Open access peer-reviewed chapter

Forward Osmosis Membranes – A Review: Part I

Written By

Murat Eyvaz, Serkan Arslan, Derya İmer, Ebubekir Yüksel and İsmail Koyuncu

Submitted: 29 March 2017 Reviewed: 07 November 2017 Published: 28 March 2018

DOI: 10.5772/intechopen.72287

From the Edited Volume

Osmotically Driven Membrane Processes - Approach, Development and Current Status

Edited by Hongbo Du, Audie Thompson and Xinying Wang

Chapter metrics overview

3,147 Chapter Downloads

View Full Metrics

Abstract

Forward osmosis (FO) is a technical term describing the natural phenomenon of osmosis: the transport of water molecules across a semi-permeable membrane. The osmotic pressure difference is the driving force of water transport, as opposed to pressure-driven membrane processes. A concentrated draw solution (DS) with osmotic pressure draws water molecules from the feed solution (FS) through a semi-permeable membrane to the DS. The diluted DS is then reconcentrated to recycle the draw solutes as well as to produce purified water. As a major disadvantage, nature of FO membranes (asymmetrical structure) causes international concentration polarization (ICP) which promotes the decrease in water flux. Therefore, the number of studies related to improving both active and support layers of FO membranes is increasing in the applications. The purpose of the chapter is to bring an overview on the FO membrane manufacturing, characterizing and application area at laboratory or full scales. This chapter is published in two parts. In the first part, which appears here, the overview of membrane technologies and the definition of forward osmosis process are stated. The manufacturing methods of support and active layers forming FO membranes are described with common and/or new modification procedures.

Keywords

  • forward osmosis
  • water flux
  • reverse salt flux
  • active layer
  • support layer
  • thin film composite
  • structural parameter
  • porosity
  • internal concentration polarization

1. Introduction

Membrane separation processes have been widely applied for many years in environmental, industrial applications, and domestic use such as water/wastewater treatment, desalination, specific industrial purposes and energy recovery. Among the concentration-driven processes, FO has recently attained many attractions due to its advantages such as less energy requirement, lower fouling tendency or easier fouling removal and higher water recovery. However, all drawbacks of FO process such as; (i) membrane fouling originated from ICP, (ii) lower flux, and (iii) reverse salt diffusion (RSD) limit the performance of the FO applications in environmental studies. Moreover, “necessity of concentrate management” and “meeting of discharge standards with high amounts of product water” oblige the developing new processes, membrane materials or modifications and finding new DS. In recent studies, developing new FO membranes by optimization of thickness, porosity, tortuosity of active/support layer of FO membrane to increase water flux and decrease ICP are mainly focused.

FO is a technical term describing the natural phenomenon of osmosis: the transport of water molecules across a semi-permeable membrane. The osmotic pressure difference is the driving force of water transport, as opposed to pressure-driven membrane processes A concentrated DS with osmotic pressure draws water molecules from the FS through a semi-permeable membrane to the DS. The diluted DS is then reconcentrated to recycle the draw solutes as well as to produce purified water. As the driving force is only the osmotic pressure difference between two solutions which means that there is no need to apply an external energy, this results in low fouling propensity of membrane and minimization of irreversible cake forming which are the main problems controverted by membrane applications -especially- in biological treatment systems (e.g. FO-MBR). However, nature of FO membranes (asymmetrical structure) causes ICP which promotes the decrease in water flux. Therefore, the number of studies related to improving both active and support layers of FO membranes is increasing in the applications.

In this first part of chapter, advantages of FO over conventional membrane processes and main drawbacks originated from the nature of FO membranes are thoroughly stated by considering both review and research articles in the available literature. The book chapter consists of three main titles firstly including introduction section, the second of which states a literature survey on early definitions of diaphragm, membrane and selectivity phenomenon by considering about last two centuries. Basic principles of FO phenomenon is also expressed in this part. However, the special aspects of FO process are discussed in the third part in view of FO membrane properties. The water flow is mainly determined by the support layer, while the selectivity is by the active layer of FO membrane. Therefore, both support and active layer designing are overemphasized by addressing new materials, manufacturing methods and modification steps to overcome the main challenges of FO processes such as low water flux and concentration polarization phenomena causing the membrane fouling.

Advertisement

2. General aspects of membrane processes

2.1. Membrane technology

Systematic studies of the membrane phenomenon can be traced to the eighteenth century philosopher scientists. For example, Abb’e Nolet prepared an ‘osmosis’ word in 1748 to describe water permeability through a diaphragm. Through the nineteenth century and beginning of the twentieth century, membranes were not used for industrial or commercial purposes, but they were used as laboratory tools to develop physical/chemical theories. For example, the measurement of solution osmotic pressure by membranes by Traube and Pfeffer was used by van’t Hoff in 1887 to improve the limit law, which describes the behavior of ideal diluted solutions; this work led directly to the van’t Hoff equation. At the same time, Maxwell and others used a perfectly selective semipermeable membrane concept in the development of the kinetic theory of gases [1].

Early membrane researchers have experimented with all sorts of diaphragms available for themselves, such as pigs, cattle or fish and sausage covers made of animal guts. Later, nitrocellulose membranes were preferred because they could be reproducibly produced. In 1907, Bechhold developed a technique for preparing nitrocellulose membranes of a graded pore size determined by the bubble test [2]. Other early workers, the technique of Bechhold and were introduced into the market of microporous nitrocellulose membranes at the beginning of the 1930s [3, 4, 5]. In the following 20 years, this early microfiltration (MF) membrane technology has expanded to other polymers, especially cellulose acetate. Membranes found their first important practice in drinking water testing at the end of the Second World War. Drinking water sources serving large communities in Germany and elsewhere in Europe were destroyed and filters were urgently needed to test water safety. The research effort to develop these filters, backed by the US Army, was later exploited by Millipore Corporation, the first and largest producer of US MF membranes. By 1960, elements of modern membrane science were developed, but membranes were used only in a few laboratories and small, specialized industrial applications. An effective membrane industry was not available and the total annual sales of the membranes for all industrial applications probably did not exceed US $ 20 million in 2003. There were four problems that prevented membranes from being widely used as a separation process: they were too unreliable, too slow, too unselective and too expensive. Solutions for each of these problems have been developed over the past 30 years and membrane separation systems have become more common [1].

The first discovery of the conversion of membrane separation into an industrial process from a laboratory appeared with defect-free, high-flux anisotropic reverse osmosis (RO) membranes produced by the Loeb-Sourirajan process at the beginning of the 1960s [6]. These membranes consisted of an ultra-thin, selective surface on a much thicker but more permeable microporous support providing mechanical strength. The flux of the first Loeb-Sourirajan RO membrane was 10 times higher than that of any available membrane, and this performance made the RO potentially a practical method for desalinating water. Loeb and Sourirajan’s work and large-scale research and development by the US Department of Interior Office of Saline Water (OSW) have been a pioneer in the commercialization of RO and this has been a major factor in the development of MF and ultrafiltration (UF) membranes. In addition, the development of electrodialysis was supported by OSW funding. With the development of these industrial applications of membranes, the development of -especially- artificial kidneys, has been provided for medical separation procedures. Kolff et al. [7] demonstrated the first successful artificial kidney in Holland in 1945. It took about 20 years for technology to be applied to large-scale works, but these developments were completed in the early 1960s. Since then, the use of membranes in artificial organs has become an important life-saving procedure [1].

Currently, more than 800,000 people are protected with artificial kidneys and about 1 million of people who have undergone open heart surgery every year through a possible procedure by developing a membrane blood oxygenator. The sales of these devices easily exceed the total industrial membrane separation market. Another important medical application of membranes is for controlled drug delivery systems. An important figure in this area was Alex Zaffaroni, who founded Alza, a company dedicated to developing these products in 1966. Membrane techniques developed by Alza and his competitors are widely used in the pharmaceutical industry to improve efficacy and safety of drug delivery. Significant stages were recorded in the membrane technology of 1960–1980 period. Using the original Loeb-Sourirajan technique, other membrane production methods including interface polymerization and multilayer composite casting and coating were developed to produce high performance membranes. Membranes produced using these methods and containing thin layers of 0.1 μm or less are now produced by many companies. Along with the membrane type, membrane modules have been developed by working on the packing volume and the number of studies for increasing the membrane stability has increased. In the 1980s, large-scale installations involving MF, UF, RO and electrodialysis began to become widespread all over the world [1].

