Next Article in Journal
New Poly(imide)s Bearing Alkyl Side-Chains: A Study on the Impact of Size and Shape of Lateral Groups on Thermal, Mechanical, and Gas Transport Properties
Next Article in Special Issue
Hydrophilic Dual Layer Hollow Fiber Membranes for Ultrafiltration
Previous Article in Journal
Computational Fluid Dynamics (CFD) Modeling and Simulation of Flow Regulatory Mechanism in Artificial Kidney Using Finite Element Method
Previous Article in Special Issue
In-Situ Combination of Bipolar Membrane Electrodialysis with Monovalent Selective Anion-Exchange Membrane for the Valorization of Mixed Salts into Relatively High-Purity Monoprotic and Diprotic Acids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Developments in Nanomaterials-Modified Membranes for Improved Membrane Distillation Performance

by
Saikat Sinha Ray
1,*,†,
Harshdeep Singh Bakshi
1,2,†,
Raghav Dangayach
1,2,†,
Randeep Singh
1,3,†,
Chinmoy Kanti Deb
2,†,
Mahesh Ganesapillai
2,*,
Shiao-Shing Chen
1,* and
Mihir Kumar Purkait
3
1
Institute of Environmental Engineering and Management, National Taipei University of Technology, Taipei City 106, Taiwan
2
School of Chemical Engineering, Vellore Institute of Technology (VIT), Vellore 632014, India
3
Department of Chemical Engineering, Indian Institute of Technology, Guwahati 781039, India
*
Authors to whom correspondence should be addressed.
The authors have contributed equally.
Submission received: 9 June 2020 / Revised: 1 July 2020 / Accepted: 2 July 2020 / Published: 3 July 2020
(This article belongs to the Special Issue Membranes: 10th Anniversary)

Abstract

:
Membrane distillation (MD) is a thermally induced membrane separation process that utilizes vapor pressure variance to permeate the more volatile constituent, typically water as vapor, across a hydrophobic membrane and rejects the less volatile components of the feed. Permeate flux decline, membrane fouling, and wetting are some serious challenges faced in MD operations. Thus, in recent years, various studies have been carried out on the modification of these MD membranes by incorporating nanomaterials to overcome these challenges and significantly improve the performance of these membranes. This review provides a comprehensive evaluation of the incorporation of new generation nanomaterials such as quantum dots, metalloids and metal oxide-based nanoparticles, metal organic frameworks (MOFs), and carbon-based nanomaterials in the MD membrane. The desired characteristics of the membrane for MD operations, such as a higher liquid entry pressure (LEPw), permeability, porosity, hydrophobicity, chemical stability, thermal conductivity, and mechanical strength, have been thoroughly discussed. Additionally, methodologies adopted for the incorporation of nanomaterials in these membranes, including surface grafting, plasma polymerization, interfacial polymerization, dip coating, and the efficacy of these modified membranes in various MD operations along with their applications are addressed. Further, the current challenges in modifying MD membranes using nanomaterials along with prominent future aspects have been systematically elaborated.

Graphical Abstract

1. Introduction

The changing environment and depleting sources of water lead to increased demand for clean and drinking water for domestic, agricultural, and industrial use. According to the National Ocean Service-US, around 97 percent of Earth’s water is in the ocean, but due to its ultra-high salinity, this water is unsuitable for direct use. In this context, membrane distillation (MD) is highly effective and efficient to desalinate ultra-high saline waters. Hence, the MD separation method offers various combinations to be used in desalination, wastewater treatment, pharmaceutical, food, and biomedical applications for different applications. The lower membrane mechanical strength, temperature, and pressure requirements distinguish MD from other membrane separation processes [1]. In addition, MD can be easily combined with various processes and renewable or alternative sources of energy, making it not only energy-efficient but also effective in separation. These properties make MD a desired membrane separation process for large-scale applications. As earlier mentioned, MD is a thermally induced membrane separation process widely utilized for desalination and wastewater treatment [1,2,3,4]. In addition, MD separation is a non-isothermal process in which a hot feed side and a cold permeate side are separated by a thin hydrophobic membrane. The temperature difference between feed and permeate results in a state of thermodynamic disequilibrium, which leads to a vapor pressure difference. Under the presence of this driving force, vapor molecules are transported from the feed side to the permeate side and condense as a pure distillate [5]. Membrane distillation can be used in different arrangements, such as air gap membrane distillation (AGMD), direct contact membrane distillation (DCMD), sweeping gas membrane distillation (SGMD), and vacuum membrane distillation (VMD) [4,5]. These different types of MD processes carry their own advantages and can be used accordingly in relation to the application type and demand. MD membranes, as mentioned earlier, face numerous challenges such as membrane fouling, surface wetting, and low permeate flux resulting in comprehensive work to address these challenges by integrating nanoparticles (NPs) in these MD membranes. These include quantum dots, metal-based NPs (titania (TiO2), silver (Ag)), silica-based NPs, metal organic frameworks (MOF), and carbon nanotubes (CNTs) that are used as either surface modifiers or blended in the polymeric membrane matrix. Nanomaterials are often used to improve membrane flux and salt rejection, enhance surface hydrophobicity, mitigate membrane fouling, and enhance the physical and chemical characteristics of the membrane to ensure superior functionality [6]. Therefore, research has focused on the engineering of superhydrophobic, anti-wetting surfaces, including the development of MD membranes with new generation nanomaterials for enhanced MD performance.

1.1. Timeline of MD Membranes

Membrane distillation has emerged as an effectual process for desalinating salty water under low temperature and pressure conditions [7], unlike other methods, such as nanofiltration and reverse osmosis, that demand high energy, hence substantial capital investment [8,9]. Despite advancements, the popularity of the MD process is declining due to the lack of availability of appropriate membranes due to various challenges in the MD operation (Figure 1) that makes it difficult to come up with an ideal MD membrane separation process compared with other membrane separation processes. The first MD patent came in the year 1963 [10] and the first research publication in 1967. Later, in the 1960s, the first set for asymmetric membranes was developed for membrane desalination applications, which was revolutionary in the field of separation technology for the next five decades [1,2,3,11]. However, there is still scope for advancement and development in the field of MD membranes in order to propose the successful application of the MD process on an industrial scale. The most important aspect of using the MD process for large-scale applications is the integration of MD with other processes and utilizing a cheaper source of energy to make it a cost-effective, eco-friendly, and efficient process. The operational cost of the membrane distillation system much depends on the source of energy, where conventional heat sources such as burning coal, petroleum, or electricity make the MD process very expensive (10.8 USD/m3 for AGMD). The operational cost of MD processes can be significantly reduced (2.68 USD/m3 for AGDM) if an alternative, cheaper heat source, such as waste heat and solar heat is utilized, making its operational cost comparable to conventional desalination technologies [11].
Remarkably, MD operations also have a higher rejection of micro-pollutants when compared with both nanofiltration (NF) and reverse osmosis (RO) processes [12]. Although MD operations can be energy-efficient and have a higher rejection percentage, membrane fouling and wetting are major concerns affecting the efficiency of the process [13,14]. Extensive research on the modification of polymeric membranes by the incorporation of nanoparticles (NPs) is being conducted in order to mitigate these concerns. NPs are ultrafine particles with at least one dimension less than 100 nm. These ultrafine particles attribute very different physicochemical characteristics compared with their bulk counterparts, despite being made of the same materials, and hence the mean particle size and size distribution play a vital role in the applications of these NPs [15,16].
A careful survey was executed to show the progressive development in the field of MD membranes in the last decade, as shown in Figure 2, in terms of research publications related to “membrane distillation” and “nanomaterials”. The database was taken from an advanced scholarly search system based on Scopus. Therefore, it can be concluded that nanomaterial-based MD membranes have gained a lot of attention in recent desalination research and development.
Although there are many advantages to MD, the progress of MD and MD membranes on a commercial scale has been literally slow due to various challenges faced, as pointed out in Figure 1. Although there are many advantages to MD, the progress of MD and MD membranes on a commercial scale has been literally slow. The progress in the first three decades was minimal, and most of the studies were basic in nature, focusing on the understanding of the MD process and its different configurations. Later in the emerging phase, many research and development sectors came along to tackle the serious issues related to the wetting and fouling of MD membranes.

1.2. New Generation Nanomaterials-based MD Membrane

Over the past decade, nanomaterials such as metalloid and metal NPs, MOFs, CNTs, QDs, and graphene and its derivatives have attracted a lot of attention of researchers and thus have emerged as pioneer candidates for potential materials to be used for MD membrane modification. These nanomaterials have been used for various purposes to mitigate membrane fouling and pore wetting as they improve the physicochemical properties of the membrane [17].
Metallic NPs possess a hierarchal structure with multilevel roughness, thus improving the hydrophobic character of the membrane [18]. Metallic NPs are often used to generate an uneven surface to the membrane surface. This uneven surface naturally leads to the formation of air pockets, which allow for improper contact between the water and the membrane surface, resulting in the improvement of the hydrophobic properties of the membrane [19,20]. Further, some of the NPs, such as Ag, TiO2, or SiO2, possess antibacterial characteristics due to their ability to react with the thiol (-SH) group present in the cell wall of microorganisms which prevents biofilm formation [21]. MOFs, on the other hand, are one of the most common fillers that are used in mixed matrix membranes. Additionally, the unique physical and chemical properties of MOFs make them the most commonly used fillers in mixed matrix membranes. The organic ligands of MOFs show strong attraction towards the polymeric chains, describing excellent filler properties [22]. Further, MOFs-based membranes exhibit tailorable pore structure and size, which makes them application-specific in terms of selectivity and permeation [23]. They possess an ordered arrangement with high pore volume and specific surface area, which modifies the membrane structure resulting in improved transmembrane flux [24]. Carbon-based nanomaterials are a superior class of nanomaterials possessing novel characteristics like a large surface area to volume ratio, high mechanical strength, and tendency to reduce fouling propensity, which makes them attractive materials to fabricate multifunctional composite membranes [25]. Quantum dots have recently attracted a lot of attention as an MD membrane modifier due to their low cost of production and highly desirable physicochemical properties [26]. Figure 3 shows the efficacy of new generation nanomaterials on the MD membrane.
New generation NPs used in MD studies have been critically reviewed in this investigation with a focus on the enhancement of the anti-fouling and superhydrophobic properties of the membrane. Modified membranes have been evaluated using characteristics such as pure water flux, salt rejection, pore size and porosity, water contact angle, liquid entry pressure (LEPw) value, and long-term operation. Further, we have thoroughly discussed membrane modification methods, which usually are neglected while discussing about MD membrane fabrication.
The main goal of this review was to give a comprehensive outlook on the various state-of-the-art membranes fabricated for MD application using novel materials such as metallic NPs, CNTs, graphene and its derivatives, and MOFs. Even though literature focused on the improvement of membrane characteristics using an NP addition has been published before, our study aims at giving detailed scientific discussions on the mechanisms which govern the improved properties of nanoparticle-modified membranes [27,28,29,30]. This review article focuses on the effectiveness and strategies used to design membrane nanostructures for improved performance in MD. In addition, this review article compares the advantages and disadvantages of emerging NPs and nanomaterial-based membranes. Finally, the research gap and problems associated with nanomaterial-based membranes in previous MD investigations have been addressed with potential solutions to fabricate these membranes for real-world applications.

2. Fundamentals of MD Membrane

Lippman’s non-isothermal liquid separation was first introduced in 1907 [31]. Subsequently, Aubert carried out a detailed investigation into this process [32]. This phenomenon of non-isothermal separation was called thermo-osmosis due to the nature of the liquid separation. As mentioned above, the separation of liquids takes place in the form of vapor in MD. Therefore, theoretically, any type of membrane, dense or porous, charged or neutral, or hydrophilic or hydrophobic, in nature may be used. Porous and hydrophobic membranes, however, are commonly used for better results, as described later in this section. In addition, the material used for the synthesis of MD membranes is selected on the basis of the properties to be imparted to the membranes [33,34,35]. Therefore, this section gives details about the various developments and advancements of the MD membrane. In later sections, commonly used membrane modifications via the incorporation of nanomaterials for the development of specific membranes for the MD process are discussed.

