Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Membrane Separation Principles and Applications: From Material Selection to Mechanisms and Industrial Uses
Membrane Separation Principles and Applications: From Material Selection to Mechanisms and Industrial Uses
Membrane Separation Principles and Applications: From Material Selection to Mechanisms and Industrial Uses
Ebook875 pages9 hours

Membrane Separation Principles and Applications: From Material Selection to Mechanisms and Industrial Uses

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Membrane Separation Principles and Applications: From Material Selection to Mechanisms and Industrial Uses, the latest volume in the Handbooks in Separation Science series, is the first single resource to explore all aspects of this rapidly growing area of study. Membrane technology is now accepted as one of the most effective tools for separation and purification, primarily due to its simple operation. The result has been a proliferation of studies on this topic; however, the relationships between fundamental knowledge and applications are rarely discussed. This book acts as a guideline for those who are interested in exploring membranes at a more progressive level.

Covering methods of pressure driving force, partial pressure driving force, concentration driving force, electrical potential driving force, hybrid processes, and more, this volume is more complete than any other known resource on membrane separations.

  • Covers membrane material selection, membrane fabrication, membrane characterization, separation mechanisms and applications in each chapter
  • Authored by contributors who are internationally recognized as experts in their respective fields
  • Organized by the driving force behind each type of membrane separation—a unique approach that more clearly links fundamental principles with their dominant applications
LanguageEnglish
Release dateSep 7, 2018
ISBN9780128128169
Membrane Separation Principles and Applications: From Material Selection to Mechanisms and Industrial Uses

Related to Membrane Separation Principles and Applications

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Membrane Separation Principles and Applications

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Membrane Separation Principles and Applications - Ahmad Fauzi Ismail

    Malaysia

    Chapter 1

    Reverse Osmosis Membrane Separation Technology

    Yi-Ning Wang⁎; Rong Wang⁎,†    ⁎ Singapore Membrane Technology Centre, Nanyang Environmental and Water Research Institute, Singapore, Singapore

    † School of Civil and Environmental Engineering, Nanyang Technological University, Singapore, Singapore

    Abstract

    Reverse osmosis (RO) is a pressure driven membrane process which has been recognized as the leading technology widely applied for seawater desalination and water reclamation. Advancements over the 30 years in RO technology including advanced membrane fabrication, membrane module and process design and optimization, installation of energy recovery devices, and the use of more efficient pumps, etc., have led to significant improvement in separation performance, dramatic energy reduction in RO process and brought in more commercial applications. This chapter aims to provide a fundamental understanding on RO technology, address the main challenges of this technology as well as its wide applications. In specific, it firstly introduces the historic development of RO technology and the basic properties of RO membranes, followed by the elaborations of various membrane fabrication and characterization methods and module/process design. The key challenging issues encountered in RO applications such as concentration polarization and membrane fouling are highlighted with provisions of mitigation and control strategies, followed by a summary of various RO applications.

    Keywords

    Membrane; Reverse osmosis; Membrane fabrication; Membrane characterization; Membrane module; Membrane fouling and control; Desalination; Water reclamation; RO applications