2.2. Conventional membrane processes

RO is primarily used to remove salts from brackish water or seawater while it can reject synthetic organic compounds. One of the latest developed membrane process, nanofiltration (NF), is used to soften fresh water and clear disinfection by-products (DBP) precursors. Electrodialysis is used to demineralize brackish and sea water and to soften fresh water. UF and MF are used to remove turbidity, pathogens and particles in fresh water. In the broadest sense, a membrane, a common element of all of these processes, can be defined as any barrier to the flow of suspended, colloidal or dissolved species in any solvent. Applicable size ranges for membrane processes are shown in Figure 1. Generally, the cost of membrane processing increases when the size of the solute is reduced. The ionic range in Figure 1 includes potable water solubles such as sodium, chloride, total hardness, maximum dissolved solids, and smaller DBP precursors. Macromolecular range includes large and small colloids, bacteria, viruses, and colors. The fine particle range includes particles that produce larger turbidity, most suspended solids, cysts, and larger bacteria. Membrane processes normally used in the ionic range can remove macromolecules and fine particles, but are not as cost effective as larger pore membranes due to some operational problems [9]. The comparison of the membrane properties with each other is given in Table 1.

Figure 1.

Membrane pore size compared with molecules, bacteria and virus [8].

Reverse osmosis Nanofiltration Ultrafiltration Microfiltration
Membrane Asymmetrical Asymmetrical Asymmetrical Symmetrical
Asymmetrical
Thickness
Thin film
150 μm
1 μm
150 μm
1 μm
150–250 μm
1 μm
10–150 μm
Pore size <0.002 μm <0.002 μm 0.2–0.02 μm 4–0.02 μm
Rejection of HMWC, LMWC
Sodium chloride
Glucose
Amino acids
HMWC
Mono-, di- and oligosaccharides
Polyvalent neg. ions
Macromolecules,
Proteins,
Polysaccharides
Vira
Particles
Clay
Bacteria
Membrane material(s) CA
Thin film
CA
Thin film
Ceramic
PSO, PVDF, CA
Thin film
Ceramic
PP, PSO, PVDF
Membrane module Tubular,
Spiral wound,
Plate-and-frame
Tubular,
Spiral wound,
Plate-and-frame
Tubular,
Hollow fiber
Spiral wound,
Plate-and-frame
Tubular,
Hollow fiber
Operating pressure 15–150 bar 5–35 bar 1–10 bar <2 bar

Table 1.

Comparing membrane processes [10].

RO is the tightest membrane process in liquid/liquid separation. In principle, water is the only substance passing through the membrane; essentially all dissolved and suspended material is rejected. The RO membranes with much larger pores are sometimes confused with NF membranes. True NF rejects ions with more than one negative charge, such as only sulfate or phosphate, while passing single charged ions. NF also rejects uncharged, dissolved materials and positively charged ions according to the molecular size and shape. Finally, the sodium chloride (NaCl) rejection varies from 0 to 50%, depending on NF and the rejected feed concentration. In contrast, “loose RO” is an RO membrane with reduced salt rejection. Such membranes are highly effective for a range of applications where moderate desalination is acceptable and, therefore, the operating pressure and power consumption are significantly reduced. Therefore, the costs are reduced in cases where complete desalination is not required.

UF is a process in which all low molecular weights compounds (LMWC) are freely permeable, while the high molecular weights compounds (HMWC), such as proteins, and suspended solids are rejected. Therefore, none of the mono- and di-saccharides, salts, amino acids, organics, inorganic acids or sodium hydroxide are rejected. Microfiltration (MF) is ideally a process where only suspended solids are rejected, and even proteins pass free through the membrane.

A wide range of products using membranes, but water desalination uses more than 80% of all membranes having ever been sold. The remaining 20% is used for -mostly- milk processing, while the rest is sold for use with many different liquids. Some liquids are waste products and some of which are very expensive pharmaceutical products. Table 2 lists some typical applications and as seen in the table the permeate as well as the concentrate can be the desired product [10].

Permeate Concentrate
RO Dyeing effluent Clean water BOD, salt, chemicals, waste products
Water Low salinity water Salty water
Whey Low BOD permeate Whey concentrate
Antibiotics Salty waste product Desalted, concentrated antibiotics
NF Dyeing effluent Clean,salty water BOD/COD, color
water softened water waste product
Whey Salty waste water Desalted whey concentrate
Antibiotics Clarified fermentation broth Waste product
Bio-gas waste Clarified liquid for discharge Microbes to be recycled
Carrageenan Waste product Concentrated carrageenan
Enzymes Waste product High value product
Milk Lactose solution Protein concentrate for cheese production
UF Oil emulsion Oil free water (<10 ppm) Highly concentrated oil emulsion
Washing effluent Clarified water Dirty water (waste product)
Water Clarified water Waste product
Whey Lactose solution Whey protein concentrate
Xantan Waste product Concentrated xantan

Table 2.

Type of membrane process for several products (the shaded area representing the main product) [10].

2.3. Forward osmosis process

van’t Hoff’s semipermeable membrane, which he assumed to promote the dilution of aqueous solution theory, is a permeable barrier to water (solvent), which is completely impermeable to dissolved solutes. For this reason, removal of the solvents results in a model barrier for all membrane filtration processes where the solutions are retained (concentrated). Like all joining properties, osmotic effects are limited to liquid solutions. Since we know nature is a watery system, the following solvent is water. When pure water and a random aqueous solution come in contact with the environment through a semi-permeable membrane, pure water is “drawn” into the solution, as if to dilute it: Osmosis. As is well known, osmosis is extremely important for the functioning of life when understood as a transport phenomenon at the molecular level. Live cell walls are osmotic barriers with improved selectivity towards inorganic and organic solutes (“biological membranes”). The direction of osmotic water transport, irrespective of the nature of the solution, indicates that the solution has a lower free energy (potential) than pure water. Specifically, the effectiveness of the solvent must be reduced by the effect of the solute, since the model barrier is assumed to communicate only through solvent [11].

Forward osmosis, an evolving separation/desalination process, has received increased interest in both academic research and industrial development in the past decade [12]. In FO, a semi-permeable membrane is placed between two solutions of different concentrations: a concentrated DS and a more dilute FS. Using the osmotic pressure differential to provide water permeation through the membrane, FO may respond to some of the deficiencies of hydraulic pressure driven membrane processes such as RO [13].

The transport of water molecules from a semipermeable membrane to the concentrate/saline solution on the other side of the membrane is referred to a technical term as forward osmosis. Contrary to conventional pressure-driven membrane processes, no pressure is applied to the side of water or concentrated solution. The difference in osmotic pressure between the aqueous medium on both sides of the membrane serves as the driving force to transport the water. [14]. Concentrated solution (DS), which pulls water molecules, is diluted during the process. The diluted DS is then re-concentrated in order to separate the water from the DS with a suitable further process. Where appropriate, the re-concentrated solution can be used again as DS. [15]. The FO process is shown in Figure 2. The main advantages of FO are it is not operated under any hydraulic pressure, that a wide range of pollutants can be rejected at a high level and have lower irreversible pollution than pressure-based membrane processes [17].

Figure 2.

Illustration of comparison between FO and RO processes [16].

As a method for water desalination, FO has been investigated for about four decades [18] and many researchers have found that (i) the selection or development of (new) membrane materials [19, 20], (ii) determining the suitable DS [21] The understanding of the mechanism of pollution [22], (iv) the characterization of concentration polarization (CP) [23]. In these pure academic publications, the FO survey and the increasing tendency of various special topics have been shown in Figures 3 and 4 for the last 10 years. As seen in Figure 3, the number of researchers has been steadily increasing, and recent research has focused heavily on membrane properties and development [16].

Figure 3.

Annual number of publications on FO since 2006 until fourth quarter of 2017 (retrieved from science direct database search) (updated and adapted from Eyvaz et al. [16]).

Figure 4.

The number of publication about FO studies different research topics since 2006 until fourth quarter of 2017 (retrieved from science direct database search) (updated and adapted from Eyvaz et al. [16]).

Advertisement

3. Manufacturing of forward osmosis membranes

Currently used membranes are mostly asymmetric porous membranes [24, 25]. In asymmetric porous membranes, the structure and transport properties change across the membrane thickness. An asymmetric membrane normally consists of a dense layer of 0.1–1 μm thick and supported by a highly porous, 100–200 μm thick support layer [24]. The dense layer provides most of the selectivity for the membrane. The separation properties are determined by the chemical structure, the size of the pores (0.4–1 nm) and the thickness of the skin layer. It is believed that the porous substrate provides mechanical support for the thin and fragile selective layer and has little effect on the separation performance of the membrane. However, recently the effects of the chemical properties of the support layer (eg, hydrophilicity/hydrophobicity) and pore structure (e.g., pore size and porosity) on composite membrane transport have been reevaluated [26, 27, 28].