2.1. Membrane Materials and Structure

The material used for fabricating the membrane is of extreme importance when designing a typical MD system because it affects both the heat and the transfer of mass across the membrane, which influences the long-term performance of the membrane [36]. Materials like polytetrafluoroethylene (PTFE), polyvinylidene fluoride (PVDF), and polypropylene (PP) are the most widely used polymers for MD membrane fabrication due to their intrinsic hydrophobic nature and good processability. A more hydrophobic surface would prevent the penetration of the liquid phase into the membrane pores, due to the surface tension forces, while allowing the vapor molecules to pass through the porous surface. Membrane porosity and thickness also vary along with the pore size, which could lead to a different conclusion on the overall effect on MD performance. It is believed that the properties such as the thickness, tortuosity, pore size, and porosity dictate the resistance to mass transfer in the MD process [37]. Figure 4 shows the basic membrane features, along with the desired range that affects the performance of MD.
Apart from the choice of the polymer, the concentration of the polymer in the dope solution can affect the resultant morphology and performance of the membrane. In a recent study, the effects of the dope concentration of polyethersulfone (PES) on the morphology and average pore size of the achieved membrane were studied. It was observed that with an increase in the concentration of PES in the spinning dope solution, the average pore size of the PES hollow fiber substrate decreased [38,39]. Therefore, a key area of scientific research focuses on the design, modification, and manufacture of hydrophobic membranes for MD processes. Nanomaterials were widely used as additives to enhance the structure of the membranes and achieve desired characteristics [40]. Some primary nanomaterial properties include superior chemical and thermal stability, high surface area to volume ratio, and exceptional mechanical resistance. Consequently, nanomaterial doping can lead to a major improvement of the membrane structure as it increases porosity, modifies membrane morphology, and also enhances surface hydrophobicity [41]. SiO2-NPs doping in PVDF led to the formation of an asymmetric arrangement with a porous sponge layer as well as a finger-like macro void layer, which resulted in improved VMD performance [42]. The incorporation of superhydrophobic alumina NPs in nanofibrous PVDF membranes improved the liquid entry pressure (LEPw) and water contact angle of the membrane structure, which helped in the treatment of highly concentrated brines in an AGMD operation [43]. Water vapor transport improved in graphene oxide-modified PTFE membranes due to the selective sorption of water onto the epoxy, hydroxyl, and carboxyl groups in GO [44]. A carbon nanotube-immobilized PP membrane showed a greater mass transfer coefficient than the pristine PP membrane, which was enabled in the separation of pure water from saline water [45]. Thus, different kinds of nanomaterials can be used to fabricate multifunctional membranes possessing novel properties, which will improve MD performance.

2.2. MD Membrane Perquisites

An MD system operation is heavily dependent upon the membrane’s intrinsic properties. It is essential to have a perfect blend of structural and physicochemical characteristics to achieve optimum performance. The characterization of different properties like surface morphology, membrane thickness, pore size, and geometry sheds light on the physical characteristics of the material used for manufacturing. On the other hand, contact angle, sliding angle, surface roughness, and LEPw measurements provide insight into the membrane’s hydrophobicity. The membrane should have low mass transfer resistance and thermal conductivity, while also maintaining high thermal and chemical stability for the performance required [46]. Table 1 provides an outline of the desired characteristics required in membranes that are currently used for the MD application.

2.3. MD Membrane Modification

Membranes for MD operations are generally manufactured using phase inversion (solution casting) and electrospinning methods, which are subsequently modified with dip coating, graft polymerization, and interfacial polymerization [56]. The phase inversion technique involves defusing/de-mixing, where under optimal conditions, a homogeneous polymeric solution is transformed into a membrane. This technique can be further classified according to the methods used for transformation, viz., immersion precipitation, thermally induced phase separation, and precipitation by controlled evaporation [2]. Thermally induced phase separation conduces to the formation of the membrane when the parent polymeric solution is de-mixed under high-temperature conditions. In the immersion precipitation method, the membrane is developed by the interchange of the solvent between the polymeric solution and coagulation solvent where the polymeric solution is immersed. In contrast, the membrane is formed during evaporative phase inversion by the volatilization of the solvent used to prepare the initial polymeric solution [57,58].
The electrospinning technique, which is a variation of the electro-spraying process, involves a polymeric solution or melt subjection to high electrical fields, which reduce the polymeric blend surface tension [59,60,61,62]. Under this condition, the polymer is stretched and diluted, resulting in the formation of nanofibers collected on a fixed surface [63]. Different methods, including surface coating, surface grafting, plasma polymerization, interfacial polymerization, and dip coating are used to incorporate NPs in the MD membrane to enhance their properties, enabling them to overcome challenges such as fouling, wetting, fluxing, and porosity by providing them with the necessary characteristics. Figure 5 indicates the schematic diagram of various methodologies for the incorporation or doping of nanomaterials in the MD membrane.

2.3.1. Surface Grafting

Surface grafting is a chemical treatment in which the membrane material chains are activated and then grafted by chemical reaction or intense radiation. These grafted macromolecular chains form covalent bonds with the surface of the membrane. These bonds established on the MD membrane surface guarantee long-term stability and prevent surface delamination [64]. In contrast to physical modification, surface grafting forms a covalent linkage between the polymeric surface of the membrane and the molecules of the nanomaterial used as a modifier; this is generally carried out in two different ways, i.e., grafting-to and grafting-from.
In the grafting-to technique, the end-functionalized active molecules react with the surface of the membrane, while in grafting-from, free radical sites are generated by the base polymer, which reacts with the modifiers to form a modified surface. Comparing these two methods, grafting-to is more convenient to control as, in this case, the pre-synthesized reactive groups can be purified before grafting, whereas grafting-from provides an edge when it comes to controlling the thickness of the grafted layer, as in this case the modifiers can be added over time. Further, in the grafting-from method, various processes are available to activate the surface of the polymeric membrane for grafting polymerization, which includes plasma treatment, UV irradiation, ozone treatment, and gamma-ray irradiation [65]. Many recent studies have shown that surface grafting can be used to modify a hydrophilic membrane surface into a hydrophobic surface, making them suitable for MD operations. In a recent study, a hydrophobic yttria-stabilized zirconia (YSZ) hollow fiber membrane was obtained by grafting fluoroalkylsilanes (FASs) on the pristine membrane surface. The achieved modified membrane attributed a high contact angle (up to 140°), high mechanical strength, and considerable chemical and thermal stability [66]. In another study, grafting induced by direct radiation was used by El-Arnaoty et.al., for the incorporation of ZnO NPs on the polyamide membrane surface to enhance its anti-biofouling properties [67].

2.3.2. Plasma Polymerization

Plasma polymerization involves electrically induced atomic disintegration of monomers that give rise to different active particles. The generated activated particles combine with the surface to form a highly branched and crosslinked structure. This deposited layer ensures a lower mass transfer coefficient and decreases the water flux of the MD membrane. Plasma polymers show stronger adhesion to the substrate along with higher thermal stability. Plasma polymerization is a one-step method and hence provides a greater advantage over other modification techniques by cutting down several steps. Further, this incorporated layer only has a thickness of 1 to 2 µm and has very minimal impact on the porosity of the MD membrane compared with other techniques. This modification technique is widely utilized in the incorporation of metal/metal oxide nanomaterials on the membrane surface, and in a recent study, silicon dioxide particles were grown on a PVDF membrane surface to enhance its stability and reduce the flux decrease [68]. Similar to this, in various other studies, the incorporation of different metal/metal oxide nanomaterials was carried out using plasma polymerization to enhance the physicochemical properties of the membrane surface without compromising its porosity. Moreover, this method of MD membrane modification is environment-friendly [69].

2.3.3. Interfacial Polymerization

Interfacial polymerization is a step-growth technique in which polymerization takes place at the interface between two non-miscible phases providing them with specific chemical and topological properties, like anisotropic shapes, hollow structures, and alternative surface chemistry [70]. The fabrication of a polymer material by interfacial polymerization includes two major aspects, which are developing an interface between the two non-miscible phases and distributing the monomers in these two phases. These interfaces can be further categorized as liquid–liquid emulsion (L–L), liquid–solid emulsion (L–S), and liquid in liquid emulsion (L–in–L). The factors affecting interfacial polymerization include temperature, humidity, and purity of the reactants. This technique has been used for the surface enhancement of membranes by incorporating carbon, graphene, metal/metal oxide nanomaterials, and their composites. In the latest study carried out by [71], graphene-oxide-TiO2 nanofillers were incorporated onto PA membranes by interfacial polymerization, providing them with higher water vapor permeance and selectivity.

2.3.4. Dip Coating

Dip coating is one of the straightest techniques available to incorporate NPs onto an MD membrane [72]. The process is carried out in three steps: (a) immersion and dwell time, (b) deposition, and (c) evaporation of the solvent. The base polymer is kept perpendicularly in the solution containing NPs until these modifiers properly settle on the surface, after which, the membrane is withdrawn out of the solution and is allowed to dry. In this same manner, commercially available polyester (PET) membranes were hydrophobically modified by PDMS-coated SiO2 NPs yielding a chemically stable, superhydrophobic membrane (dip-PET) [73]. The coating thickness, structural integrity, and pore size of the post-modification membrane are dependent on several process parameters like dipping time, the concentration of the dipping polymer, and concentration of the crosslinking solution [74].

2.4. State of the Art of Nanomaterials Doped MD Membrane

These specialized MD membranes achieved better results both in terms of water flux and the salt rejection rates compared with the pristine membrane, as mentioned earlier. Doping certain nanomaterials transforms the chemical aspect of the membrane; therefore, it helps to achieve better performance in the MD operation. A short summary of various nanomaterials incorporated in membranes for various MD applications is demonstrated in Table 2.

3. Incorporation of Nanomaterials for Enhanced Performance

Nanotechnology has played a crucial part in the development and advancement of membrane science. Nanotechnology has provided excellent materials, such as metallic NPs, carbon nanotubes, graphene, and metal organic frameworks, as membrane modifiers so as to achieve the required membrane functional and structural characteristics. These materials provided membranes with exceptional characteristics, such as selective permeability, chemical resistance, anti-fouling nature, mechanical strength, and thermal stability, resulting in better operational and functional properties of the membranes. The use of nanotechnology has contributed to the advancement of the membrane life span, which reduces the overall cost of a separation process. This makes the membrane separation processes competitive with other traditional processes and proposes their use in applications of a large scale.
In addition to that, the incorporation of NPs as membrane modifiers to improve fouling resistance and minimize wettability is focused. Few have reported the utilization of silica-based NPs and TiO2-NPs for improved membrane hydrophobicity [75,77]. Therefore, the use and effects of different nanomaterials on membrane modifications are discussed in detail in this section. The section further discusses the various applications of modified membranes. As before, MD-based studies were virtually non-existent in the 1980s, but during the early 2000s, most of the developments and advances were directed towards the growth of the MD technique. The recent research boom has been observed in MD membranes where it is directed towards commercialization, with a particular emphasis on improvements in material and structure involving the use of different types of nanomaterials. Figure 6 shows the current evolution of MD membranes involving different groups of new generation nanomaterials for efficient MD output.

