 2015, 346, 134–146. [Google Scholar] [CrossRef]
  404. Zhu, K.; Mu, Y.; Zhang, M.; Liu, Y.; Na, R.; Xu, W.; Wang, G. Mixed matrix membranes decorated with in situ self-assembled polymeric nanoparticles driven by electrostatic interaction. J. Mater. Chem. A 2018, 6, 7859–7870. [Google Scholar] [CrossRef]
  405. Xu, H.-P.; Yu, Y.-H.; Lang, W.-Z.; Yan, X.; Guo, Y.-J. Hydrophilic modification of polyvinyl chloride hollow fiber membranes by silica with a weak in situ sol–gel method. RSC Adv. 2015, 5, 13733–13742. [Google Scholar] [CrossRef]
  406. Patala, R.; Mahlangu, O.T.; Nyoni, H.; Mamba, B.B.; Kuvarega, A.T. In Situ Generation of Fouling Resistant Ag/Pd Modified PES Membranes for Treatment of Pharmaceutical Wastewater. Membranes 2022, 12, 762. [Google Scholar] [CrossRef]
  407. Huang, M.-h.; Li, Y.-m.; Gu, G.-w. Chemical composition of organic matters in domestic wastewater. Desalination 2010, 262, 36–42. [Google Scholar] [CrossRef]
  408. Wulan, D.R.; Hamidah, U.; Komarulzaman, A.; Rosmalina, R.T.; Sintawardani, N. Domestic wastewater in Indonesia: Generation, characteristics and treatment. Environ. Sci. Pollut. Res. 2022, 29, 32397–32414. [Google Scholar]
Figure 1. Schematic of phase-inversion processes for the formation of flat-sheet antifouling porous membranes via blending modification.
Figure 1. Schematic of phase-inversion processes for the formation of flat-sheet antifouling porous membranes via blending modification.
Membranes 13 00058 g001
Figure 2. Literature survey of antifouling membranes prepared via blending modification from 2012–2022 obtained from the Web of Science.
Figure 2. Literature survey of antifouling membranes prepared via blending modification from 2012–2022 obtained from the Web of Science.
Membranes 13 00058 g002
Figure 3. Schematic illustration of the different groups of additives used to modify polymeric membranes by blending.
Figure 3. Schematic illustration of the different groups of additives used to modify polymeric membranes by blending.
Membranes 13 00058 g003
Figure 4. Schematic effect of the antifouling additive on the position of the demixing gap. Red arrows highlight the shift of the binodal line towards the solvent/polymer axis.
Figure 4. Schematic effect of the antifouling additive on the position of the demixing gap. Red arrows highlight the shift of the binodal line towards the solvent/polymer axis.
Membranes 13 00058 g004
Figure 5. Influence of the addition of 1 wt% PS30-b-PEGMA68 copolymer to a casting solution containing PVDF (NMP as solvent) on the cross-sectional morphology of LIPS membranes (Tdissolution: 40 °C; Solvent content fixed: 75 wt%; non-solvent: DI water; casting thickness: 300 µm). Left image partially reproduced with permission from [357]. Copyright 2014 Elsevier.
Figure 5. Influence of the addition of 1 wt% PS30-b-PEGMA68 copolymer to a casting solution containing PVDF (NMP as solvent) on the cross-sectional morphology of LIPS membranes (Tdissolution: 40 °C; Solvent content fixed: 75 wt%; non-solvent: DI water; casting thickness: 300 µm). Left image partially reproduced with permission from [357]. Copyright 2014 Elsevier.
Membranes 13 00058 g005
Figure 6. Influence of the addition of 1 wt% PS60-b-PEGMA108 copolymer to a casting solution containing 20 wt% PVDF (NMP as solvent) on the surface morphology of VIPS membranes (Tdissolution: 32 °C; RH: 70%; Tchamber: 25 °C; texposure: 20 min; casting thickness: 300 µm).
Figure 6. Influence of the addition of 1 wt% PS60-b-PEGMA108 copolymer to a casting solution containing 20 wt% PVDF (NMP as solvent) on the surface morphology of VIPS membranes (Tdissolution: 32 °C; RH: 70%; Tchamber: 25 °C; texposure: 20 min; casting thickness: 300 µm).