In thin film composite (TFC) membranes, the porous support layer is generally an integrally skinned membrane formed by a non-solvent induced phase separation (NIPS) process. The skin layer is typically formed by either interfacial polymerization (IP) or dip coating followed by crosslinking [24]. The most common thin film chemistry for RO membranes is based on a completely aromatic polyamide (PA) formed by the IP of meta-phenylenediamine (MDP) and trimesoyl chloride (TMC). In contrast, the popular PA NF membranes are formed by IP of piperazine and TMC [29]. It is assumed that the dense selective layer formed by the IP is heterogeneous (20–200 nm) throughout the thickness and is highly cross-linked. The surface properties of a PA film are different from those of the PA dense layer because the polymer density is not uniformly distributed [30]. The PA dense layer is extremely negatively charged because acyl chloride groups are not completely converted to amide during the formation process; however, recent direct titration experiments have demonstrated the presence of both positive and negative fixed charges in the dense layer of composite PA NF membranes [31]. According to [28], Freger and Srebnik suggested that the fixed charge is not uniform and that the film is actually a “sandwich” with two oppositely charged layers [32].

The dense coating layer has been treated as a non-porous film in the past. More advanced identification techniques such as atomic force microscopy, scanning electron microscopy, transmission electron microscopy (TEM), X-ray photoelectron spectroscopy, electron spin resonance, nuclear magnetic resonance (NMR), small angle X-ray scattering, and molecular dynamics simulations have been used to state the structure of the dense layer. In the literature [28], a highly cross-linked PA skin layer structure with sub-nanoscale pores (0.2–1 nm) and low porosity has been reported [33, 34].

Wang et al. [28] stated that mixed matrix membranes contain both organic and inorganic phases. The first mixed matrix membranes were produced to enhance the performance of the gas separation membranes by providing interconnected flow paths of materials with a high diffusion rate [35]. In aqueous separations, mixed matrix membranes are typically formed of a polymer matrix in which inorganic particles are dispersed. Classically, micron-scale inorganic fillers (eg, zeolites and silicalite) have been added to polymer membranes to create preferential flow paths for rapid transport of certain molecules [36]. When nanomaterials (eg, metal and zeolites nanoparticles) are used as the inorganic filler, these membranes are called nanocomposite membranes [37].

Inorganic particles may be present throughout the thickness of a symmetric or integrally-skinned membrane or exclusively in the coating film of a composite membrane. Theoretically, mixed matrices add an additional degree of freedom to membrane production because the advantages of a particular filler material can be imbued into a bulk membrane material [35]. Mixed matrices have been used to enhance the mechanical and chemical stability of organic membranes and to add specific functionality to the interface of polymer membranes, such as desired degradation, reduced fouling or increased selectivity [38, 39]. Organic and inorganic hybrid membranes are very interested in using it as a new generation of membrane materials for water treatment. According to Wang et al. [28], scientists have begun to use nanoparticles TiO2 [40], carbon nanotubes [41], zeolites [42], clay [43], nonporous amorphous silica [44] and such as, to increase the water flux.

In another study [45] a new nanocomposite FO membrane is designed to perform oil/water separation and desalination at the same time. This nanocomposite FO membrane consists of an oil-repelling and salt-rejecting hydrogel selective layer on the top surface of graphene oxide (GO) nanosheets grafted into a polymeric support layer (Figure 5). This selective layer demonstrates strong underwater oleophobicity, which leads to superior anti-fouling properties under various oil/water emulsions and ICP can be decreased by and can be significantly reduced by GO in view of membrane structural parameter (S) decrease by about 20%. Compared to the commercial FO membrane, this new FO membrane has a markedly low membrane fouling tendency, having higher removal rates for oils and salts (>99.9% in oil and >99.7% in multivalent ions) for treatment of simulated shale gas wastewater (Figure 6). These combined advantages will endorse this new FO membrane in the treatment of highly saline and oily wastewater.

Figure 5.

Illustration of the synthetic process and work mechanisms of hydrogel/GO FO membrane [45].

Figure 6.

The study of simultaneously deoiling and desalting shale gas wastewaters [45].

Xu et al. [46] reported that the availability of suitable FO membranes is crucial for the development of FO technology. Problems such as high RSD, high concentration polarization (CP) and poor mechanical strength are frequently encountered in FO processes. Meanwhile, although FO tends to exhibit a lower membrane fouling than pressure driven membrane processes, fouling is still the most serious problem that adversely affects FO performance. To overcome these problems, many new FO membranes have been fabricated or ready-made membranes have been modified by means of surface chemistry in recent years [47].

Chung et al. [48] stated that a few comprehensive reviews on FO membrane development are available in the literature [12, 18, 49]. Basically, most FO membranes are fabricated with conventional phase inversion [49] and TFC by IP processes [50]. Each layer of FO membranes have been investigated, but reverse solute diffusions (RSD) tend to be high [51]. Employing hydrophilic materials as substrates in FO membranes is crucial to increase water flux [52]. Recently, TFC-FO-membranes synthesized on nanofiber [53, 54] and multi-bore [20] surfaces with good mechanical properties have also been reported. Future R & D should focus on innovative membranes with low fouling and ICP. Until recently, double-skinned FO membranes with dense RO skin and a loose RO skin, have been promised reduced membrane fouling and ICP [55].

Xu et al. [46] declared that the FO membrane serves as a selective barrier to control the water transport and solute retention to maintain the separation efficiency. The initial attempt to use the RO membrane in the FO process faced with some operational limitations; such as low flux, due to the thick sponge-like substrate and compact support of the RO membrane hindering mass transfer and causing severe ICP within the support layer [28, 29, 30]. Hydration Technologies, Inc. (HTI) developed the first commercial FO membranes [56], one of which has a characteristic structure embedding a thin polyester mesh support in cellulose triacetate (CTA) (Figure 7). These membranes provide significantly better separation performance than commercially available RO membranes. In addition to the commercial CTA FO membranes, HTI has then introduced the TFC FO membrane to the market; The flux of the spiral element was twice that of existing CTA membranes. This is thought to be a new criterion in future studies on FO membranes [58]. However, FO membranes with superior water permeability and salt rejection are still subjects to be developed for commercialization of FO technology. SEM images of the other some commercial FO membranes are presented in Figure 8.

Figure 7.

Cross-sectional SEM images of hydration technologies CA (the bar in each SEM image is 100 μm) (adapted from McCutcheon and Elimelech [57]).

Figure 8.

Cross-sectional SEM images for the (a) Osmonics CE, (b) Filmtec SW30 XLE and (c) Filmtec SW30 XLE, (d) Osmonics CE membrane with the fabric layer removed (the bar in each SEM image is 100 μm) (adapted from McCutcheon and Elimelech [58]).

According to the research in literature, it is shown that the adjustment of the sub-layers is of great potential in tailoring PA-TFC membranes. In addition, the flexibility of PA-TFC membrane structures has a positive effect on the improvement of the sub-layers, since each of the surface active and support layers can be individually constituted for a particular purpose (Figure 9) [59].

Figure 9.

Flexibility in tailoring PA-TFC membranes by adjusting the surface active and support layers individually [59].

3.1. Support layer properties and manufacturing techniques

Li et al. [60] have recently reviewed recent researches on polymer and polymer composite membranes for RO and FO processes comprehensively. In one of these studies, a TFC FO flat membrane has a thin selective layer on top of a flat porous polymeric support that is produced by phase inversion with/without a thin nonwoven layer [50]. More recently, nanofiber mats with high porosity have been proposed as a support layer to reduce the ICP to a minimum [53, 54, 61]. Bucky-papers made from CNTs are also being tested as support layer candidates due to their flexibility, strength and high porosity, it is also recommended to investigate other low cost and high porosity materials such as metal oxide nanotubes [62]. Parallel to the studies with TFC flat plate modules, the number of research related to hollow fiber configuration is also increasing due to its advantages such as high packing density and enhanced flow pattern and self-supporting structure [20, 63]. Under the same drive force effect, higher fluxes were obtained with hollow fiber membranes than with flat sheet membranes [64].

Similar to the characteristics of the support layers required to produce high performance PA TFC RO membranes, it is desirable that the support layers of the TFC FO membranes have high hydrophilicity, stability and mechanical strength, [65]. In addition, resistance to chemicals, temperature and oxidation, as well as low fouling tendencies, increase the potential use of FO membranes in harsh industrial environmental conditions. [66]. However, the thin support layer, high porosity, and low tortuosity will help reduce ICP [67]. For this reason, some of the research related to FO membranes have focused on support layer fabrication and modification. In these studies, fabrication parameters such as optimization of polymer concentration, solvent composition and functional additives have been considered in the synthesis of the support layer [68]. For example, a simultaneous casting of two polymer solutions with a co-casting technique, a synthesized support layer has played an important role in reducing ICP, improving water flux and reducing reverse salt flux [69]. Furthermore, even the selection of non-woven fabrics underneath the polymeric support layer significantly affects the adhesion of the support layer polymer to this non-woven sheet. For example, selecting a fabric with high tortuosity, large thickness, and low porosity leads to an undesirable decrease in the water flux of the FO membrane. Moreover, the addition of foreign components to the substrate casting solution can help improve the substrate properties. For example, when lignin content is incorporated into the polysulfone (PSf) substrate, the bulk porosity enhanced, shorter diffusion pathway is provided and TFC membrane performance is improved [70]. By using diethylene glycol as a pore forming agent in the PSf/sulfonated poly (ether ketone)/N-methylpyrrolidone (NMP) casting solution, greater porosity and wider pore size distribution were obtained which reduced the resistance of the support layer to the solution and the reverse salt flux was also relatively controlled [65]. Addition of PEG to the preparation of the CAP substrate increased the connectivity of the pores and prevented macrovoids, as well as positively interacted with the cast glass blade. The resulting support is suitable for high performance TFC FO membrane fabrication since it has a high porosity bottom and a tight top surface [52]. The result is that the TFC membrane with macrovoids (or finger-like structure) support facilitates mass transport and reduces ICP in FO [67]. However, these porous structures may become mechanically weak points in the membrane structure and may, in practice, worsen membrane separation under continuous water flow or backwash conditions [71]. The highly porous support can also increase the difficulty in forming an excellent active selective layer with the necessary mass transport properties [67]. However, the sponge-like structure of small pores surrounded by dense walls may be convenient to form an integral thin active layer and may exhibit better mechanical stability on the finger-like property; however, it increases resistance to mass transfer [64]. Experimental studies suggest that the ideal support with a sponge-like film on a finger-like substrate is very important in the fabrication of high-performance TFC FO membranes [50].