    Outline

    1.1Introduction of Reverse Osmosis

    1.1.1Historic Development of RO

    1.1.2Basic Properties of RO Membrane

    1.2RO Membrane Fabrication

    1.2.1Cellulose Acetate Membrane

    1.2.2TFC Polyamide Membrane

    1.2.3Membrane With a Polyelectrolyte Multilayer Film

    1.2.4Recent Advances in Membranes

    1.3Membrane Properties and Characterizations

    1.3.1Membrane Properties

    1.3.2Membrane Characterizations

    1.4Membrane Modules and Process Operation

    1.4.1Membrane Modules

    1.4.2Process Operation

    1.5Concentration Polarization

    1.6Membrane Fouling and Control

    1.6.1Factors Affecting Membrane Fouling

    1.6.2Fouling Mitigation

    1.7RO Applications

    1.7.1Desalination and Water Reclamation

    1.7.2Ultrapure Water Production

    1.7.3Solute Concentration

    1.7.4Organic Solvent Separation

    1.8Conclusions

    References

    Further Reading

    List of Symbols

    A 

    water permeability coefficient

    B 

    solute permeability coefficient

    c 

    solute concentration

    cb 

    bulk concentration

    cf 

    feed concentration

    cm 

    solute concentration near membrane surface

    cp 

    solute concentration in the permeate

    Cwm 

    the concentration of water inside the selective layer

    dh 

    hydraulic diameter

    D 

    diffusion coefficient of solute

    Dsm 

    diffusion coefficient of solute inside RO rejection layer

    Dwm 

    diffusion coefficient of water inside RO rejection layer

    J 

    water flux

    Js 

    solute flux

    k 

    mass transfer coefficient

    Ksm 

    solute partitioning coefficient in the selective layer

    L 

    channel length

    lm 

    thickness of a selective layer

    P 

    hydraulic pressure

    R 

    rejection

    Re 

    Reynolds number

    Rg 

    universal gas constant (8.31 J mol− 1 K− 1)

    Rint 

    intrinsic rejection of a membrane

    Rm 

    membrane hydraulic resistance

    Sc 

    Schemidt number

    Sh 

    Sherwood number

    T 

    absolute temperature (K)

    u 

    cross-flow velocity

    Vw 

    molar volume of water

    v 

    kinematic viscosity

    η 

    dynamic viscosity of water

    δ 

    boundary layer thickness

    π 

    osmotic pressure

    1.1 Introduction of Reverse Osmosis

    Osmosis, also called forward osmosis (FO), a natural phenomenon discovered as early as 1748, is a diffusion of fluid (usually water) through a semi-permeable membrane from a solution with a low solute concentration to a solution with a higher solute concentration until an equilibrium of fluid concentration on both sides of the membrane is reached [1]. The driving force for water transport is the chemical potential difference (i.e., osmotic pressure difference) across the membrane. Reverse osmosis (RO) is achieved by applying a pressure in excess of the osmotic pressure gradient to drive the water to flow from the high solute concentration side to the low solute concentration side, i.e., in an oppose direction to the automatic water flow in the FO process. Fig. 1.1 illustrates four osmotic membrane processes including pressure assisted osmosis (PAO), FO, pressure retarded osmosis (PRO) and RO. The four processes are similar in a way that a semi-permeable membrane is placed in between a solution of low osmotic pressure and a solution of high osmotic pressure. The first three (PAO, FO, and PRO) are the processes (partially) driven by osmotic pressure difference, where water flows from the low osmotic pressure side to the high osmotic pressure side. They are not able to achieve desalination by themselves. Only when the water flow direction is reversed in the RO process, can clean and desalted water be directly obtained by this pressure driven membrane process.

    Fig. 1.1 Osmotic membrane processes including pressure assisted osmosis (PAO), forward osmosis (FO), pressure retarded osmosis (PRO) and reverse osmosis (RO).

    1.1.1 Historic Development of RO

    The use of RO as a feasible separation process was first demonstrated in the early 1960s. The milestones of RO membrane development are shown in Fig. 1.2. The first practical RO membrane made of cellulose acetate (CA) was reported by Loeb and Sourirajan in 1962 [2]. This membrane exhibited high flux and salt rejection owing to the asymmetric structure (a thin skin layer on a porous substrate). Cadotte and Petersen later made the first efficient thin-film composite (TFC) membrane based on the reaction of phenylene diamine and trimesoyl chloride, which was a key breakthrough to make the large-scale application of RO membranes economically feasible [3]. Today, the state-of-art desalination RO membranes are still TFC polyamide (PA) RO membranes [4]. Although they are produced based on the original chemistry discovered during the 1980s, the performance of RO membranes have been improved significantly over the 30 years [5], by closer monitoring of the recipes for the substrates and selective layer and/or applying coating, etc. [6]. More than sevenfold decrease in salt passage has expanded the range of saline feeds, while the increased lifespan (2.3 times) and increased water permeability (2.5 times) have greatly reduced the real cost (one twelfth) [5, 7]. Currently, TFC RO membranes are produced in flat sheet configuration and mostly supplied in spiral wound module (SWM). With the aim of further reducing energy cost in desalination (theoretical minimal energy required to desalinate 35 g/L seawater at 50% recovery is ~ 1.06 kWh/m³, while the real energy consumption is ~ 2–5 kWh/m³[8, 9]), extensive researches are focused on developing more superior performing, fouling resistant and chlorine resistant RO membranes, in addition to the optimizations of the process and membrane module design.