3.1. Metalloid and Metal Oxides Based Nanoparticles

Metalloids and metal oxide NPs have the potential to produce synergistic effects when combined with different types of materials and hence are extensively used in enhancing the properties of MD membranes. In recent studies, various oxides-based NPs like SiO2, TiO2, ZnO, and Al2O3 were used to enhance the membrane flux and wetting resistance for MD operations [53,77]. Among these, TiO2 NPs have gained a lot more consideration due to their high availability, and excellent physicochemical, antibacterial, and anti-fouling properties [86,87]. TiO2 NPs are added to prevent fouling by rendering hydrophilicity to these hydrophobic MD membranes, reducing the surface interaction of these organic foulants. TiO2 NPs can be incorporated either by mixing them in bulk with the membrane substrate or by coating them over the surface. A superhydrophobic MD membrane was developed by coating a TiO2 nanoparticle on the surface of a polyvinylidene fluoride (PVDF) membrane, followed by fluorosilanization of the surface. The generated FTCS-TiO2-PVDF membrane had a significantly higher LEPw value (190 kPa) and contact angle (163 ± 3°). The incorporation of TiO2 NPs provided the surface with a hierarchical arrangement and -OH functional groups, ensuring uniform functionalization of perfluorododecyltrichlorosilane (FTCS). As illustrated in Figure 7, the incorporation of FTCS on the rough surface hydrolyzed the hydrophilic end on the TiO2 incorporated PVDF surface, exposing the hydrophobic fluorinated carbon chain [77]. In a similar manner, the FAS-PVDF-SiO2 nanofiber membrane was prepared, by the electrospinning of the PVDF-SiO2 solution, followed by fluorosilanization, by immersing the obtained PVDF-SiO2 nanofiber membrane into a fluoroalkylsilane (FAS)-containing solution. In this case, the achieved superhydrophobic surface featured an LEPw value of 195 kPa and a contact angle close to 160.5° [88]. In another study, a PVDF/SiO2 NPs composite membrane was developed by the phase inversion technique for a VMD operation. The incorporation of these superhydrophobic SiO2 NPs enhanced the hydrophobicity of the membrane. Though the composite membranes could not feature a superhydrophobic surface, they all showed higher hydrophobicity than the pristine PVDF membrane, with a maximum contact angle of 94°. The generated nano-composite membrane could achieve a maximum flux value of approximately 3 L m−2 h−1, 99.98% salt rejection rate, and an LEPw value greater than 3 bar, making it suitable for VMD operations for seawater desalination [42].
The incorporation of metal oxide NPs on a polymeric MD membrane can significantly increase its surface roughness. In a study put forward by Dong Ma et al., the surface roughness of the developed FAS-PVDF-SiO2 membrane significantly raised from 92 to 218 nm as the weight percentage of the SiO2 nanomaterial in the dope solution increased from 0% to 8% [88]. This increase in surface roughness reduces the effective solid–liquid interface area along with an increase in the presence of air pockets, formed when the two layers collect a certain amount of vapor/air mixture. In the presence of external hydraulic pressure, these air pockets still keep deforming. They ensure that the water droplets move quickly and hence decrease the water sliding angles of the membrane. Further, the presence of pocket-like cells creates a resistance to heat transfer, which helps maintain a high temperature in the feed side, ensuring a higher vapor pressure gradient, thus increasing the water flux of the MD membrane [89,90]. Figure 8 demonstrates the effects of air pockets that result in superhydrophobicity and anti-wetting properties.

3.2. Carbon Materials

Carbon-based nanomaterials such as carbon nanotubes (CNTs), graphene, and graphene oxide are extensively utilized in membrane separation techniques due to their unique characteristics that enhance membrane permeability and selectivity [91]. Such nanomaterials allow water molecules to be transported swiftly and impart an anti-fouling character to the membrane, making them a successful candidate for MD technology. The following section illuminates the state-of-the-art MD membranes produced using these nanomaterials, with a focus on enhanced activity due to their addition.

3.2.1. Carbon Nanotubes (CNTs)

Carbon nanotubes (CNTs) possess a wide range of superlative physical, chemical, and thermal characteristics, which make them a perfect filler material for polymeric membrane composites [92]. Carbon nanotubes are allotropes of carbon possessing a hexagonal lattice rolled up in a hollow cylindrical structure [93]. They have diameters ranging from 1 to 100 nm and up to a few millimeters in length, which may result in increased permeability within composite membranes [94]. CNTs are widely classified based on the number of sheet layers of graphene used for manufacture, i.e., single-walled CNTs (SWCNTs) and CNTs with multiple walls (MWCNTs) [95]. The introduction of CNTs offers alternative routes for transporting water vapor across the membrane, thereby dramatically enhancing the efficiency of the membrane [96]. Several CNT studies show that they improve the behavior of adsorption–desorption and reduce the frictional behavior of the walls, which provides favorable conditions for selective transmembrane flux [97].
CNTs also possess a nonpolar aromatic structure that is hydrophobic in nature and hence is used to improve and enhance the anti-wetting ability of the MD membrane, along with a significant contribution to enhance the mechanical strength, self-cleaning functions, and anti-fouling capacity of the MD membrane [25]. CNTs either form beads on the surface of the membrane, or they protrude out wholly which increases the microstructure roughness, contributing directly to surface hydrophobicity [97]. Silva et al., 2015, investigated the effects of MWCNT doping onto PVDF membranes and found an optimum 0.2 wt% that resulted in large-sized pores having a sponge-like consistency that resulted in a permeate flux of 9.5 × 10−3 L m−2 s−1 with a salt rejection of 100% over 60 min [98].
In another study, CNTs promoted the repulsion force required for Knudsen and molecular diffusion, enabled in the ease of surface diffusion, and deterred the boundary layer effect in viscous flow, allowing for increased vapor flux and anti-wetting characteristics [99]. CNTs in their bucky paper configuration also exhibit favorable properties such as high porosity (90%) and higher contact angle (113°) with low thermal conductivity. The CNT-bucky paper-based modified membranes seem to be equivalent to that of commercially available PTFE membranes. However, CNT-bucky paper-based membranes exhibit a lower lifespan and show flux decline in the MD operation due to micro-crack formation [100]. Thus, CNT-based membranes show interesting prospects in MD research, but a lot of investigation is pending regarding its viability.

3.2.2. Graphene and Graphene Oxide (GO)

Graphene and graphene oxide (GO) are an emerging class of nanofiller materials used for MD application. They possess unique structural and functional properties that set them apart from traditionally used NPs [101]. Graphene modification leads to added hierarchal microstructure roughness that helps in fabricating robust anti-wetting surfaces. Furthermore, like CNTs, they also show selective permeability because of the various functional groups present on their surface. When graphene-based membranes were tested under a mixture of saline water containing various surfactants, they possessed exceptional anti-fouling characteristics, which are attributed to the charge neutrality of the fabricated membranes [102]. Woo et al. studied the effects of varying the graphene doping concentration onto PVDF and found that 0.5 wt% graphene had a superior contact angle, pore size, porosity, and LEPw as compared with neat membranes. However, increasing it further than 0.7 wt% resulted in the aggregation of graphene particles that was counterproductive for the AGMD operation [103].
Graphene oxide is a graphene derivative that has a novel architecture with the availability of various chemical functional groups on their surface that shows potential for selective separation processes. PVDF/GO membranes show higher fluxes due to the presence of higher porosity than pristine PVDF with favorable nanofiller–water interactions. They also tend to show enhanced mechanical strength as compared with pristine PVDF [104]. However, GO is inherently hydrophilic, which makes it necessary to functionalize it in order to achieve exceptional MD performance [105]. GO-ODA (octadecylamine-functionalized graphene oxide) membranes show high membrane flux due to an additional path of transport for vapor molecules [106]. The functionalization of GO by n-butylamine before introducing it into flat sheet PVDF membranes results in high mechanical strength due to the uniform crystal structure and superior interfacial interactions [106]. The blending of reduced GO into the PVDF membrane matrix resulted in enhanced MD performance due to morphological and functional modifications in the membrane [107]. Figure 9 gives a brief overview on the enhancement of membrane properties due to the addition of GO NPs.

3.2.3. Quantum Dots (QDs)

Quantum dots have recently attracted much attention as an MD membrane modifier because of their physicochemical properties such as high biocompatibility, complete chemical inertness, ultra-small size, rich surface functional groups, and high anti-fouling ability. QDs are relatively more straightforward and inexpensive to incorporate on the membrane surface than other surface modifiers due to their ultra-small size and rich functionality. These modifiers can be selectively incorporated just on the substrate layer or the entire membrane depending upon the cost, material usage, and desired properties of the membrane. A QDs-enhanced MD membrane has shown significantly improved permeability, porosity, antibacterial, and anti-fouling capacity. In a study carried out by Jafari, A., graphene quantum dots (GQDs)-incorporated PVDF membranes, synthesized by the electrospinning method, were tested in an AGDM system. The addition of GQDs NPs caused a slight decline in the liquid entry pressure (LEPw) and contact angle, which can be related to the lower hydrophobicity of the membrane after modification because of the existence of hydrophilic functional groups in GQDs NPs. Overall, the obtained GQDs/PVDF electrospun membrane exhibited a more compact structure, acceptable contact angle, LEPw, and porosity. The modified membrane exhibited better anti-wetting properties than base PVDF membranes in a 60-h test. Further, the modified membrane GQD3P (0.25 wt% GQDs) achieved a higher flux value (17.6 L m−2 h−1) and a salt rejection of about 99.7%, compared with a pure PVDF membrane, which could achieve a membrane flux of 14.7 L m−2 h−1 and salt rejection rate of about 94.5%, which could be related to blockages caused due to a higher concentration of NPs, greater compactness of nanofibers ensuring lesser distance between the layers, and higher LEPw [84]. In another study, GOQDs were coated on the surface of the PVDF membrane through helium plasma-induced surface grafting of polyethylene glycol (PEG), followed by treatment with APTMS and covalent linkage between amine groups available on the PVDF membrane and carboxylic groups available on the GOQDs. The achieved GOQDs-modified PVDF membrane showed significantly enhanced water permeability, stability, antibacterial, and anti-biofouling properties. The improved anti-biofouling capacity was attributed to the presence of uniformly dispersed GOQDs NPs with a large number of active edges present on their surface, causing physical piercing of bacterial cells upon direct contact. The modified membrane also exhibited a lesser flux drop (23.4%) compared with the pristine PVDF membrane (65.7%) after 12 h [108]. Figure 10 illustrates the methodology and application of QDs into the MD membrane.
Based on a few reported studies where QDs have been utilized in an MD membrane, Table 3 represents the membrane performance in terms of membrane water flux, salt rejection rate, contact angle, and other applications. In short, it can be concluded that QDs-doped membranes seem to be efficient for improved anti-biofouling as well as long-term MD performance.

3.3. Metal Organic Framework (MOFs)

In recent decades, membrane-based separation methods have attracted a lot of attention due to their applicability in a variety of methods for water and wastewater purification. As earlier mentioned, the MD method is reliable, flexible, and energy-efficient; however, the existence of inorganic and organic pollutants in water makes these methods highly prone to fouling and wetting that reduces the selectivity and overall efficiency of the membrane [112]. Therefore, in an effort to counter the aforementioned concerns, membranes are coated, impregnated, or mixed with nanomaterials, oxides of graphene, or metal oxide frameworks (MOFs) [113]. All these materials notably enhanced the hydrophobicity, selectivity, and performance of membranes in MD operations [113]. Conversely, MOFs have received substantial attention in MD operations owing to their desired physio-chemical characteristics like high surface area, porosity, and intensive capacity to combine inorganic metal centers with organic linkers by coordinate bonds without altering the framework [114,115]. Briefly, MOFs are hybrid inorganic-centered organic microporous crystalline structures. Although the applicability of MOFs is widespread, these materials are thermodynamically unstable. In recent investigations, MOFs with clustered centers of aluminum (Al), zirconium (Zr), and iron (Fe) exhibited stable characteristics for water treatment processes [116]. Prominently, till date, MOFs-incorporated membranes are only used for DCMD and VMD [117]. Figure 11 indicates the advantageous features of MOFs-doped MD and the process involved in the incorporation of MOFs in polymeric membranes.
As per the few reports, the incorporation of various MOFs reverts the membrane hydrophobicity to hydrophilicity, hence improving its wetting resistance and anti-fouling properties. The wettability of the membrane can be studied by determining its LEPw value. An increase in the LEPw value can be linked to the reduced pore size and increased hydrophobicity of the membrane, hence enhancing its wetting resistance. In a recent study, the LEPw value of an MOF-modified MD membrane was significantly raised to 3 bar, owing to the growth of hydrophobic floss on the surface [116]. In addition to that, MOFs-enhanced polymeric membranes also exhibited higher permeability and selectivity [118]. A summary of MOFs-doped polymeric membranes for MD processes is shown in Table 4.