Recently, nanofiber support layers with scaffold-like and interconnected porous structures have been seen as promising alternatives to overcoming the disadvantages of sponge-like structures. The nanofiber supported PA TFC membranes exhibited much lower S (~80 μm) than a commercial HTI FO membrane (S = 620 μm) and thus with a low molarity (0.5 M NaCl) DS and a DI water as the FS, it has been observed that the water flux has increased by five times [72].

An FO membrane with tubular nanofiber support was manufactured for the first time in the study of Arslan et al. [53]. In the first stage of the manufacturing, the support layer (polyacrylonitrile (PAN) nanofiber) was coated on the hollow braided rope (backing layer) by electrospinning method. In the second step, the active layer called the TFC layer was coated on the formed nanofiber by the IP process. Schematic illustration of FO membrane manufacturing is shown in Figure 10 and SEM images are presented in Figure 11. The TFC layer is the main selective barrier that prevents the transfer of the salt to the diluted side and allows the water molecules to diffusion into the DS side.

Figure 10.

Demonstration of tubular nanofiber FO membrane manufacturing steps: (a) Steps of TFC process, 1— immersion of the tubular nanofiber membrane into the MPD solution, 2—air drying, 3—immersion of the tubular nanofiber membrane into the TMC solution, 4—heat treatment, 5— tubular nanofiber FO membrane. (b) Reaction mechanisms of the polyamide formation from MPD and TMC. (c) Schematic representation of cross-section of tubular nanofiber FO membrane [54].

Figure 11.

SEM images of (a) nanofibers in support layer at magnification of 10,000×, (b) gel-like formation of the polyamide layer on nanofibers at magnification of 50,000×, (c) cross-section of tubular nanofiber supported FO membrane at magnification of 10,000× (adapted from Arslan et al. [53]).

According to aforementioned review [60], Han et al. [65] pointed out that the hydrophilicity and support layer thickness are critical parameters in controlling of water transport. It is reported that the TFC membrane with support layer which is completely sponge-like and has a hydrophilic upper surface, provides a higher water flux than a TFC membrane with support layer which is completely hydrophobic and has finger-like structure. In order to prepare or modify the support layer, hydrophilic materials such as sulfonated polysulfone (sPSf), sulfonated copolymer made of polyethersulfone (PES) and polyphenylsulfone (PESU-co-sPPSU), sulfonated poly(ether ketone), poly- dopamine (PDA) or poly(vinyl alcohol) (PVA) have been explored [73, 74, 75]. Emadzadeh et al. improved the mass transfer and reduced ICP by applying TFC on the PSF support layer containing TiO2, thus increasing the water flow of the FO membrane [76].

Liang et al. [77] for the first time in the production of TFC-FO membranes, it has been proposed to use vertical porous substrates as a support layer. The addition of acetone in aqueous phase promotes IP on vertical porous substrates. Positron annihilation lifetime analyses indicated that new FO membranes in the study have thicker and dense selective layers than conventional FO membranes with asymmetric substrates. These new FO membranes have a low structural parameter, indicating a greatly reduced ICP effect.

The selective layer exhibits an unprecedented water flux up to 93.6 L/m2.h (LMH) (Figure 12) when driven by a 2 M NaCl as DS. This performance is evidenced by the FO membranes reported in the literature and commercially available. The authors suggested according to the results that substrates with vertically oriented porous structure are ideal supports for developing FO membranes with lower ICP and ultra-high water flux. Proposed lower ICP in vertically oriented porous layer compared with a tortuous sponge-like structure in FO membranes by means of salt concentration profile are illustrated in Figure 13 [77].

Figure 12.

(a) Water flux and (b) reverse salt flux of the FO membranes in a process mode of the selective layer facing the FS (VOPS: vertically oriented porous substrates) [77].

Figure 13.

The salt concentration profiles of (a) VOPS-TFC and (b) PI-TFC FO membranes. πD and πF denote the osmotic pressure of draw solution and FS, respectively. Δπideal indicates the osmotic pressure difference between the bulk feed and the bulk draw solution. Δπeff means the effective osmotic pressure driving force due to the presence of ICP effect [77].

In a more recent study, Kwon et al. [78] produced a highly permeable and mechanically resistant TFC-FO membrane with a new support layer which has been already commercialized porous polyethylene (PE) membrane as the lithium ion battery separator. The very open and interconnected pore structure of the PE support, when combined with the thickness (~8 μm), is useful for alleviating the ICP, thus increasing the FO water flux. The use of a suitable plasma treatment and a surfactant in the PE support resulted in a stable formation of a PA permselective layer on the support by IP process. The prepared PE supported TFC (PE-TFC) membrane exhibited high water flux and low reverse salt flux performance due to its significantly low structural parameter. The performance values obtained in this study are also compared with other flux values in the literature (Table 3). The PE-TFC membrane has superior mechanical properties compared to the much thicker commercial FO membrane due to the exceptionally high mechanical integrity of the PE support. The proposed strategy offers a new material platform for FO membranes with strong commercial potential and excellent performance and durability.

Membranes S (μm) Jw (LMH) Js/Jw (g/L)
(FO/PRO)
Refs.
PE-TFC 161 41.9/45.1 0.50/0.49 [78]
PES hollow fiber-TFC 219 26.5/37.6 0.17/0.14 [64]
PAN flat sheet-TFC 350 28.8/36.3 0.10/0.13 [79]
Cellulose ester flat sheet-TFC 32 56.9/89.5 0.14/0.12 [52]
PESU-co-sPPSU flat sheet-TFC 324 20.0/25.0 0.12/0.13 [80]
PTA-co-POD flat sheet-TFC 236 37.5/78.4 0.15/0.15 [81]
PSf/SPPO (50 wt% SPEK) flat sheet-TFC 381 16.0/32.0 0.28/0.19 [82]
PSU/SPEK (50 wt% SPEK) flat sheet-TFC 107 23.0/34.0 0.18/0.20 [65]
PES/SPES (50 wt% SPES) flat sheet-TFC 245 25.2/33.7 0.28/0.26 [83]
Polydopamine-coated PAI flat sheet-TFC 456 14.0/48.0 0.44/0.17 [84]
PSf-silica NPs flat sheet-TFC 216 31.0/60.5 0.24/0.26 [85]
PSf-zeolite flat sheet-TFC 340 33.0/65 0.55/0.47 [86]
PSf-LDH flat sheet-TFC 148 18.1/34.6 0.45/0.36 [87]
PES nanofiber-TFC 106 46.0/50.0 [72]
Nylon 6,6 nanofiber-TFC 21.0/27.0 0.24/0.44 [88]
PVDF nanofiber-TFC 193 22.0/31.0 0.17/0.43 [89]

Table 3.

Comparison of the FO performance of RO-grade membranes (FS: DI water, DS: 1.0 M NaCl) [78].

In another recent study on the support layer, Zhang et al. [90] manufactured hollow fiber FO membranes with improved thermal stability using IP process on the lumen side of the co-poly (phthalazinone biphenyl ether sulfone) (PPBES) substrate. The increase in water flow in the PPBES substrate also increased the flow in the FO membrane. IP preparation parameters such as solvent, monomer concentrations, reaction time and curing conditions have been shown to seriously affect the development of composite FO membrane properties. The water flux of composite FO membranes increased from 24.0 to 66.5 LMH without a significant change in salt flux/water flux (Js/Jw) ratio when the draw solution temperature was raised from 23 to 85°C (Figure 14).

Figure 14.

Water flux changing of thermally-stabile FO membrane with draw solution temperature [90].