    Fig. 1.2 Milestones of RO membrane development. Adapted from Baker RW. Reverse osmosis. Membrane technology and applications. John Wiley & Sons, Ltd; 2004. p. 191–235.

    Since 2007, the emerging thin-film nanocomposite (TFN) membranes/mixed matrix membranes (MMMs), i.e., TFC membranes with incorporated nanoparticles/inorganic particles in the selective layer and/or the support layer, have attracted great interests in fabricating high performance RO membranes [10, 11], due to their enhanced performance as compared with pristine TFC membranes [12].

    1.1.2 Basic Properties of RO Membrane

    The basic properties of RO membrane are summarized in Table 1.1 along with other types of pressure driven membranes such as nanofiltration (NF), ultrafiltration (UF), and microfiltration (MF), which are categorized based on the pore size or operating pressure range. MF and UF membranes are the low-pressure membranes with relatively larger pores. They are usually adopted in the pretreatment of RO/NF processes to protect the RO/NF membranes. They are also used as the support for fabricating TFC RO and NF membranes. RO membranes have the tightest surface pores (< 2 nm (diameter), generally considered as non-porous membrane) and are able to retain dissolved ions (including monovalent ions Na+ and Cl−) and small organic molecules. RO membranes can be further divided into seawater RO (SWRO) and brackish water RO (BWRO) membranes. SWRO membranes have high NaCl rejection (> 99%) and are used for seawater desalination. The trade-off for this high rejection is the low water permeability (< 1 L/m² h/bar). High pressure of above 60 bar is generally required in real operation to overcome the seawater osmotic pressure to produce reasonably high-water flux. BWRO membranes have relatively low NaCl rejection (> 95%) and higher water permeability (1–10 L/m² h/bar), and are used for water/wastewater treatment where feed salt content is much lower compared to that of seawater. NF membranes are similar to RO membranes, but the bigger pore size of NF membranes as compared with RO membranes leads to the higher water permeability and much lower rejection to the monovalent ions (e.g., 10%–90% rejection to Na+ depending strongly on the surface charges of the membrane). NF membranes can effectively remove di- and multi-valent ions, which makes them useful for water softening.

    Table 1.1

    Note: MWCO, molecular weight cut-off.

    Adapted from Fane AG, Wang R, Tang C. Membrane technology for water: microfiltration, ultrafiltration, nanofiltration, and reverse osmosis. Treatise on water science, vol. 4. Amsterdam; Hackensack, NJ: Elsevier Science; 2011. p. 301; Mulder M. Basic principles of membrane technology. The Netherlands: Kluwer Academic Publishers; 1996.

    During a pressure driven membrane process, a feed stream is divided into two streams, i.e., retentate stream (or concentrate stream) that is retained by the membrane and permeate stream that passes through the membrane. It is widely accepted that the transport of molecules through the selective skin layer of a RO membrane is best described by the solution-diffusion model [14, 15]. It assumes that both the solvent and the solute are absorbed to the skin layer, and then diffuse through the non-porous layer independent of each other under their respective chemical potential gradient. A separation is achieved between the solvent and the solute (or among different permeants) because of the differences in the amount of the molecules that dissolve in the membrane (solubility) and the rate at which the molecules diffuse through the membrane (diffusivity) [14]. RO membranes generally allow high absorption of water molecules and fast diffusion of them through the selective layer, as compared with sodium chloride [15].

    1.2 RO Membrane Fabrication

    RO membranes are generally categorized into integrally skinned asymmetric membranes and TFC membranes based on the structure and materials. The common materials used for RO membrane fabrication are summarized in Table 1.2. An integrally skinned asymmetric membrane is made of one polymeric material, and CA membranes (made of cellulose esters) are the most widely used examples [15]. The TFC RO membrane is the state-of-the-art RO membrane, which comprises two or more polymer materials. The porous membrane support is typically made of polysulfone (PSf) or polyethersulfone (PES). The ultrathin selective layer is generally made of polyamide (PA), which has relatively high water flux and high solute rejection, as well as good chemical, mechanical, and thermal stability (Table 1.2) [9].