4. Nanomaterials for Fouling Control in MD Process

Mitigating membrane fouling and surface wetting are amongst the most prominent areas of research in MD studies. The main aim while fabricating membranes for long-term operations is to design anti-wetting and superhydrophobic surfaces. The wetting resistance of the surface is controlled by the surface free energy of the membrane or membrane geometry. Thus, in order to minimize surface wettability, the two most commonly used approaches are (1) generating a hierarchal structure with micro/nano roughness or (2) lowering the surface free energy for the membrane by chemical modifications [121]. Altering the surface chemistry to generate membranes with low surface free energy alters membrane–liquid interactions, while fabricating hierarchal structures creates air pockets that help in achieving anti-wetting behavior. The superhydrophobic layer also shows anti-wetting ability against feed solutions possessing components having low surface tension that can quickly enter the pores of the membrane and thus contribute to pore wetting and fouling. Furthermore, these membranes tend to have higher permeability, thus improving membrane performance [122]. In this context, nanomaterials have been widely used to achieve superior liquid repellency, thus tackling the problem of membrane fouling and pore wetting [123]. Nanomaterial modifications onto electrospun nanofibrous membranes (ENMs) are often used to generate a hierarchal surface with several air pockets to repel water droplets and achieve the lotus leaf effect [124]. Inorganic metalloid and metal oxides such as titanium dioxide and silicon dioxide are extensively used to generate superhydrophobic membranes for MD application. TiO2 and SiO2 have an abundant amount of hydroxyl surface groups present on their surface that makes it easy to functionalize them with fluoroalkylsilanes and fabricate multifunctional membranes [125,126]. Application of carbon-based nanomaterials in membrane fabrication has enabled achieving robust anti-wetting surfaces owing to their excellent physicochemical characteristics [127,128]. The incorporation of MOFs results in an increment of the surface roughness of the membrane and thus enhances the anti-wetting behavior. The effects of anti-wetting behavior due to the incorporation of NPs are showcased in Table 5.

5. Current Challenges and Future Outlook

Membrane modifications utilizing nanomaterials, such as NPs, nanotubes, nano-composites, and organic metal frameworks, face challenges that prove futile for the membrane as well as for the overall process [1,135]. The most prominent issues are discussed as follows:
  • Efficient synthesis methods: The use of nanomaterials poses many barriers to successful membrane modification, such as membrane pore blocking and non-uniform dispersal of membrane nanomaterials, among others [1,136]. Hence, it is important to test and devise an efficient membrane synthesis method according to the requirements for successful membrane modification. Otherwise, the entire process may fail because the membranes will not be efficient and successful in their position;
  • Appropriate integration of the materials: The nanomaterials used to modify the membranes must adhere or embed in the membranes appropriately; otherwise, they may leach out over time [136]. The different commonly used nanomaterials, such as CNTs and metal oxide NPs, among others, could be linked to the membranes via different groups, like -OH, present on their surface with the functional groups present in the membrane matrix via hydrogen or covalent bonds. This will not only make the membrane effective and efficient but also enhance its stability and life;
  • Stability of the fillers: The stability of the filler material is very important as it defines the overall membrane effectiveness and efficiency [135,136,137]. They must be firmly embedded with the aid of various bonds in the membrane matrix. The firmly embedded filler material in the membrane satisfies that while in use, the filler material will not disintegrate from the membrane. The membrane can, therefore, be utilized for longer periods of time without losing its structural and functional integrity. However, presently much work is needed for developing such filler materials as well as synthesis methods that approve the stability of the fillers to cent percent;
  • Conservation of the functional properties: It is very important to conserve the required functional attributes of membranes for which they are sought. Currently, however, there is a definite loss of functional characteristics during membrane synthesis, meaning that the membranes cannot show the theoretical extent of their functional attributes. In addition, there is a loss of functionality over time during membrane operations. Therefore, there is a requirement of a sustainable synthesis method that is capable of withholding the functional capacity of the membranes to the best;
  • Membrane strength: The filler materials, sometimes instead of increasing the strength of the membrane, make them weak and brittle. Therefore, this aspect should also be studied carefully before using any filler for the modification of membranes;
  • Membrane fouling: Fouling is one of the most important factors in membrane science that inhibits the use of membranes in large-scale applications. However, there are lots of studies carried out specifically to eradicate this single problem, but it still persists [2]. Therefore, there is a need to tackle this problem appropriately for better employment of membranes for large-scale applications.
The recently reported literature based on the incorporation of nanomaterials suggested a few prominent future aspects which can be explored, such as:
(a)
Risk of particle aggregation while incorporating nanomaterials onto the membrane;
(b)
Utilization of cheap nanomaterials in order to reduce the overall cost;
(c)
Rougher membrane surface may enhance membrane fouling, which must be optimized;
(d)
Risk of peeling off or particle washout after a long-term experiment.
Figure 12 summarizes the serious issues in the manufacture of nanomaterial-based polymeric MD membranes, which need high attention and must be systematically discussed to attain improved MD membrane performance and scalability.
Thus, extensive research must be encouraged for the fabrication of superhydrophobic and anti-wetting MD membranes via the incorporation of new generation nanomaterials as inadequate suppliers are available. In addition to that, collaborative research work is highly needed between academia and industrial organizations as these membranes need to be commercialized at the industrial scale.

6. Concluding Remarks

The main goal of this review was to study the new generation nanomaterials currently being used for the development of state-of-the-art MD membranes. In MD research, nanomaterials such as TiO2, SiO2, MOFs, CNTs, and graphene and its derivatives were evaluated from a holistic perspective as additives. According to extensive literature studies, the introduction of NPs generally improves the superhydrophobic property of the membrane, reduces the fouling tendency, increases the porosity of the membrane, and improves the mechanical and chemical stability, which ensures the production of a superior MD membrane class. Improvement of these properties enables the fabrication of robust anti-wetting surfaces for long-term MD operations. The methods of incorporation of NPs as a part of the membrane fabrication process are also briefly discussed. Even though using NPs as filler materials results in synergistic effects by imparting novel structural and functional characteristics, several problems such as high fabrication costs, scalability, hazardous nature, and arduous methods of designing are commonly observed. Therefore, in order to extend the work done in small-scale laboratories to large-scale industries, it is desired that future researchers seek the discovery of more viable and successful solutions. Thus, nanomaterial-based membrane engineering is strongly encouraged to reduce the research gap towards developing a supreme class of MD membranes.

Author Contributions

Conceptualization, S.S.R., M.G., S.-S.C., M.K.P., R.S.; methodology, H.S.B., R.D., C.K.D., writing—original draft preparation, S.S.R., H.S.B., R.D., R.S., C.K.D.; writing—review and editing, S.S.R. and H.S.B.; visualization, R.D.; supervision, S.-S.C.; project administration, M.G.; funding acquisition, S.-S.C. All authors have read and agreed to the published version of the manuscript.

Funding

Ministry of Science and Technology (MOST), Taiwan, Republic of China (ROC), under grant number 107-2221-E-027-001-MY3.