3.2. Active layer properties and manufacturing techniques

Li et al. [60] in their comprehensive review stated some recent applications on manufacturing or modifying active layer of FO membranes. Accordingly, the preparation of PA TF FO membranes is also similar to the preparation of TFC RO membranes. It is necessary to optimize the parameters such as the reaction time and the air drying duration and compositions of the monomers. Klaysom et al. [91] noted that both the surfactant additive and the drying of excess amine solution prior to the reaction are two critical parameters in PAN support preparation to control membrane properties. The addition of SDS increases polymerization and helps to form a uniform and highly cross-linked PA film. Thus, the rate of salt uptake in this study increased from 57% to over 95%, and the decrease in permeability did not occur. On the other hand, the removal of the excess amine solution before interacting with TMC resulted in the formation of a less rough membrane with improved salt rejection [91]. Due to the ionic interaction between cetyltrimethylammonium chloride (CTAC) and m-phenylenediamine (MPD) in the aqueous solution, CTAC may alter the reaction of the monomers of the presence and polymer molecular aggregation. Increasing the CTAC content improves the formation of the linear PA structure and microcrystalline structure of the active layer, but consequently the water flux of the PA TFC hollow fiber membrane with PES support layer is reduced, despite the high reverse salt selectivity [92]. Thermal annealing after SDS/glycerol treatment on TFC FO membranes facilitates the removal of residual unreacted monomers from the surface of the active layer, increasing the free volume size/fractional free volume ratio and reducing the total membrane thickness; so that the water flux can be improved without losing the rejection performance of the membrane [52]. Another major problem encountered in FO processes is membrane fouling, although it is less severe and reversible compared to RO processes. The structures of the support layers also significantly affect the active layer properties and hence the fouling characteristics of the TFC FO membranes. Surfaces with high roughness and large leaf-like structures are more prone to foulant accumulation and exhibit a dramatic decline in flux through these membranes, making it more difficult to improve the flux by physical cleaning of the membrane [93]. When the TFC FO membrane surface is modified, for example by covalent attachment of PEG, the tendency of surface contamination is significantly reduced due to surface barriers that adsorb pollutants [74]. On the other hand, attachment of the functionalized silica on TFC membrane via covalent amide bonds between amine groups of functionalized nanoparticles and the carboxyl groups of the TFC surface improve the fouling resistance and reduce the BSA or alginate adhesion. This is explained by the presence of the tightly bound hydration layer and the reduction of the charged carboxyl groups on the TFC membrane surface [94].

A more recent research in the literature has produced nanoporous thin-film inorganic (TFI) FO membranes with a tetraethylorthosilicate-driven sol-gel process (Figure 15(a)). The produced membrane was used for the removal of four typical ionic divalent heavy metals. In laboratory scale FO process, 69 LMH flux was obtained by using four heavy metal containing FS at pH 4.5 and 2 M NaCl as DS (Figure 15(b)). An average of 94% metal removal from the 200 mg/L FS solution was obtained. Since the hydrated ion diameters of the metals are smaller than the membrane pore size, the charge-interaction should be responsible for heavy metal rejection. Based on the classical Debye-Hückel theory and the Gouy-Chapman model, You et al. [95] have shown the importance of double-layer overlap in the membrane pore induced by electrostatic interaction between heavy metal ions and silica-made pore walls. Thus, the selectivity of the TFI membrane depends primarily on the function of the membrane pore size, the surface potential of the membrane pore wall, and Debye length (Figure 16). This study not only confirms the feasibility of the TFI membrane in the treatment of acidic heavy metal wastewater without pH adjustment, but it also suggests a simple theoretical scheme for better understanding and design of the charged membrane for FO applications.

Figure 15.

(a) Schematic diagram of TFI membrane formation (SSM: Stainless steel mesh, TEOS: Tetraethylorthosilicate), (b) FO water flux obtained under AL-FS and AL-DS mode at initial heavy metal concentration of 200 mg/L and pH 4.5 ± 0.5 for TFI membrane (adapted from You et al. [95]).

Figure 16.

Schematic illustration of proposed mechanisms for rejection of heavy metal ions in FO process based on charge exclusion effect within the pores of TFI membrane [95].

Salehi et al. [96] in their work, fabricated a new and highly efficient FO membrane by using electrostatic interaction on a porous support layer employing layer-by-layer (LbL) assembly of positive chitosan (CS) and negative GO nanosheets. The support layer was prepared by mixing the hydrophilic sulfonated polyethersulfone (sPES) and PES using the wet phase inversion process (Figure 17).

Figure 17.

Schematic illustration of CS/GO LbL assembly procedure [96].

Various characterization techniques have been used to confirm that the LbL membrane has been successfully fabricated. The number of layers formed in the SPES-PES support layer was easily adjusted by repeating the CS and GO deposition cycles. A TFC membrane with the same SPES-PES support layer and PA active layer was also prepared to compare membrane performances. Water permeability and salt rejection of the fabricated membranes were obtained with two types of DS (including Na2SO4 and sucrose) for two different membrane orientations. The results showed that the membrane coated by a CS/GO double layer had a flow rate of 2–4 orders of magnitude as much as the TFC. By increasing the number of CS/GO double layers, the selectivity of the LbL membrane was improved. The newly fabricated LbL membrane showed better fouling resistance than the TFC in the FS containing 200 ppm sodium alginate as the foulant model (Figure 18).

Figure 18.

Fouling behavior and flux recovery of the TFC and 10-LbL membranes [96].

Xu et al. [60] reported that generally, the flux obtained in the active layer facing draw solution (AL-DS) (PRO) configuration is higher than in the active layer facing feed solution (AL-FS) (FO) mode, but more fouling may occur in the PRO mode if the FS containing scalants/foulants is easily transported to the porous support layer. Two active layered hollow fiber membranes, one at the top of the high porosity support layer and one at the bottom, have been proposed by Wang’s group [97] so that scaling or fouling can be controlled without reducing water flux in the AL-DS mode. The hollow fiber membranes with RO and NF-like scales fabricated on a PAI support were subjected to high water flux and reverse salt flux values (41.3 and 5.2 LMH) using 2 M NaCl DS and DI as FS in AL-DS mode after IP reaction and polyethyleneimine (PEI) modification. In addition, the presence of the NF-like layer on the support layer can greatly increase the resistance to scaling in the AL-DS mode. A double-skinned hollow fiber membrane with CaHPO4 scaling with a 2-hour backwash recovers 96% of the water flux while a hollow fiber membrane with a single RO selective layer has a recovery of 78% [97]. Recent studies have used polyelectrolyte LbL to form an NF-like skin in the support layer studies conducted without chemical modification, while PA-RO-like layers have also been formed. Since the resulting NF-like skin does not directly contact the FS and support layer, it prevents the transport of pollutants such as humic acid, dextran and lysozyme and thus the pore clogging. As a result, for a double-skinned hollow fiber membrane, the decrease in water flux was less than 30% during 4 hours operation, whereas for RO layer hollow fiber membranes, this reduction was 30–40% when 200 ppm foulant was used [98].

Advertisement

4. Conclusions

The active layer of an ideal FO membrane must be very thin and dense to achieve high salt retention. In order for the membrane to be able to be operated for a long time and the internal concentration polarization to be low, the support layer should be thin, hydrophilic, porous and exhibit mechanical strength as possible. The hydrophilicity must be high so that high flux and low fouling can be achieved. According to the current studies, utilizing novel nanomaterials, substrates and layer-by-layer assumptions in manufacturing of FO membrane undoubtedly enhance the water flux, rejection of the pollutants and minimize the membrane fouling but using synthetic wastewater -generally- containing one model foulant or DI water as feed solution makes it difficult to predict how FO membranes will act in real wastewaters or harsh environmental conditions. Therefore working with complex foulants and real wastewaters to better understanding of membrane behaviors, using modeling tools for fouling prediction and new cleaning strategies are essential to mitigate intrinsic challenges of the FO membranes.

In on-going researches, the developed new support layers appears continue to increase water flux slightly, however, lower water flux remains as a main challenge of the process when compared the conventional membrane systems. It is also a fact that the diffusion provided by draw solution in the process is not effective alone to increase product water volume, therefore, some promotive factors such as rehabilitated hydrodynamic behaviors or simultaneous filtration could be provided together with diffusion phenomena in further researches.

Advertisement

Acknowledgments

This work was supported by the Scientific and Technological Research Council of Turkey (TUBITAK), grant number: CAYDAG-113Y340.