Membranes 13 00058 g006
Table 1. Polymer/organic-based antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Table 1. Polymer/organic-based antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Polymer/Organic-Based
Additives
MaterialsRef.
Hydrophilic or amphiphilicPoly(ethylene glycol) and its derivatives[1,41,42,86,87,88,89,90,91,92,93]
Poly(vinylpyrrolidone)[43,94,95,96,97]
Cellulose nanocrystals[98,99,100,101,102,103]
Poly(vinyl alcohol)[104,105,106,107]
Poly(acrylic acid)[44,108,109]
Polydopamine[110,111,112,113,114]
Zwitterion/amphiphilic zwitterion[115,116,117,118,119,120,121,122,123,124]
Amphiphilic copolymer[95,125,126,127,128,129,130]
Nature-derived biopolymersVanillin[131,132,133]
Chitosan[134,135,136,137,138]
Chitin[139,140]
Lignin[141,142]
Lignocellulose[143]
Caramel[144]
Capsaicin[145,146]
Acacia gum[147]
Isocyanate[148,149]
Organic acids[45,150,151]
p-aramid[152]
Table 2. Metal-based antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Table 2. Metal-based antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Metal-Based
Additives
MaterialsRef.
MetallicAg and bio-derived Ag[172,173,174,175,176]
Au[177]
Pd[178]
Biphasic metalsZrO2[179,180,181,182]
ZnO[183,184,185,186,187,188]
TiO2[38,58,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203]
SiO2[204,205,206,207,208,209,210,211,212,213,214,215,216]
CuO[186,187,188,217,218]
Fe2O3, Fe3O4[14,51,205,219,220,221,222,223,224,225,226,227]
Al2O3[55,228,229]
MXenes[230]
Bimetallic oxidesFe-Mn oxide[231]
Table 3. Ceramic-based antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Table 3. Ceramic-based antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Ceramic-Based
Additives
MaterialsRef.
NanoclayBoehmite[241,242,243]
Montmorillonite[244,245,246,247]
Bentonite[63,248,249]
Attapulgite[250]
Hydroxyappatite[251,252,253,254]
Goethite[151]
Cloisite 15A and 30B[60,255,256]
Pyrochlores[255]
SBA-15[257,258,259,260,261,262]
Halloysite[158,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279]
Layered double hydroxide[64,65,280,281]
Table 4. Carbon allotropes and porous nanomaterial as antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Table 4. Carbon allotropes and porous nanomaterial as antifouling additives used in the last 10 years for the blending modification of polymeric membranes.
Carbon Allotropes and
Porous Nanomaterials
MaterialsRef.
Carbon allotropesCarbon quantum dot[286,287]
Nanodiamond[72,288,289,290,291,292]
Carbon nanotubes[293,294,295,296,297,298,299,300,301,302,303]
Graphene and its derivatives[157,304,305,306,307,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323,324,325,326,327,328]
Porous nanomaterialsMetal-organic framework[68,70,329,330,331,332,333,334,335,336,337,338,339,340]
Zeolites/ZIF[73,341,342]
Cyclodextrin[343]
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

Geleta, T.A.; Maggay, I.V.; Chang, Y.; Venault, A. Recent Advances on the Fabrication of Antifouling Phase-Inversion Membranes by Physical Blending Modification Method. Membranes 2023, 13, 58. https://doi.org/10.3390/membranes13010058

AMA Style

Geleta TA, Maggay IV, Chang Y, Venault A. Recent Advances on the Fabrication of Antifouling Phase-Inversion Membranes by Physical Blending Modification Method. Membranes. 2023; 13(1):58. https://doi.org/10.3390/membranes13010058

Chicago/Turabian Style

Geleta, Tesfaye Abebe, Irish Valerie Maggay, Yung Chang, and Antoine Venault. 2023. "Recent Advances on the Fabrication of Antifouling Phase-Inversion Membranes by Physical Blending Modification Method" Membranes 13, no. 1: 58. https://doi.org/10.3390/membranes13010058

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