Acknowledgments

All the authors are grateful for the financial support from the Ministry of Science and Technology (MOST), Taiwan, Republic of China (ROC), under grant number 107-2221-E-027-001-MY3.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Purkait, M.K.; Singh, R.; Haldar, D.; Mondal, P. Thermal Induced Membrane Separation Processes; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
  2. Purkait, M.K.; Singh, R. Membrane Technology in Separation Science; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  3. Purkait, M.K.; Sinha, M.K.; Mondal, P.; Singh, R. Stimuli Responsive Polymeric Membranes: Smart Polymeric Membranes; Academic Press: Amsterdam, The Netherlands, 2018. [Google Scholar]
  4. Alkhudhiri, A.; Darwish, N.; Hilal, N. Membrane distillation: A comprehensive review. Desalination 2012, 287, 2–18. [Google Scholar] [CrossRef]
  5. Ray, S.S.; Chen, S.S.; Sangeetha, D.; Chang, H.M.; Thanh, C.N.D.; Le, Q.H.; Ku, H.M. Developments in forward osmosis and membrane distillation for desalination of waters. Environ. Chem. Lett. 2018, 16, 1247–1265. [Google Scholar] [CrossRef]
  6. Kochkodan, V.; Hilal, N. A comprehensive review on surface modified polymer membranes for biofouling mitigation. Desalination 2015, 356, 187–207. [Google Scholar] [CrossRef]
  7. Gilron, J.; Song, L.; Sirkar, K.K. Design for cascade of crossflow direct contact membrane distillation. Ind. Eng. Chem. Res. 2007, 46, 2324–2334. [Google Scholar] [CrossRef]
  8. Lalia, B.S.; Kochkodan, V.; Hashaikeh, R.; Hilal, N.J.D. A review on membrane fabrication: Structure, properties and performance relationship. Desalination 2013, 326, 77–95. [Google Scholar] [CrossRef]
  9. Ray, S.S.; Dommati, H.; Wang, J.-C.; Chen, S.-S. Solvent based Slurry Stereolithography 3D printed hydrophilic ceramic membrane for ultrafiltration application. Ceram. Int. 2020, 46, 12480–12488. [Google Scholar] [CrossRef]
  10. Bodell, B.R. Distillation of Saline Water Using Silicone Rubber Membrane. U.S. Patent US3361645A, 2 January 1968. [Google Scholar]
  11. Amaya-Vías, D.; López-Ramírez, J.A. Techno-economic assessment of air and water gap membrane distillation for Seawater desalination under different heat source scenarios. Water 2019, 11, 2117. [Google Scholar] [CrossRef] [Green Version]
  12. Camacho, L.M.; Dumée, L.; Zhang, J.; Li, J.D.; Duke, M.; Gomez, J.; Gray, S. Advances in membrane distillation for water desalination and purification applications. Water 2013, 5, 94–196. [Google Scholar] [CrossRef] [Green Version]
  13. Ray, S.S.; Deb, C.K.; Chang, H.M.; Chen, S.S.; Ganesapillai, M. Crosslinked PVDF-HFP-based hydrophobic membranes incorporated with CNF for enhanced stability and permeability in membrane distillation. J. Appl. Polym. Sci. 2019, 136, 48021. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Wang, X.; Cui, Z.; Drioli, E.; Wang, Z.; Zhao, S. Enhancing wetting resistance of poly (vinylidene fluoride) membranes for vacuum membrane distillation. Desalination 2017, 415, 58–66. [Google Scholar] [CrossRef]
  15. Li, W.R.; Xie, X.B.; Shi, Q.S.; Zeng, H.Y.; You-Sheng, O.Y.; Chen, Y.B. Antibacterial activity and mechanism of silver nanoparticles on Escherichia coli. Appl. Microbiol. Biotechnol. 2010, 85, 1115–1122. [Google Scholar] [CrossRef] [PubMed]
  16. Prabhu, S.; Poulose, E.K. Silver nanoparticles: Mechanism of antimicrobial action, synthesis, medical applications, and toxicity effects. Int. Nano Lett. 2012, 2, 32. [Google Scholar] [CrossRef] [Green Version]
  17. Nthunya, L.N.; Gutierrez, L.; Derese, S.; Nxumalo, E.N.; Verliefde, A.R.; Mamba, B.B.; Mhlanga, S.D. A review of nanoparticle-enhanced membrane distillation membranes: Membrane synthesis and applications in water treatment. J. Chem. Technol. Biotechnol. 2019, 94, 2757–2771. [Google Scholar] [CrossRef]
  18. Wang, Y.; He, G.; Shao, Y.; Zhang, D.; Ruan, X.; Xiao, W.; Li, X.; Wu, X.; Jiang, X.J.S.; Technology, P. Enhanced performance of superhydrophobic polypropylene membrane with modified antifouling surface for high salinity water treatment. Sep. Purif. Technol. 2019, 214, 11–20. [Google Scholar] [CrossRef]
  19. Barati, G. Science and Engineering of Superhydrophobic Surfaces: Review of Corrosion Resistance. Chem. Mech. Stab. Arab. J. Chem. 2018, 13, 1763–1802. [Google Scholar] [CrossRef]
  20. Deka, B.J.; Guo, J.; Khanzada, N.K.; An, A.K. Omniphobic re-entrant PVDF membrane with ZnO nanoparticles composite for desalination of low surface tension oily seawater. Water Res. 2019, 165, 114982. [Google Scholar] [CrossRef]
  21. Natarajan, K.J.C. Antibiofilm activity of epoxy/Ag-TiO2 polymer nanocomposite coatings against Staphylococcus aureus and Escherichia coli. Coatings 2015, 5, 95–114. [Google Scholar]
  22. Furukawa, H.; Cordova, K.E.O.; Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [Green Version]
  23. Lin, R.; Hernandez, B.V.; Ge, L.; Zhu, Z. Metal organic framework based mixed matrix membranes: An overview on filler/polymer interfaces. J. Mater. Chem. A 2018, 6, 293–312. [Google Scholar] [CrossRef]
  24. Kadhom, M.; Deng, B. Metal-organic frameworks (MOFs) in water filtration membranes for desalination and other applications. Appl. Mater. Today 2018, 11, 219–230. [Google Scholar] [CrossRef]
  25. Das, R.; Ali, M.E.; Abd Hamid, S.B.; Ramakrishna, S.; Chowdhury, Z.Z. Carbon nanotube membranes for water purification: A bright future in water desalination. Desalination 2014, 336, 97–109. [Google Scholar] [CrossRef]
  26. Zhao, D.L.; Chung, T.-S. Applications of carbon quantum dots (CQDs) in membrane technologies: A review. Water Res. 2018, 147, 43–49. [Google Scholar] [CrossRef] [PubMed]
  27. Zodrow, K.; Brunet, L.; Mahendra, S.; Li, D.; Zhang, A.; Li, Q.; Alvarez, P.J. Polysulfone ultrafiltration membranes impregnated with silver nanoparticles show improved biofouling resistance and virus removal. Water Res. 2009, 43, 715–723. [Google Scholar] [CrossRef] [Green Version]
  28. Dumee, L.; Velleman, L.; Sears, K.; Hill, M.; Schutz, J.; Finn, N.; Duke, M.; Gray, S. Control of porosity and pore size of metal reinforced carbon nanotube membranes. Membranes 2011, 1, 25–36. [Google Scholar] [CrossRef] [Green Version]
  29. Nasreen, S.A.A.N.; Sundarrajan, S.; Nizar, S.A.S.; Balamurugan, R.; Ramakrishna, S. Advancement in electrospun nanofibrous membranes modification and their application in water treatment. Membranes 2013, 3, 266–284. [Google Scholar] [CrossRef]
  30. Ma, L.; Dong, X.; Chen, M.; Zhu, L.; Wang, C.; Yang, F.; Dong, Y. Fabrication and water treatment application of carbon nanotubes (CNTs)-based composite membranes: A review. Membranes 2017, 7, 16. [Google Scholar] [CrossRef] [Green Version]
  31. Lippmann, G. Endosmose entre deux liquides de même composition chimique et de températures différentes. Compt. Rend 1907, 145, 104–105. [Google Scholar]
  32. Aubert, M. Thermo-osmose. Verlag Nicht Ermittelbar; Gauthier-Villars: Paris, France, 1912. [Google Scholar]
  33. Villalobos García, J.; Dow, N.; Milne, N.; Zhang, J.; Naidoo, L.; Gray, S.; Duke, M. Membrane distillation trial on textile wastewater containing surfactants using hydrophobic and hydrophilic-coated polytetrafluoroethylene (PTFE) membranes. Membranes 2018, 8, 31. [Google Scholar] [CrossRef] [Green Version]
  34. Nawi, M.; Izati, N.; Bilad, M.R.; Zolkhiflee, N.; Nordin, N.A.H.; Lau, W.J.; Narkkun, T.; Faungnawakij, K.; Arahman, N.; Mahlia, T.M.I. Development of a novel corrugated polyvinylidene difluoride membrane via improved imprinting technique for membrane distillation. Polymers 2019, 11, 865. [Google Scholar] [CrossRef] [Green Version]
  35. Gryta, M. Effectiveness of water desalination by membrane distillation process. Membranes 2012, 2, 415–429. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, P.; Chung, T.S. Recent advances in membrane distillation processes: Membrane development, configuration design and application exploring. J. Membr. Sci. 2015, 474, 39–56. [Google Scholar] [CrossRef]
  37. Eykens, L.; De Sitter, K.; Dotremont, C.; Pinoy, L.; Van der Bruggen, B. How to optimize the membrane properties for membrane distillation: A review. Ind. Eng. Chem. Res. 2016, 55, 9333–9343. [Google Scholar] [CrossRef]
  38. Khayet, M.; Matsuura, T. Pervaporation and vacuum membrane distillation processes: Modeling and experiments. AIChE J. 2004, 50, 1697–1712. [Google Scholar] [CrossRef]
  39. Liu, X.; Liu, H.; Li, P. Effect of polymer dope concentration on the morphology and performance of PES/PDMS hollow fiber composite membrane for gas separation. PES 2017, 500, 72. [Google Scholar] [CrossRef]
  40. Kim, J.; Van der Bruggen, B.J.E.P. The use of nanoparticles in polymeric and ceramic membrane structures: Review of manufacturing procedures and performance improvement for water treatment. Environ. Pollut. 2010, 158, 2335–2349. [Google Scholar] [CrossRef]
  41. Baghbanzadeh, M.; Rana, D.; Lan, C.Q.; Matsuura, T.J.S.; Technology, P. Effects of hydrophilic silica nanoparticles and backing material in improving the structure and performance of VMD PVDF membranes. Sep. Purif. Technol. 2016, 157, 60–71. [Google Scholar] [CrossRef]
  42. Efome, J.E.; Baghbanzadeh, M.; Rana, D.; Matsuura, T.; Lan, C.Q.J.D. Effects of superhydrophobic SiO2 nanoparticles on the performance of PVDF flat sheet membranes for vacuum membrane distillation. Desalination 2015, 373, 47–57. [Google Scholar] [CrossRef]
  43. Attia, H.; Alexander, S.; Wright, C.J.; Hilal, N.J.D. Superhydrophobic electrospun membrane for heavy metals removal by air gap membrane distillation (AGMD). Desalination 2017, 420, 318–329. [Google Scholar] [CrossRef] [Green Version]
  44. Bhadra, M.; Roy, S.; Mitra, S.J.D. Desalination across a graphene oxide membrane via direct contact membrane distillation. Desalination 2016, 378, 37–43. [Google Scholar] [CrossRef]
  45. Roy, S.; Bhadra, M.; Mitra, S.J.S.; Technology, P. Enhanced desalination via functionalized carbon nanotube immobilized membrane in direct contact membrane distillation. Sep. Purif. Technol. 2014, 136, 58–65. [Google Scholar] [CrossRef]
  46. Huang, F.Y.; Arning, A. Performance comparison between polyvinylidene fluoride and polytetrafluoroethylene hollow fiber membranes for direct contact membrane distillation. Membranes 2019, 9, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Eykens, L.; De Sitter, K.; Dotremont, C.; Pinoy, L.; Van der Bruggen, B.J.S.; Technology, P. Membrane synthesis for membrane distillation: A review. Sep. Purif. Technol. 2017, 182, 36–51. [Google Scholar] [CrossRef]
  48. El-Bourawi, M.S.; Ding, Z.; Ma, R.; Khayet, M. A framework for better understanding membrane distillation separation process. J. Membr. Sci. 2006, 285, 4–29. [Google Scholar] [CrossRef]
  49. Ma, M.; Hill, R.M. Superhydrophobic surfaces. Curr. Opin. Colloid Interface Sci. 2006, 11, 193–202. [Google Scholar] [CrossRef]
  50. Johnson, D.J.; Oatley-Radcliffe, D.L.; Hilal, N.J.D. State of the art review on membrane surface characterisation: Visualisation, verification and quantification of membrane properties. Desalination 2018, 434, 12–36. [Google Scholar] [CrossRef] [Green Version]
  51. Laganà, F.; Barbieri, G.; Drioli, E. Direct contact membrane distillation: Modelling and concentration experiments. J. Membr. Sci. 2000, 166, 1–11. [Google Scholar] [CrossRef]