References

  1. 1. Baker RW. Membrane Technology and Applications. 2nd ed. UK: John Wiley & Sons, Ltd.; 2004. 538 p. DOI: 10.1002/0470020393
  2. 2. Bechhold H. Kolloidstudien mit der Filtrationsmethode. Zeitschrift für Physikalische Chemie. 1907;60:527-533. DOI: 10.1002/bbpc.19070133207
  3. 3. Elford WJ. Principles governing the preparation of membranes having graded porosities. The properties of ‘Gradocol’ membranes as Ultrafilters. Transactions of the Faraday Society. 1937;33:1094-1104. DOI: 10.1039/TF9373301094
  4. 4. Zsigmondy R, Bachmann W. Uber Neue Filter. Zeitschrift für Anorganische und Allgemeine Chemie. 1918;103:119-128. DOI: 10.1002/zaac.19181030107
  5. 5. Ferry JD. Ultrafilter membranes and ultrafiltration. Chemical Reviews. 1936;18:373-455. DOI: 10.1021/cr60061a001
  6. 6. Loeb S, Sourirajan S. Sea water demineralization by means of an osmotic membrane. In: Gould RF, editor. Saline Water Conversion–II, Advances in Chemistry Series Number 28. Washington, DC: American Chemical Society; 1963. pp. 117-132. DOI: 10.1021/ba-1963-0038.ch009
  7. 7. Kolff WJ, Berk HTHJ, Welle NM, van der Ley AJW, van Dijk EC, van Noordwijk J. The artificial kidney: A dialyser with great area. Acta Medica Scandinavica. 1944;117:121-134. DOI: 10.1111/j.0954-6820.1944.tb03951.x
  8. 8. McGowan W. Water Processing: Residential, Commercial, Light Industrial. 3rd ed. Water Quality Association: Illinois; 2001
  9. 9. Taylor JS, Wiesner M. Membranes. In: Letterman RD, editor. Water Quality and Treatment: A Handbook of Community Water Supplies. 5th ed. New York: AWWA-McGraw-Hill; 1999. pp. 11.1-11.71
  10. 10. Wagner J. Membrane Filtration Handbook Practical Tips and Hints. 2nd ed. Minnetonka, MN: Osmonics Inc.; 2001. p. 127
  11. 11. Böddeker KW. Liquid Separations with Membranes: An Introduction to Barrier Interference. Berlin Heidelberg: Springer; 2007. p. 146
  12. 12. Zhao S, Zou L, Tang CY, Mulcahy D. Recent developments in forward osmosis: Opportunities and challenges. Journal of Membrane Science. 2012;396:1-21. DOI: 10.1016/j.memsci.2011.12.023
  13. 13. Shaffer DL, Yip NY, Gilron J. Elimelech M. Seawater desalination for agriculture by integrated forward and reverse osmosis: improved product water quality for potentially less energy. Journal of Membrane Science. 2012;415-416:1-8. DOI: doi.org/10.1016/j.memsci.2012.05.016
  14. 14. Lutchmiah K, Lauber L, Roest K, Harmsen DJH, Post JW, Rietveld LC, van Lier JB, Cornelissen ER. Zwitter ions as alternative draw solutions in forward osmosis for application in wastewater reclamation. Journal of Membrane Science. 2014;460:82-90. DOI: 10.1016/j.memsci.2014.02.032
  15. 15. Liu Y, Mi B. Combined fouling of forward osmosis membranes: Synergistic foulant interaction and direct observation of fouling layer formation. Journal of Membrane Science. 2012;407-408:136-144. DOI: 10.1016/j.memsci.2012.03.028
  16. 16. Eyvaz M, Aslan T, Arslan S, Yuksel E, Koyuncu İ. Recent developments in forward osmosis membrane bioreactors: A comprehensive review. Desalination and Water Treatment. 2016;57(59):28610-28645. DOI: 10.1080/19443994.2016.1193448
  17. 17. Holloway RW, Childress AE, Dennett KE Cath TY. Forward osmosis for concentration of anaerobic digester centrate. Water Research. 2007;41:4005-4014. DOI: 10.1016/j.watres.2007.05.054
  18. 18. Alsvik IL, Hägg MB. Pressure retarded osmosis and forward osmosis membranes: Materials and methods. Polymer. 2013;5:303-327. DOI: 10.3390/polym5010303
  19. 19. Flanagan MF, Escobar IC. Novel charged and hydrophilized polybenzimidazole (PBI) membranes for forward osmosis. Journal of Membrane Science. 2013;434:85-92. DOI: 10.1016/j.memsci.2013.01.039
  20. 20. Luo L, Wang P, Zhang S, Han G, Chung TS. Novel thin-film composite tri-bore hollow fiber membrane fabrication for forward osmosis. Journal of Membrane Science. 2014;461:28-38. DOI: 10.1016/j.memsci.2014.03.007
  21. 21. Ou R, Wang Y, Wang H, Xu T. Thermo-sensitive polyelectrolytes as draw solutions in forward osmosis process. Desalination. 2013;318:48-55. DOI: 10.1016/j.desal.2013.03.022
  22. 22. Linares RV, Yangali-Quintanilla V, Li Z, Amy G. NOM and TEP fouling of a forward osmosis (FO) membrane: Foulant identification and cleaning. Journal of Membrane Science. 2012;421-422:217-224. DOI: 10.1016/j.memsci.2012.07.019
  23. 23. Tan CH, Ng HY. Revised external and internal concentration polarization models to improve flux prediction in forward osmosis process. Desalination. 2013;309:125-140. DOI: 10.1016/j.desal.2012.09.022
  24. 24. Mulder M. Basic Principles of Membrane Technology. London: Kluwer Academic Publishers; 1996. p. 564
  25. 25. Loeb S, Sourirajan S. High Flow Semipermeable Membrane for Separation of Water from Saline Solutions, US Patent. 3,133,132. May 12, 1964
  26. 26. Ghosh AK, Hoek EMV. Impacts of support membrane structure and chemistry on polyamide-polysulfone interfacial composite membranes. Journal of Membrane Science. 2009;336:140-148. DOI: 10.1016/j.memsci.2009.03.024
  27. 27. Pendergast MM, Hoek EMV. A review of water treatment membrane nanotechnologies. Energy & Environmental Science. 2011;4:1946-1971. DOI: 10.1039/C0EE00541J
  28. 28. Wang J, Dlamini DS, Mishra AK, Pendergast MTM, Wong MCY, Mamba BB, Freger V, Verliefde ARD, Hoek EMV. A critical review of transport through osmotic membranes. Journal of Membrane Science. 2014;454:516-537. DOI: 10.1016/j.memsci.2013.12.034
  29. 29. Cadotte JE, Petersen RJ, Larson RE, Erickson EE. New thin-film composite seawater reverse-osmosis membrane. Desalination. 1980;32:25-31. DOI: 10.1016/S0011-9164(00)86003-8
  30. 30. Freger V. Nanoscale heterogeneity of polyamide membranes formed by interfacial polymerization. Langmuir. 2003;19:4791-4797. DOI: 10.1021/la020920q
  31. 31. Coronell O, Marinas BJ, Cahill DG. Depth heterogeneity of fully aromatic polyamide active layers in reverse osmosis and nanofiltration membranes. Environmental Science and Technology. 2011;45:4513-4520. DOI: 10.1021/es200007h
  32. 32. Freger V, Srebnik S. Mathematical model of charge and density distributions in interfacial polymerization of thin films. Journal of Applied Polymer Science. 2003;88:1162-1169. DOI: 10.1002/app.11716
  33. 33. Hilal N, Al-Zoubi H, Darwish NA, Mohammad AW. Characterisation of nanofiltration membranes using atomic force microscopy. Desalination. 2005;177:187-199. DOI: 10.1016/j.desal.2004.12.008
  34. 34. Boussu K, De Baerdemaeker J, Dauwe C, Weber M, Lynn KG, Depla D, Aldea S, Vankelecom IFJ, Vandecasteele C, Van der Bruggen B. Physico-chemical characterization of nanofiltration membranes. Chemphyschem. 2007;8:370-379. DOI: 10.1002/cphc.200600512
  35. 35. Zimmerman C, Singh A, Koros W. Tailoring mixed matrix composite membranes for gas separations. Journal of Membrane Science. 1997;137:145-154. DOI: 10.1016/S0376-7388(97)00194-4
  36. 36. Jia MD, Peinemann KV, Behling RD. Molecular-sieving effect of the zeolite-filled silicone-rubber membranes in gas permeation. Journal of Membrane Science. 1991;57:289-296. DOI: 10.1016/S0376-7388(00)80684-5
  37. 37. Brunet L, Lyon DL, Zodrow K, Rouch JC, Caussat B, Serp P, Remigy JC, Wiesner MR, Alvarez PJJ. Properties of membranes containing semi-dispersed carbon nanotubes. Environmental Engineering Science. 2008;25:565-575. DOI: 10.1089/ees.2007.0076
  38. 38. Aerts P, Greenberg AR, Leysen R, Krantz WB, Reinsch VE, Jacobs PA. The influence of filler concentration on the compaction and filtration properties of Zirfon (R)-composite ultrafiltration membranes. Separation and Purification Technology. 2001;22-23(1-3):663-669. DOI: 10.1016/S1383-5866(00)00165-9