  52. Srisurichan, S.; Jiraratananon, R.; Fane, A.G. Mass transfer mechanisms and transport resistances in direct contact membrane distillation process. J. Membr. Sci. 2006, 277, 186–194. [Google Scholar] [CrossRef]
  53. Razmjou, A.; Mansouri, J.; Chen, V. The effects of mechanical and chemical modification of TiO2 nanoparticles on the surface chemistry, structure and fouling performance of PES ultrafiltration membranes. J. Membr. Sci. 2011, 378, 73–84. [Google Scholar] [CrossRef]
  54. Ray, S.S.; Lee, H.-K.; Kwon, Y.-N. Review on blueprint of designing anti-wetting polymeric membrane surfaces for enhanced membrane distillation performance. Polymers 2020, 12, 23. [Google Scholar]
  55. Ju, J.; Fejjari, K.; Cheng, Y.; Liu, M.; Li, Z.; Kang, W.; Liao, Y. Engineering hierarchically structured superhydrophobic PTFE/POSS nanofibrous membranes for membrane distillation. Desalination 2020, 486, 114481. [Google Scholar] [CrossRef]
  56. Kang, G.D.; Cao, Y.M. Application and modification of poly (vinylidene fluoride)(PVDF) membranes–a review. J. Membr. Sci. 2014, 463, 145–165. [Google Scholar] [CrossRef]
  57. Nguyen, Q.T.; Alaoui, O.T.; Yang, H.; Mbareck, C. Dry-cast process for synthetic microporous membranes: Physico-chemical analyses through morphological studies. J. Membr. Sci. 2010, 358, 13–25. [Google Scholar] [CrossRef]
  58. Kim, J.-K.; Taki, K.; Ohshima, M. Preparation of a unique microporous structure via two step phase separation in the course of drying a ternary polymer solution. Langmuir 2007, 23, 12397–12405. [Google Scholar] [CrossRef] [PubMed]
  59. Jaworek, A.; Krupa, A.; Czech, T. Modern electrostatic devices and methods for exhaust gas cleaning: A brief review. J. Electrost. 2007, 65, 133–155. [Google Scholar] [CrossRef]
  60. Ray, S.S.; Chen, S.-S.; Li, C.-W.; Nguyen, N.C.; Nguyen, H.T. A comprehensive review: Electrospinning technique for fabrication and surface modification of membranes for water treatment application. RSC Adv. 2016, 6, 85495–85514. [Google Scholar] [CrossRef]
  61. Ray, S.S.; Chen, S.-S.; Nguyen, N.C.; Hsu, H.-T.; Nguyen, H.T.; Chang, C.-T. Poly (vinyl alcohol) incorporated with surfactant based electrospun nanofibrous layer onto polypropylene mat for improved desalination by using membrane distillation. Desalination 2017, 414, 18–27. [Google Scholar] [CrossRef]
  62. Ray, S.S.; Chen, S.-S.; Nguyen, N.C.; Nguyen, H.T. Electrospinning: A Versatile Fabrication Technique for Nanofibrous Membranes for Use in Desalination. In Nanoscale Materials in Water Purification; Elsevier: Amsterdam, The Netherlands, 2019; pp. 247–273. [Google Scholar]
  63. Al-Furaiji, M.; Arena, J.T.; Ren, J.; Benes, N.; Nijmeijer, A.; McCutcheon, J.R. Triple-layer nanofiber membranes for treating high salinity brines using direct contact membrane distillation. Membranes 2019, 9, 60. [Google Scholar] [CrossRef] [Green Version]
  64. Zare, S.; Kargari, A. Membrane properties in membrane distillation. In Emerging Technologies for Sustainable Desalination Handbook; Elsevier: Amsterdam, The Netherlands, 2018; pp. 107–156. [Google Scholar]
  65. Low, S.-C.; Ng, Q.-H. Progress of stimuli responsive membranes in water treatment. In Advanced Nanomaterials for Membrane Synthesis and its Applications; Elsevier: Amsterdam, The Netherlands, 2019; pp. 69–99. [Google Scholar]
  66. Wei, C.C.; Li, K. Preparation and characterization of a robust and hydrophobic ceramic membrane via an improved surface grafting technique. Ind. Eng. Chem. Res. 2009, 48, 3446–3452. [Google Scholar] [CrossRef]
  67. El-Arnaouty, M.; Abdel Ghaffar, A.; Eid, M.; Aboulfotouh, M.E.; Taher, N.; Soliman, E.-S. Nano-modification of polyamide thin film composite reverse osmosis membranes by radiation grafting. J. Radiat. Res. Appl. Sci. 2018, 11, 204–216. [Google Scholar] [CrossRef] [Green Version]
  68. Lee, H.K.; Kim, W.; Kim, Y.M.; Kwon, Y.N. Surface modification of polyvinylidene fluoride membrane for enhanced wetting resistance. Appl. Surf. Sci. 2019, 491, 32–42. [Google Scholar] [CrossRef]
  69. Akhavan, B.; Jarvis, K.; Majewski, P. Plasma polymer-functionalized silica particles for heavy metals removal. ACS Appl. Mater. Interfaces 2015, 7, 4265–4274. [Google Scholar] [CrossRef]
  70. Song, Y.; Fan, J.-B.; Wang, S. Recent progress in interfacial polymerization. Mater. Chem. Front. 2017, 1, 1028–1040. [Google Scholar] [CrossRef]
  71. Baig, M.I.; Ingole, P.G.; Jeon, J.D.; Hong, S.U.; Choi, W.K.; Lee, H.K. Water vapor transport properties of interfacially polymerized thin film nanocomposite membranes modified with graphene oxide and GO-TiO2 nanofillers. Chem. Eng. J. 2019, 373, 1190–1202. [Google Scholar] [CrossRef]
  72. Ray, S.S.; Chen, S.-S.; Chang, H.-M.; Thanh, C.N.D.; Le, H.Q.; Nguyen, N.C. Enhanced desalination using a three-layer OTMS based superhydrophobic membrane for a membrane distillation process. RSC Adv. 2018, 8, 9640–9650. [Google Scholar] [CrossRef] [Green Version]
  73. Han, S.W.; Kim, K.D.; Seo, H.O.; Kim, I.H.; Jeon, C.S.; An, J.E.; Kim, J.H.; Uhm, S.; Kim, Y.D. Oil–Water Separation Using Superhydrophobic PET Membranes Fabricated Via Simple Dip-Coating of PDMS–SiO2 Nanoparticles. Macromol. Mater. Eng. 2017, 302, 1700218. [Google Scholar] [CrossRef]
  74. Yimsiri, P.; Mackley, M.R.J.C.E.S. Spin and dip coating of light-emitting polymer solutions: Matching experiment with modelling. Chem. Eng. Sci. 2006, 61, 3496–3505. [Google Scholar] [CrossRef]
  75. Nassrullah, H.; Makanjuola, O.; Janajreh, I.; AlMarzooqi, F.A.; Hashaikeh, R. Incorporation of nanosized LTL zeolites in dual-layered PVDF-HFP/cellulose membrane for enhanced membrane distillation performance. J. Membr. Sci 2020, 118298. [Google Scholar] [CrossRef]
  76. Hamzah, N.; Leo, C.J.S.; Technology, P. Fouling prevention in the membrane distillation of phenolic-rich solution using superhydrophobic PVDF membrane incorporated with TiO2 nanoparticles. Sep. Purif. Technol. 2016, 167, 79–87. [Google Scholar] [CrossRef]
  77. Razmjou, A.; Arifin, E.; Dong, G.; Mansouri, J.; Chen, V. Superhydrophobic modification of TiO2 nanocomposite PVDF membranes for applications in membrane distillation. J. Membr. Sci. 2012, 415, 850–863. [Google Scholar] [CrossRef]
  78. Ye, H.; Li, X.; Deng, L.; Li, P.; Zhang, T.; Wang, X.; Hsiao, B.S. Silver nanoparticle-enabled photothermal nanofibrous membrane for light-driven membrane distillation. Ind. Eng. Chem. Res. 2019, 58, 3269–3281. [Google Scholar] [CrossRef]
  79. Chen, L.H.; Huang, A.; Chen, Y.R.; Chen, C.H.; Hsu, C.C.; Tsai, F.Y.; Tung, K.L. Omniphobic membranes for direct contact membrane distillation: Effective deposition of zinc oxide nanoparticles. Desalination 2018, 428, 255–263. [Google Scholar] [CrossRef]
  80. Huang, Y.X.; Wang, Z.; Hou, D.; Lin, S. Coaxially electrospun super-amphiphobic silica-based membrane for anti-surfactant-wetting membrane distillation. J. Membr. Sci. 2017, 531, 122–128. [Google Scholar] [CrossRef]
  81. Han, M.; Dong, T.; Hou, D.; Yao, J.; Han, L. Carbon nanotube based Janus composite membrane of oil fouling resistance for direct contact membrane distillation. J. Membr. Sci. 2020, 118078. [Google Scholar] [CrossRef]
  82. Yan, K.K.; Jiao, L.; Lin, S.; Ji, X.; Lu, Y.; Zhang, L. Superhydrophobic electrospun nanofiber membrane coated by carbon nanotubes network for membrane distillation. Desalination 2018, 437, 26–33. [Google Scholar] [CrossRef]
  83. Grasso, G.; Galiano, F.; Yoo, M.J.; Mancuso, R.; Park, H.B.; Gabriele, B.; Figoli, A.; Drioli, E. Development of graphene-PVDF composite membranes for membrane distillation. J. Membr. Sci. 2020, 604, 118017. [Google Scholar] [CrossRef]
  84. Jafari, A.; Kebria, M.R.S.; Rahimpour, A.; Bakeri, G. Graphene quantum dots modified polyvinylidenefluride (PVDF) nanofibrous membranes with enhanced performance for air Gap membrane distillation. Chem. Eng. Process. Process Intensif. 2018, 126, 222–231. [Google Scholar] [CrossRef]
  85. Zhou, R.; Rana, D.; Matsuura, T.; Lan, C.Q. Effects of multi-walled carbon nanotubes (MWCNTs) and integrated MWCNTs/SiO2 nano-additives on PVDF polymeric membranes for vacuum membrane distillation. Sep. Purif. Technol. 2019, 217, 154–163. [Google Scholar] [CrossRef]
  86. Damodar, R.A.; You, S.-J.; Chou, H.-H. Study the self cleaning, antibacterial and photocatalytic properties of TiO2 entrapped PVDF membranes. J. Hazard. Mater. 2009, 172, 1321–1328. [Google Scholar] [CrossRef]
  87. Monteserín, C.; Blanco, M.; Murillo, N.; Pérez-Márquez, A.; Maudes, J.; Gayoso, J.; Manuel Laza, J.; Hernáez, E.; Aranzabe, E.; Vilas, J.L. Novel Antibacterial and Toughened Carbon-Fibre/Epoxy Composites by the Incorporation of TiO2 Nanoparticles Modified Electrospun Nanofibre Veils. Polymers 2019, 11, 1524. [Google Scholar]
  88. Dong, Z.-Q.; Ma, X.-H.; Xu, Z.-L.; Gu, Z.-Y. Superhydrophobic modification of PVDF–SiO2 electrospun nanofiber membranes for vacuum membrane distillation. RSC Adv. 2015, 5, 67962–67970. [Google Scholar] [CrossRef]
  89. Samaha, M.A.; Gad-el-Hak, M. Polymeric Slippery Coatings: Nature and Applications. Polymers 2014, 6, 1266–1311. [Google Scholar] [CrossRef] [Green Version]
  90. Chen, Y.; Zheng, R.; Wang, J.; Liu, Y.; Wang, Y.; Li, X.-M.; He, T. Laminated PTFE membranes to enhance the performance in direct contact membrane distillation for high salinity solution. Desalination 2017, 424, 140–148. [Google Scholar] [CrossRef]
  91. Ray, S.S.; Gandhi, M.; Chen, S.-S.; Chang, H.-M.; Dan, C.T.N.; Le, H.Q. Anti-wetting behaviour of a superhydrophobic octadecyltrimethoxysilane blended PVDF/recycled carbon black composite membrane for enhanced desalination. Environ. Sci. Water Res. Technol. 2018, 4, 1612–1623. [Google Scholar] [CrossRef]
  92. Lee, J.K.Y.; Chen, N.; Peng, S.; Li, L.; Tian, L.; Thakor, N.; Ramakrishna, S. Polymer-based composites by electrospinning: Preparation functionalization with nanocarbons. Prog. Polym. Sci. 2018, 86, 40–84. [Google Scholar] [CrossRef]
  93. Liu, W.-W.; Chai, S.-P.; Mohamed, A.R.; Hashim, U. Synthesis and characterization of graphene and carbon nanotubes: A review on the past and recent developments. J. Ind. Eng. Chem. 2014, 20, 1171–1185. [Google Scholar] [CrossRef]
  94. Elmarghany, M.R.; El-Shazly, A.H.; Rajabzadeh, S.; Salem, M.S.; Shouman, M.A.; Sabry, M.N.; Matsuyama, H.; Nady, N. Triple-layer nanocomposite membrane prepared by electrospinning based on modified PES with carbon nanotubes for membrane distillation applications. Membranes 2020, 10, 15. [Google Scholar] [CrossRef] [Green Version]
  95. Rashid, M.; Ralph, S.F. Carbon nanotube membranes: Synthesis, properties, and future filtration applications. Nanomaterials 2017, 7, 99. [Google Scholar] [CrossRef] [Green Version]
  96. Gethard, K.; Mitra, S. Carbon nanotube enhanced membrane distillation for online preconcentration of trace pharmaceuticals in polar solvents. Analyst 2011, 136, 2643–2648. [Google Scholar] [CrossRef]
  97. Bhadra, M.; Roy, S.; Mitra, S. A bilayered structure comprised of functionalized carbon nanotubes for desalination by membrane distillation. ACS Appl. Mater. Interfaces 2016, 8, 19507–19513. [Google Scholar] [CrossRef]
  98. Silva, T.L.; Morales-Torres, S.; Figueiredo, J.L.; Silva, A.M. Multi-walled carbon nanotube/PVDF blended membranes with sponge-and finger-like pores for direct contact membrane distillation. Desalination 2015, 357, 233–245. [Google Scholar] [CrossRef]