  39. 39. Wu L, Shamsuzzoha M, Ritchie SMC. Preparation of cellulose acetate supported zero-valent iron nanoparticles for the dechlorination of trichloroethylene in water. Journal of Nanoparticle Research. 2005;7:469-476. DOI: 10.1007/s11051-005-4271-5
  40. 40. Luo M, Wen Q, Liu J, Liu H, Jia Z. Fabrication ofSPES/nano-TiO2 composite ultrafiltration membrane and its anti-fouling mechanism. Chinese Journal of Chemical Engineering. 2011;19:45-51. DOI: 10.1016/S1004-9541(09)60175-0
  41. 41. Celik E, Park H, Choi H, Choi H. Carbon nanotube blended polyethersulfone membranes for fouling control in water treatment. Water Research. 2011;45:274-282. DOI: 10.1016/j.watres.2010.07.060
  42. 42. Jeong BH, Hoek EMV, Yan YS, Subramani A, Huang XF, Hurwitz G, Ghosh AK, Jawor A. Interfacial polymerization of thin film nanocomposites: A new concept for reverse osmosis membranes. Journal of Membrane Science. 2007;294:1-7. DOI: 10.1016/j.memsci.2007.02.025
  43. 43. Nandi BK, Uppaluri R, Purkait MK. Effects of dip coating parameters on the morphology and transport properties of cellulose acetate-ceramic composite membranes. Journal of Membrane Science. 2009;330:246-258. DOI: 10.1016/j.memsci.2008.12.071
  44. 44. Pendergast MTM, Nygaard JM, Ghosh AK, Hoek EMV. Using nanocomposite materials technology to understand and control reverse osmosis membrane compaction. Desalination. 2010;261:255-263. DOI: 10.1016/j.desal.2010.06.008
  45. 45. Qin D, Liu Z, Sun DD, Song X, Bai H. A new nanocomposite forward osmosis membrane custom-designed for treating shale gas wastewater. Scientific Reports. 2015;5(14530):1-14. DOI: 10.1038/srep14530
  46. 46. Xu W, Chen Q, Ge Q. Recent advances in forward osmosis (FO) membrane: Chemical modifications on membranes for FO processes. Desalination. 2017;419:101-116. DOI: 10.1016/j.desal.2017.06.007
  47. 47. Ulbricht M. Advanced functional polymer membranes. Polymer. 2006;47:2217-2262. DOI: 10.1016/j.polymer.2006.01.084
  48. 48. Chung T, Luo L, Feng C, Cui Y, Amy G. What is next for forward osmosis (FO) and pressure retarded osmosis (PRO). Separation and Purification Technology. 2015;156:856-860. DOI: 10.1016/j.seppur.2015.10.063
  49. 49. Chung TS, Li X, Ong RC, Ge QC, Wang HL, Han G. Emerging forward osmosis (FO) technologies and challenges ahead for clean water and clean energy applications. Current Opinion in Chemical Engineering. 2012;1:246-257. DOI: 10.1016/j.coche.2012.07.004
  50. 50. Yip NY, Tiraferri A, Phillip WA, Schiffman JD, Elimelech M. High performance thin-film composite forward osmosis membrane. Environmental Science and Technology. 2010;44:3812-3818. DOI: 10.1021/es1002555
  51. 51. Qiu CQ, Qi SR, Tang CY. Synthesis of high flux forward osmosis membranes by chemically crosslinked layer-by-layer polyelectrolytes. Journal of Membrane Science. 2011;381:74-80. DOI: 10.1016/j.memsci.2011.07.013
  52. 52. Ong RC, Chung TS, de Wit JS, Helmer BJ. Novel cellulose ester substrates for high performance flat-sheet thin-film composite (TFC) forward osmosis (FO) membranes. Journal of Membrane Science. 2015;473:63-71. DOI: 10.1016/j.memsci.2014.08.046
  53. 53. Arslan S, Aslan T, Eyvaz M, Güçlü S, Yüksekdağ A, Yüksel E, Koyuncu İ. Water and reverse salt flux performances of a novel nanofiber forward osmosis membrane. In: 5th IWA Regional Conference on Membrane Technology (IWA-RMTC2016); Kunming City, Yunnan Province China: IWA Publishing; Aug 22-24, 2016
  54. 54. Arslan S, Eyvaz M, Aslan T, Koyuncu İ, Yüksel E. Manufacturing of high performance forward osmosis by innovative manufacturing methods and their applications in submerged membrane bioreactors (Poster Presentation in Turkish). In: Institute of Science and Technology, Graduate Researches-Promotion Days. Gebze Technical University, Department of Architecture, Kocaeli/Turkey. Abstracts Book. GTU Publishing; May 17-18, 2016. p. 46
  55. 55. Wang KY, Ong RC, Chung TS. Double-skinned forward osmosis membranes for reducing internal concentration polarization within the porous sublayer. Industrial & Engineering Chemistry Research. 2010;49:4824-4831. DOI: 10.1021/ie901592d
  56. 56. Chung TS, Zhang S, Wang KY, Su J, Ling MM. Forward osmosis processes: Yesterday, today and tomorrow. Desalination. 2012;287:78-81. DOI: 10.1016/j.desal.2010.12.019
  57. 57. McCutcheon JR, Elimelech M. Influence of membrane support layer hydrophobicity on water flux in osmotically driven membrane processes. Journal of Membrane Science. 2008;318:458-466. DOI: 10.1016/j.memsci.2008.03.021
  58. 58. Yuan H, Abu-Reesh IM, He Z. Enhancing desalination and wastewater treatment by coupling microbial desalination cells with forward osmosis. Chemical Engineering Journal. 2015;270:437-443. DOI: 10.1016/j.cej.2015.02.059
  59. 59. Xu GR, Xu JM, Feng HJ, Zhao HL, Wu SB. Tailoring structures and performance of polyamide thin film composite (PA-TFC) desalination membranes via sublayers adjustment-a review. Desalination. 2017;417:19-35. DOI: 10.1016/j.desal.2017.05.011
  60. 60. Li D, Yan Y, Wang H. Recent advances in polymer and polymer composite membranes for reverse and forward osmosis processes. Progress in Polymer Science. 2016;61:104-155. DOI: 10.1016/j.progpolymsci.2016.03.003
  61. 61. Tian EL, Zhou H, Ren YW, Za M, Wang XZ, Xiong SW. Novel design of hydrophobic/hydrophilic interpenetrating network composite nanofibers for the support layer of forward osmosis membrane. Desalination. 2014;347:207-214. DOI: 10.1016/j.desal.2014.05.043
  62. 62. Dumée L, Lee J, Sears K, Tardy B, Duke M, Gray S. Fabrication of thin film composite poly(amide)-carbon-nanotube supported membranes for enhanced performance in osmotically driven desalination systems. Journal of Membrane Science. 2013;427:422-430. DOI: 10.1016/j.memsci.2012.09.026
  63. 63. Wang R, Shi L, Tang CY, Chou S, Qiu C, Fane AG. Characterization of novel forward osmosis hollow fiber membranes. Journal of Membrane Science. 2010;355:158-167. DOI: 10.1016/j.memsci.2010.03.017
  64. 64. Sukitpaneenit P, Chung T-S. High performance thin-film composite forward osmosis hollow fiber membranes with macrovoid-free and highly porous structure for sustainable water production. Environmental Science and Technology. 2012;46:7358-7365. DOI: 10.1021/es301559z
  65. 65. Han G, Chung T-S, Toriida M, Tamai S. Thin-film composite forward osmosis membranes with novel hydrophilic supports for desalination. Journal of Membrane Science. 2012;423-424:543-555. DOI: 10.1016/j.memsci.2012.09.005
  66. 66. Yasukawa M, Mishima S, Shibuya M, Saeki D, Takahashi T, Miyoshi T, Matsuyama H. Preparation of a forward osmosis membrane using a highly porous polyketone microfiltration membrane as a novel support. Journal of Membrane Science. 2015;487:51-59. DOI: 10.1016/j.memsci.2015.03.043
  67. 67. Tiraferri A, Yip NY, Phillip WA, Schiffman JD, Elimelech M. Relating performance of thin-film composite forward osmosis membranes to support layer formation and structure. Journal of Membrane Science. 2011;367:340-352. DOI: 10.1016/j.memsci.2010.11.014
  68. 68. Huang L, McCutcheon JR. Impact of support layer pore size on performance of thin film composite membranes for forward osmosis. Journal of Membrane Science. 2015;483:25-33. DOI: 10.1016/j.memsci.2015.01.025
  69. 69. Xiao P, Nghiem LD, Yin Y, Li X-M, Zhang M, Chen G, Song J, He T. A sacrificial-layer approach to fabricate polysulfone support for forward osmosis thin-film composite membranes with reduced internal concentration polarisation. Journal of Membrane Science. 2015;481:106-114. DOI: 10.1016/j.memsci.2015.01.036
  70. 70. Vilakati GD, Wong MCY, Hoek EMV, Mamba BB. Relating thin film composite membrane performance to support membrane morphology fabricated using lignin additive. Journal of Membrane Science. 2014;469:216-224. DOI: 10.1016/j.memsci.2014.06.018
  71. 71. Peng N, Widjojo N, Sukitpaneenit P, Teoh MM, Lipscomb GG, Chung T-S, Lai J-Y. Evolution of polymeric hollow fibers as sustainable technologies: Past, present, and future. Progress in Polymer Science. 2012;37:1401-1424. DOI: 10.1016/j.progpolymsci.2012.01.001
  72. 72. Song X, Liu Z, Sun DD. Nano gives the answer: Breaking the bottleneck of internal concentration polarization with a nanofiber composite forward osmosis membrane for a high water production rate. Advanced Materials. 2011;23:3256-3260. DOI: 10.1002/adma.201100510
  73. 73. Wang KY, Chung T-S, Amy G. Developing thin-film-composite forward osmosis membranes on the PES/SPSf substrate through interfacial polymerization. AICHE Journal. 2012;58:770-781. DOI: 10.1002/aic.12635
  74. 74. Arena JT, McCloskey B, Freeman BD, McCutcheon JR. Surface modification of thin film composite membrane support layers with polydopamine: Enabling use of reverse osmosis membranes in pressure retarded osmosis. Journal of Membrane Science. 2011;375:55-62. DOI: 10.1016/j.memsci.2011.01.060
  75. 75. Huang L, Bui N-N, Meyering MT, Hamlin TJ, McCutcheon JR. Novel hydrophilic nylon 6,6 microfiltration membrane supported thin film composite membranes for engineered osmosis. Journal of Membrane Science. 2013;437:141-149. DOI: 10.1016/j.memsci.2013.01.046
  76. 76. Emadzadeh D, Lau WJ, Matsuura T, Ismail AF, Rahbari-Sisakht M. Synthesis and characterization of thin film nanocomposite forward osmosis membrane with hydrophilic nanocomposite sup- port to reduce internal concentration polarization. Journal of Membrane Science. 2014;449:74-85. DOI: 10.1016/j.memsci.2013.08.014
  77. 77. Liang HQ, Hung WS, Yu HH, Hu CC, Lee KR, Lai JY, Xu ZK. Forward osmosis membranes with unprecedented water flux. Journal of Membrane Science. 2017;529:47-54. DOI: 10.1016/j.memsci.2017.01.056
  78. 78. Kwon SJ, Park SH, Park MS, Lee JS, Lee JH. Highly permeable and mechanically durable forward osmosis membranes prepared using polyethylene lithium ion battery separators. Journal of Membrane Science. 2017;544:213-220. DOI: 10.1016/j.memsci.2017.09.022
  79. 79. Kwon SB, Lee JS, Kwon SJ, Yun ST, Lee S, Lee JH. Molecular layer-by-layer assembled forward osmosis membranes. Journal of Membrane Science. 2015;488:111-120. DOI: doi.org/10.1016/j.memsci.2015.04.015
  80. 80. Widjojo M, Chung TS, Weber M, Maletzko C, Warzelhan V. The role of sulphonated polymer and macrovoid-free structure in the support layer for thin-film composite (TFC) forward osmosis (FO) membranes. Journal of Membrane Science. 2011;383:214-223. DOI: 10.1016/j.memsci.2011.08.041
  81. 81. Duong PHH, Chisca S, Hong PY, Cheng H, Nunes SP, Chung TS. Hydroxyl functionalized polytriazole-co-polyoxadiazole as substrates for forward osmosis membranes. ACS Applied Materials & Interfaces. 2015;7:3960-3973. DOI: 10.1021/am508387d
  82. 82. Zhou Z, Lee JY, Chung TS. Thin film composite forward-osmosis membranes with enhanced internal osmotic pressure for internal concentration polarization reduction. Chemical Engineering Journal. 2014;249:236-245. DOI: 10.1016/j.cej.2014.03.049
  83. 83. Sahebi S, Phuntsho S, Woo YC, Park MJ, Tijing LD, Hong S, Shon HK. Effect of sulphonated polyethersulfone substrate for thin film composite forward osmosis membrane. Desalination. 2013;389:129-136. DOI: 10.1016/j.desal.2015.11.028
  84. 84. Li X, Zhang S, Fu F, Chung TS. Deformation and reinforcement of thin-film composite (TFC) polyamide-imide (PAI) membranes for osmotic power generation. Journal of Membrane Science. 2013;434:204-217. DOI: 10.1016/j.memsci.2013.01.049
  85. 85. Liu X, Ng HY. Fabrication of layered silica-polysulfone mixed matrix substrate membrane for enhancing performance of thin-film composite forward osmosis membrane. Journal of Membrane Science. 2015;481:148-163. DOI: 10.1016/j.memsci.2015.02.012
  86. 86. Ma N, Wei J, Qi S, Zhao Y, Gao Y, Tang CY. Nanocomposite substrates for controlling internal concentration polarization in forward osmosis membranes. Journal of Membrane Science. 2013;441:54-62. DOI: 10.1016/j.memsci.2013.04.004
  87. 87. Lu P, Liang S, Qiu L, Gao Y, Wang Q. Thin film nanocomposite forward osmosis membranes based on layered double hydroxide nanoparticles blended substrates. Journal of Membrane Science. 2016;504:196-205. DOI: 10.1016/j.memsci.2015.12.066
  88. 88. Huang L, McCutcheon JR. Hydrophilic nylon 6,6 nanofibers supported thin film composite membranes for engineered osmosis. Journal of Membrane Science. 2014;457:162-169. DOI: 10.1016/j.memsci.2014.01.040
  89. 89. Huang L, Arena JT, McCutcheon JR. Surface modified PVDF nanofiber supported thin film composite membranes for forward osmosis. Journal of Membrane Science. 2016;499:352-360. DOI: 10.1016/j.memsci.2015.10.030
  90. 90. Zhang H, Jiang W, Cui H. Performance of anaerobic forward osmosis membrane bioreactor coupled with microbial electrolysis cell (AnOMEBR) for energy recovery and membrane fouling alleviation. Chemical Engineering Journal. 2017;321:375-383. DOI: 10.1016/j.cej.2017.03.134
  91. 91. Klaysom C, Hermans S, Gahlaut A, Van Craenenbroeck S, Vankelecom IFJ. Polyamide/polyacrylonitrile (PA/PAN) thin film composite osmosis membranes: Film optimization, characterization and performance evaluation. Journal of Membrane Science. 2013;445:25-33. DOI: 10.1016/j.memsci.2013.05.037
  92. 92. Jia Q, Han H, Wang L, Liu B, Yang H, Shen J. Effects of CTAC micelles on the molecular structures and separation performance of thin-film composite (TFC) membranes in forward osmosis processes. Desalination. 2014;340:30-41. DOI: 10.1016/j.desal.2014.02.017
  93. 93. Lu X, Arias Chavez LH, Romero-Vargas Castrillón S, Ma J, Elimelech M. Influence of active layer and support layer surface structures on organic fouling propensity of thin-film composite forward osmosis membranes. Environmental Science and Technology. 2015;49:1436-1444. DOI: 10.1021/es5044062
  94. 94. Tiraferri A, Kang Y, Giannelis EP, Elimelech M. Highly hydrophilic thin-film composite forward osmosis membranes functionalized with surface-tailored nanoparticles. ACS Applied Materials & Interfaces. 2012;4:5044-5053. DOI: 10.1021/am301532g
  95. 95. You S, Lu J, Tang CY, Wang X. Rejection of heavy metals in acidic wastewater by a novel thin-film inorganic forward osmosis membrane. Chemical Engineering Journal. 2017;320:532-538. DOI: 10.1016/j.cej.2017.03.064
  96. 96. Salehi H, Rastgar M, Shakeri A. Anti-fouling and high water permeable forward osmosis membrane fabricated via layer by layer assembly of chitosan/graphene oxide. Applied Surface Science. 2017;413:99-108. DOI: 10.1016/j.apsusc.2017.03.271
  97. 97. Fang W, Wang R, Chou S, Setiawan L, Fane AG. Composite forward osmosis hollow fiber membranes: Integration of RO- and NF-like selective layers to enhance membrane properties of anti- scaling and anti-internal concentration polarization. Journal of Membrane Science. 2012;394-395:140-150. DOI: 10.1016/j.memsci.2011.12.034
  98. 98. Fang W, Liu C, Shi L, Wang R. Composite forward osmosis hollow fiber membranes: Integration of RO- and NF-like selective layers for enhanced organic fouling resistance. Journal of Membrane Science. 2015;492:147-155. DOI: 10.1016/j.memsci.2015.05.045

Written By

Murat Eyvaz, Serkan Arslan, Derya İmer, Ebubekir Yüksel and İsmail Koyuncu

Submitted: 29 March 2017 Reviewed: 07 November 2017 Published: 28 March 2018