  99. An, A.K.; Lee, E.-J.; Guo, J.; Jeong, S.; Lee, J.-G.; Ghaffour, N. Enhanced vapor transport in membrane distillation via functionalized carbon nanotubes anchored into electrospun nanofibres. Sci. Rep. 2017, 7, 41562. [Google Scholar]
  100. Dumée, L.F.; Sears, K.; Schütz, J.; Finn, N.; Huynh, C.; Hawkins, S.; Duke, M.; Gray, S. Characterization and evaluation of carbon nanotube Bucky-Paper membranes for direct contact membrane distillation. J. Membr. Sci. 2010, 351, 36–43. [Google Scholar] [CrossRef] [Green Version]
  101. Alsebaeai, M.K.; Otitoju, T.A.; Makhtar, M.B.M.Z.; Ahmad, A.L. Membrane distillation: Progress in the improvement of dedicated membranes for enhanced hydrophobicity and desalination performance. J. Ind. Eng. Chem. 2020, 86, 13–34. [Google Scholar] [CrossRef]
  102. Seo, D.H.; Pineda, S.; Woo, Y.C.; Xie, M.; Murdock, A.T.; Ang, E.Y.; Jiao, Y.; Park, M.J.; Lim, S.I.; Lawn, M. Anti-fouling graphene-based membranes for effective water desalination. Nat. Commun. 2018, 9, 1–12. [Google Scholar] [CrossRef] [Green Version]
  103. Woo, Y.C.; Kim, Y.; Shim, W.-G.; Tijing, L.D.; Yao, M.; Nghiem, L.D.; Choi, J.-S.; Kim, S.-H.; Shon, H.K. Graphene/PVDF flat-sheet membrane for the treatment of RO brine from coal seam gas produced water by air gap membrane distillation. J. Membr. Sci. 2016, 513, 74–84. [Google Scholar] [CrossRef]
  104. Leaper, S.; Abdel-Karim, A.; Faki, B.; Luque-Alled, J.M.; Alberto, M.; Vijayaraghavan, A.; Holmes, S.M.; Szekely, G.; Badawy, M.I.; Shokri, N. Flux-enhanced PVDF mixed matrix membranes incorporating APTS-functionalized graphene oxide for membrane distillation. J. Membr. Sci. 2018, 554, 309–323. [Google Scholar] [CrossRef]
  105. Wang, Z.; Yu, H.; Xia, J.; Zhang, F.; Li, F.; Xia, Y.; Li, Y. Novel GO-blended PVDF ultrafiltration membranes. Desalination 2012, 299, 50–54. [Google Scholar] [CrossRef]
  106. Zahirifar, J.; Karimi-Sabet, J.; Moosavian, S.M.A.; Hadi, A.; Khadiv-Parsi, P. Fabrication of a novel octadecylamine functionalized graphene oxide/PVDF dual-layer flat sheet membrane for desalination via air gap membrane distillation. Desalination 2018, 428, 227–239. [Google Scholar] [CrossRef]
  107. Ma, J.; Ping, D.; Dong, X. Recent developments of graphene oxide-based membranes: A review. Membranes 2017, 7, 52. [Google Scholar]
  108. Zeng, Z.; Yu, D.; He, Z.; Liu, J.; Xiao, F.-X.; Zhang, Y.; Wang, R.; Bhattacharyya, D.; Tan, T.T.Y. Graphene oxide quantum dots covalently functionalized PVDF membrane with significantly-enhanced bactericidal and antibiofouling performances. Sci. Rep. 2016, 6, 20142. [Google Scholar] [CrossRef] [Green Version]
  109. Lei, S.; Zeng, M.; Huang, D.; Wang, L.; Zhang, L.; Xi, B.; Ma, W.; Chen, G.; Cheng, Z. Synergistic High-flux Oil–Saltwater Separation and Membrane Desalination with Carbon Quantum Dots Functionalized Membrane. ACS Sustain. Chem. Eng. 2019, 7, 13708–13716. [Google Scholar] [CrossRef]
  110. Zhao, G.; Hu, R.; Li, J.; Zhu, H. Graphene oxide quantum dots embedded polysulfone membranes with enhanced hydrophilicity, permeability and antifouling performance. Sci. China Mater. 2019, 62, 1177–1187. [Google Scholar] [CrossRef] [Green Version]
  111. Lecaros, R.L.G.; Deseo, K.M.; Hung, W.-S.; Tayo, L.L.; Hu, C.-C.; An, Q.-F.; Tsai, H.-A.; Lee, K.-R.; Lai, J.-Y. Influence of integrating graphene oxide quantum dots on the fine structure characterization and alcohol dehydration performance of pervaporation composite membrane. J. Membr. Sci. 2019, 576, 36–47. [Google Scholar] [CrossRef]
  112. Li, W.; Zhang, Y.; Li, Q.; Zhang, G. Metal−organic framework composite membranes: Synthesis and separation applications. Chem. Eng. Sci. 2015, 135, 232–257. [Google Scholar] [CrossRef]
  113. James, S.L. Metal-organic frameworks. Chem. Soc. Rev. 2003, 32, 276–288. [Google Scholar] [CrossRef] [PubMed]
  114. Lalia, B.S.; Guillen, E.; Arafat, H.A.; Hashaikeh, R. Nanocrystalline cellulose reinforced PVDF-HFP membranes for membrane distillation application. Desalination 2014, 332, 134–141. [Google Scholar] [CrossRef]
  115. Li, X.; Liu, Y.; Wang, J.; Gascon, J.; Li, J.; Van der Bruggen, B. Metal–organic frameworks based membranes for liquid separation. Chem. Soc. Rev. 2017, 46, 7124–7144. [Google Scholar] [CrossRef]
  116. Zuo, J.; Chung, T.-S. Metal–organic framework-functionalized alumina membranes for vacuum membrane distillation. Water 2016, 8, 586. [Google Scholar] [CrossRef] [Green Version]
  117. Cheng, D.; Zhao, L.; Li, N.; Smith, S.J.; Wu, D.; Zhang, J.; Ng, D.; Wu, C.; Martinez, M.R.; Batten, M.P. Aluminum fumarate MOF/PVDF hollow fiber membrane for enhancement of water flux and thermal efficiency in direct contact membrane distillation. J. Membr. Sci. 2019, 588, 117204. [Google Scholar] [CrossRef]
  118. Yang, F.; Efome, J.E.; Rana, D.; Matsuura, T.; Lan, C. Metal–organic frameworks supported on nanofiber for desalination by direct contact membrane distillation. ACS Appl. Mater. Interfaces 2018, 10, 11251–11260. [Google Scholar] [CrossRef]
  119. Su, P.; Zhang, X.; Li, Y.; Chen, H.; Meng, Q.; Zhang, G. Distillation of alcohol/water solution in hybrid metal–organic framework hollow fibers. AIChE J. 2019, 65, e16693. [Google Scholar] [CrossRef]
  120. Efome, J.E.; Rana, D.; Matsuura, T.; Yang, F.; Cong, Y.; Lan, C.Q. Triple-layered nanofibrous metal-organic frameworks-based membranes for desalination by direct contact membrane distillation. ACS Sustain. Chem. Eng. 2020, 8, 6601–6610. [Google Scholar] [CrossRef]
  121. Hamzah, N.; Leo, C.; Ooi, B. Superhydrophobic PVDF/TiO2-SiO2 membrane with hierarchical roughness in membrane distillation for water recovery from phenolic rich solution containing surfactant. Chin. J. Polym. Sci. 2019, 37, 609–616. [Google Scholar] [CrossRef]
  122. Liao, Y.; Wang, R.; Fane, A.G. Engineering superhydrophobic surface on poly (vinylidene fluoride) nanofiber membranes for direct contact membrane distillation. J. Membr. Sci. 2013, 440, 77–87. [Google Scholar] [CrossRef]
  123. Hu, C.; Yang, Z.; Sun, Q.; Ni, Z.; Yan, G.; Wang, Z. Facile Preparation of a Superhydrophobic iPP Microporous Membrane with Micron-Submicron Hierarchical Structures for Membrane Distillation. Polymers 2020, 12, 962. [Google Scholar] [CrossRef] [Green Version]
  124. Liao, Y.; Loh, C.-H.; Wang, R.; Fane, A.G. Electrospun superhydrophobic membranes with unique structures for membrane distillation. ACS Appl. Mater. Interfaces 2014, 6, 16035–16048. [Google Scholar] [CrossRef] [PubMed]
  125. Lee, E.-J.; An, A.K.; Hadi, P.; Lee, S.; Woo, Y.C.; Shon, H.K. Advanced multi-nozzle electrospun functionalized titanium dioxide/polyvinylidene fluoride-co-hexafluoropropylene (TiO2/PVDF-HFP) composite membranes for direct contact membrane distillation. J. Membr. Sci. 2017, 524, 712–720. [Google Scholar] [CrossRef]
  126. Nthunya, L.N.; Gutierrez, L.; Verliefde, A.R.; Mhlanga, S.D. Enhanced flux in direct contact membrane distillation using superhydrophobic PVDF nanofibre membranes embedded with organically modified SiO2 nanoparticles. J. Chem. Technol. Biotechnol. 2019, 94, 2826–2837. [Google Scholar] [CrossRef]
  127. Fan, X.; Liu, Y.; Quan, X.; Zhao, H.; Chen, S.; Yi, G.; Du, L. High desalination permeability, wetting and fouling resistance on superhydrophobic carbon nanotube hollow fiber membrane under self-powered electrochemical assistance. J. Membr. Sci. 2016, 514, 501–509. [Google Scholar] [CrossRef]
  128. Tijing, L.D.; Woo, Y.C.; Shim, W.-G.; He, T.; Choi, J.-S.; Kim, S.-H.; Shon, H.K. Superhydrophobic nanofiber membrane containing carbon nanotubes for high-performance direct contact membrane distillation. J. Membr. Sci. 2016, 502, 158–170. [Google Scholar] [CrossRef]
  129. Guo, J.; Deka, B.J.; Kim, K.-J.; An, A.K. Regeneration of superhydrophobic TiO2 electrospun membranes in seawater desalination by water flushing in membrane distillation. Desalination 2019, 468, 114054. [Google Scholar] [CrossRef]
  130. Hou, D.; Lin, D.; Ding, C.; Wang, D.; Wang, J. Fabrication and characterization of electrospun superhydrophobic PVDF-HFP/SiNPs hybrid membrane for membrane distillation. Sep. Purif. Technol. 2017, 189, 82–89. [Google Scholar] [CrossRef]
  131. Zhang, W.; Li, Y.; Liu, J.; Li, B.; Wang, S. Fabrication of hierarchical poly (vinylidene fluoride) micro/nano-composite membrane with anti-fouling property for membrane distillation. J. Membr. Sci. 2017, 535, 258–267. [Google Scholar] [CrossRef]
  132. Salehi, S.; Jahanshahi, M.; Peyravi, M. Poly (vinylidene difluoride) Membrane Assisted by Modified ZnO/ZIF Nanoparticles for Membrane Distillation. Chem. Eng. Technol. 2018, 41, 1994–2004. [Google Scholar] [CrossRef]
  133. Ragunath, S.; Roy, S.; Mitra, S. Carbon nanotube immobilized membrane with controlled nanotube incorporation via phase inversion polymerization for membrane distillation based desalination. Sep. Purif. Technol. 2018, 194, 249–255. [Google Scholar] [CrossRef]
  134. Intrchom, W.; Roy, S.; Humoud, M.S.; Mitra, S. Immobilization of graphene oxide on the permeate side of a membrane distillation membrane to enhance flux. Membranes 2018, 8, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Singh, R.; Purkait, M.K. Role of poly (2-acrylamido-2-methyl-1-propanesulfonic acid) in the modification of polysulfone membranes for ultrafiltration. J. Appl. Polym. Sci. 2017, 134, 45290. [Google Scholar] [CrossRef]
  136. Singh, R.; Yadav, V.; Purkait, M.K. Cu2O photocatalyst modified antifouling polysulfone mixed matrix membrane for ultrafiltration of protein and visible light driven photocatalytic pharmaceutical removal. Sep. Purif. Technol. 2019, 212, 191–204. [Google Scholar] [CrossRef]
  137. Singh, R.; Purkait, M.K. Evaluation of mPEG effect on the hydrophilicity and antifouling nature of the PVDF-co-HFP flat sheet polymeric membranes for humic acid removal. J. Water Process Eng. 2016, 14, 9–18. [Google Scholar] [CrossRef]
Figure 1. Challenges of an ideal membrane distillation (MD) process in the field of water treatment and desalination.
Figure 1. Challenges of an ideal membrane distillation (MD) process in the field of water treatment and desalination.
Membranes 10 00140 g001
Figure 2. Survey on the number of published research papers since 2010. Database acquired from the advanced Scopus scholar search system with the terms “membrane distillation” and “nanomaterials” as in April 2020.
Figure 2. Survey on the number of published research papers since 2010. Database acquired from the advanced Scopus scholar search system with the terms “membrane distillation” and “nanomaterials” as in April 2020.
Membranes 10 00140 g002
Figure 3. Effectiveness of nanomaterials on the MD membrane.
Figure 3. Effectiveness of nanomaterials on the MD membrane.
Membranes 10 00140 g003
Figure 4. Basic membrane characteristics and their corresponding impacts that affect MD performance.
Figure 4. Basic membrane characteristics and their corresponding impacts that affect MD performance.
Membranes 10 00140 g004
Figure 5. Methodologies for the incorporation of nanomaterials onto an MD membrane.
Figure 5. Methodologies for the incorporation of nanomaterials onto an MD membrane.
Membranes 10 00140 g005
Figure 6. Recent advancements towards the application of MD in seawater desalination by utilizing various classes of new generation nanomaterials.
Figure 6. Recent advancements towards the application of MD in seawater desalination by utilizing various classes of new generation nanomaterials.
Membranes 10 00140 g006
Figure 7. Schematic diagram for the superhydrophobic modification of FTCS-TiO2-PVDF membranes.
Figure 7. Schematic diagram for the superhydrophobic modification of FTCS-TiO2-PVDF membranes.
Membranes 10 00140 g007
Figure 8. Illustration of air layer formation due to enhanced surface roughness.
Figure 8. Illustration of air layer formation due to enhanced surface roughness.
Membranes 10 00140 g008
Figure 9. Property enhancement in MD membranes due to addition of graphene oxide nanoparticles.
Figure 9. Property enhancement in MD membranes due to addition of graphene oxide nanoparticles.
Membranes 10 00140 g009
Figure 10. Demonstration of methodology to incorporate quantum dots (QDs) into a hydrophobic membrane.
Figure 10. Demonstration of methodology to incorporate quantum dots (QDs) into a hydrophobic membrane.
Membranes 10 00140 g010
Figure 11. Incorporation of metal organic frameworks (MOFs) by interfacial polymerization utilized in the MD process.
Figure 11. Incorporation of metal organic frameworks (MOFs) by interfacial polymerization utilized in the MD process.
Membranes 10 00140 g011
Figure 12. Summarization of serious issues related to the manufacture of nanomaterial-based polymeric MD membranes.
Figure 12. Summarization of serious issues related to the manufacture of nanomaterial-based polymeric MD membranes.
Membranes 10 00140 g012
Table 1. Overview of desired membrane characteristics for MD application.
Table 1. Overview of desired membrane characteristics for MD application.
CriteriaDescriptionDesired ValueRef.
Liquid Entry Pressure (LEPw)LEPw is the pressure required for the liquid to overcome the forces of hydrophobicity and penetrate the pores of the membrane. It is desired for the external pressure to be less than LEPw to allow the proper functioning of an MD system.
It is expressed using
LEPw = 2 B γ l   C o s θ r m a x
B is a geometric pore coefficient (equal to 1 for cylindrical pores), γ is liquid surface tension, θ is the contact angle, and rmax is the maximum pore radius.
LEPw > 250 kPa[47]
Mean Pore Size and Pore Size distributionPermeability depends on the mean pore size. Larger pore size allows a greater area for mass transfer, thereby increasing the overall membrane flux. However, increasing the pore size reduces the LEPw, hence it is necessary to find the optimum pore size to find a balance between LEPw and membrane permeability
Pore size distribution (PSD) indicates the variation in pore size and hence the variation in mass transfer and heat transfer mechanism with it, throughout the surface. Overall, PSD has a minimal effect on MD performance.
Mean Pore Size =
100 nm–1 μm
[48]
HydrophobicityHydrophobicity is a crucial aspect when the fabrication material for the membrane is chosen. It is quantified with respect to the contact angle (θCA) of water between the liquid surface and the membrane surface.θCA > 90° (Hydrophobic)
θCA > 150° (Superhydrophobic)
[49]
Chemical ResistanceThe material used for membrane fabrication must show good resistance to chemicals (acids, bases, surfactants) to prevent membrane fouling and consequent wetting.-[50]
Thermal ConductivityMembranes are desired to have a low thermal conductivity in the MD operation as it directly relates to the heat transfer through the membrane. Increased heat transfer would affect the vapor pressure equilibrium, thereby reducing the transmembrane flux.0.1–0.5 W m−1 K−1 is the range commonly observed in the literature[47]
Membrane ThicknessOptimum membrane thickness is required as it has major effects on the thermal conductivity and the membrane flux. Even though reducing the membrane thickness increases the membrane flux, it severely reduces the thermal resistance.30–60 μm[51]
Membrane Porosity
( ε )
Membrane porosity refers to the fraction of voids present in the membrane to the total volume of the membrane.
Increasing membrane porosity improves the flux transfer as well as the thermal resistance of the membrane; however, it is achieved at the expense of the mechanical strength of the membrane.
ε = 1   ρ m ρ p o l
where ρm and ρpol are the densities of the membrane and polymer, respectively
ɛ > 80%[4]
Tortuosity
( τ )
The irregularities of membrane pores from the ideal cylindrical pores are quantified by tortuosity. Highly tortuous structures result in lower flux as the vapor molecules suffer deviation from the direct path of transport.
τ = ( 2 ε ) 2 ε
1.1–3.9 has been observed for most MD systems[52]
Tensile StrengthThe membrane material should possess adequate tensile strength to be assembled and fixed in membrane modules as the operational pressures are much less compared with RO, UF, and MF.3.4–54.9 MPa is commonly observed for most MD membranes.[53]
Sliding AngleSliding angle is another criterion along with contact angle used to measure surface hydrophobicity.
Lower sliding angle indicates higher hydrophobicity as the water droplets do not adhere to the membrane surface.
<10°[54]
Surface RoughnessMicrostructure roughness results in the formation of air pockets which results in improving membrane hydrophobicity.Optimized surface roughness provides air layers which ultimately leads to higher hydrophobicity[55]
Table 2. State-of-the-art of nanomaterial-based MD membranes for improved desalination performance.
Table 2. State-of-the-art of nanomaterial-based MD membranes for improved desalination performance.
MembraneNanomaterialsMD TypeCategoryPore Size (µm)Flux
(L m−2 h−1)
Contact Angle (°)Ref.
PVDF-HFP/Si(NPs)SilicaDCMDMetalloid1.2848.6>150°[75]
PVDF- TiO2(NPs)Titanium dioxideDCMDMetal oxide0.4 ± 0.052.5140°[76]
FTCS-TiO2-PVDFTitanium dioxideDCMDMetal oxide0.4530163 ± 3°[77]
S-PVDF-20SilverUVMDMetallic0.4752.1148 ± 2.1°[78]
OMNI
(ZnO-GF)
Zinc oxideDCMDMetal oxide0.411.4 ± 0.9152.8 ± 1°[79]
FAS-SiNPs-SFMSilicaDCMDMetalloid0.8521.9 ± 1.2-[80]
PVDF-SiO2(NPs)Silicon dioxideVMDMetalloid0.142.894°[42]
PVDF-Al2O3(NPs)Aluminium oxideAGMDMetal oxide0.37020153°[43]
PVDF-M-CNTCarbon nanotubesDCMDCarbon0.1435.1 ± 0.7-[81]
PVDF-CNTsCarbon nanotubesVMDCarbon0.2028.5159°[82]
GNP-PolyetheneGrapheneDCMDCarbon0.1516.7123°[83]
GQDs-PVDFGraphene quantum dotsAGMDQuantum dots0.004917.6>125°[84]
MWCNTs/
SiO2-PVDF
Multi-walled carbon nanotubes and silicon dioxideVMDCarbon0.092.591 ± 2.1°[85]
Table 3. An overview of feasible methodologies to cast quantum dots (QDs)-doped hydrophobic polymeric membranes and the overall outcome.
Table 3. An overview of feasible methodologies to cast quantum dots (QDs)-doped hydrophobic polymeric membranes and the overall outcome.
MembraneQD TypeContact Angle (°)ApplicationRef.
(GQDs)/PVDFGraphene quantum dots>125°
  • Higher flux
  • Improved salt rejection
[84]
C18-CQDsCarbon quantum dots152.2 ± 1.25°
  • High water vapor permeability
  • High salt rejection rate (up to 99%)
  • Highly efficient oil–water separation (up to 99%)
[109]
GOQDs-PVDFGraphene oxide quantum dots34.3 ± 2.6°
  • Improved antibacterial and anti-biofouling properties
  • Lower flux drop
[108]
GOQDs-PSFGraphene oxide quantum dots65°
  • Enhanced permeability
  • Enhanced flux
  • Improved anti-fouling properties
[110]
PVAx-GOQD300Graphene oxide quantum dots53.8 ± 0.1°
  • Improved stability
[111]
Table 4. A brief summary of MOFs-doped MD membranes used in various MD modules.
Table 4. A brief summary of MOFs-doped MD membranes used in various MD modules.
MembraneMOFs TypeMembrane TypeMD ModuleContact Angle (°)LEPw (KPa)Ref.
MOF-functionalized alumina tubNH2-MIL-53(Al)Tubular(VMD)
Vacuum membrane distillation
-300[116]
ZIF-8/PDMSZIF-8Hollow fiber(DCMD)
Direct contact membrane distillation
130°-[119]
(Iron-BTC)/PVDFIron-BTCFlat sheet membrane(DCMD)
Direct contact membrane distillation
138.06 ± 2.18°82.73[118]
AlFu MOF/PVDFAlFuHollow fiber(DCMD)
Direct contact membrane distillation
>100°-[117]
MOFs/SiO2-PVDFMOF-808Flat sheet membrane(DCMD)
Direct contact membrane distillation
140.8°86.2 ± 3.2[120]
Table 5. Effect of nanoparticle addition on anti-wetting and anti-fouling behavior.
Table 5. Effect of nanoparticle addition on anti-wetting and anti-fouling behavior.
Base PolymerNanomaterialMode of FabricationConfigurationWater Contact Angle (°)LEPw (kPa)Mean Pore Diameter (µm)Performance CharacteristicsRef.
PVDFTiO2Phase inversionDCMD112 ± 1.464 ± 30.44 ± 0.02Self-cleaning effects under UV light with a higher flux recovery ratio as compared with unmodified PVDF[76]
PVDF-co-HPFFTES-functionalized TiO2Electrospinning followed by electrospraying to coat TiO2 NPsDCMD157 ± 1.6-0.52Mitigation of membrane fouling with regenerative abilities for long-term performance[129]
PVDF-HPFSiElectrospinningDCMD> 15076.41.70Total of 99.99% salt rejection over 240 h of desalination experiments showing long-term permeability[130]
PVDFPFOTS-modified SiO2Immersion depositionDCMD161.5-0.2 ± 0.01Steady operation over 156 h with feed consisting of NaCl (100 g/L), CaCl2 (1.26 g/L), and humic Acid (10 mg/L)[131]
PVDFFAS-modified SiO2 (8% wt)ElectrospinningVMD160.5 ± 2.31950.26 ± 0.02Pore wetting prevented due to high LEPw value showing a permeate flux of around 30 l. m−2h−1[88]
PVDFAluminum fumarate MOF (1%)Dry-jet wet phase inversionDCMD> 100>2000.3Stable salt rejection of 99.9% for 3.5 wt% NaCl solution over 50 h of operation[117]
PVDFZnO NPs modified by silane and coupled with ZIF-8 crystalPhase inversionDCMD70100-Modified membrane did not have a definite trend for permeate flux due to blocking of pores possibly due to wetting and scaling[132]
PVDFTriple-layered membrane with SiO2 (hydrophobic) blended in PVDF, PAN-MOFs, SiO2 (hydrophilic) blended in PVDFElectrospinningDCMD140.8 ± 9.986.20.31−1.22Hydrophobic SiO2-NPs increase the permeate flux while MOFs increase the pore size of the middle layer that contributed to superior DCMD performance for 5 h with low permeate conductivity[120]
PVDFCNTsElectrospinning followed by spray gun to coat CNTsVMD159.31880.2Even though the membrane had a stable performance for 14 h of operation, increasing the CNT loading beyond a point did not improve pure water flux due to the increased thickness of the membrane[133]
PTFEGODropwise coating of PVDF-GO onto flat sheet PTFE membranesDCMD75 ± 2-0.2Hydrophilic properties of GO improve mass transfer coefficient, thus improving membrane flux with stable performance for 60 days of operation[134]

Share and Cite

MDPI and ACS Style

Sinha Ray, S.; Singh Bakshi, H.; Dangayach, R.; Singh, R.; Deb, C.K.; Ganesapillai, M.; Chen, S.-S.; Purkait, M.K. Recent Developments in Nanomaterials-Modified Membranes for Improved Membrane Distillation Performance. Membranes 2020, 10, 140. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes10070140

AMA Style

Sinha Ray S, Singh Bakshi H, Dangayach R, Singh R, Deb CK, Ganesapillai M, Chen S-S, Purkait MK. Recent Developments in Nanomaterials-Modified Membranes for Improved Membrane Distillation Performance. Membranes. 2020; 10(7):140. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes10070140

Chicago/Turabian Style

Sinha Ray, Saikat, Harshdeep Singh Bakshi, Raghav Dangayach, Randeep Singh, Chinmoy Kanti Deb, Mahesh Ganesapillai, Shiao-Shing Chen, and Mihir Kumar Purkait. 2020. "Recent Developments in Nanomaterials-Modified Membranes for Improved Membrane Distillation Performance" Membranes 10, no. 7: 140. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes10070140

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