    Table 1.2

    1.2.1 Cellulose Acetate Membrane

    Cellulose esters, including cellulose diacetate and cellulose triacetate, are used to prepare the integrally skinned asymmetric RO membranes via phase inversion. Phase inversion refers to a process whereby a polymer solution (liquid form; in which the solvent system is the continuous phase) inverts into a swollen three-dimensional macromolecular network or gel (solid form; where the polymer is the continuous phase) [9]. In other words, a liquid polymer solution is precipitated into two phases: a polymer-rich phase that forms the matrix of the membrane, and a polymer-lean phase that forms the membrane pores in an unstable nascent membrane structure. For RO membrane/RO substrate preparation, the precipitation is generally achieved by the immersion of a polymer solution into a non-solvent coagulant bath, such as water (solvent containing water coagulant bath is usually adopted for hollow fiber membrane preparation [16]). Both flat-sheet membranes and hollow fiber membranes can be fabricated using this method.

    Despite the ease of preparation (i.e., one step casting), it is generally difficult for the integrally skinned asymmetric RO membranes to achieve both high water permeability and high solute rejection. Formation of a dense skin layer requires a relatively high polymer concentration, which results in a less porous/less permeable support and therefore an increased overall hydraulic resistance of the membrane. It has been reported that the polymer concentration for making integrally skinned asymmetric CA membrane is higher than that of TFC membrane with CA support [17, 18]. A common practice for increasing the solute rejection of CA membrane is to adopt an annealing post-treatment, such as heating in a bath of hot water for a few minutes [19]. This process modifies the selective layer by eliminating the micropores and producing a denser skin, resulting in a decrease in water flux and an increase in salt rejection [19]. The annealing temperature determines the final properties of the membrane and a higher temperature generally leads to a denser skin layer. To compensate the low permeability, the commercial CA RO membrane is preferentially made into hollow fiber modules (e.g., Toyobo hollow fiber membrane with a very small outer diameter of ~ 165 μm [20]), which can provide sufficiently high water productivity owing to the high packing density (high specific membrane area) [21]. Considerable research efforts have been made to optimize CA membranes via changing the compositions of polymer solution (e.g., type and concentration of polymer/additives) and casting/spinning conditions (e.g., composition and temperature of coagulant bath), and modifying the surface properties, etc. [22–24]. However, their drawbacks of poor stability against chemicals, bacteria, and temperature and lower water permeability make it less competitive than the TFC polyamide membranes (see Tables 1.2 and 1.3).

    Table 1.3

    Adapted from Fane AG, Wang R, Tang C. Membrane technology for water: microfiltration, ultrafiltration, nanofiltration, and reverse osmosis. Treatise on water science, vol. 4. Amsterdam; Hackensack, NJ: Elsevier Science; 2011. p. 301.

    1.2.2 TFC Polyamide Membrane

    A TFC polyamide (PA) RO membrane generally comprises three layers, including an aromatic PA selective layer (typically < 200 nm thick), a microporous support (usually a UF membrane, 20–50 μm), and a non-woven fabric (backing layer, 100–200 μm) [9]. A schematic of a typical TFC RO membrane is shown in Fig. 1.3. The selective layer and the support are produced separately, enabling the optimization of each individual layer for the respective specific function, i.e., the top selective layer is thin but dense enough to attain high solvent flux and high solute rejection; and the support is porous enough to provide low resistance to permeate while maintaining high mechanical strength [25]. This approach provides great flexibility for utilizing different materials to tailor the membrane structure and properties.

    Fig. 1.3 Schematic of a thin-film composite PA membrane.

    The PA layer of a TFC RO membrane is formed on a porous membrane support via interfacial polymerization between amines (e.g., m-phenylenediamine, MPD) and aromatic acyl chlorides (e.g., trimesoyl chloride, TMC). In the process of preparation (Fig. 1.4), a membrane support (UF or MF membrane) is immersed in an aqueous solution containing MPD. After removing the excess MPD, the membrane surface is immersed in/contacted with a water-immiscible solvent containing dissolved TMC. Due to the rapid hydrolysis of acyl chloride in the aqueous phase and the asymmetric solubility, the reaction is diffusion-controlled and comprises three stages [26]: At the incipient stage, a loose polymer film begins to emerge in the interface between two immiscible solvents; then amines in the aqueous phase diffuse to the organic phase and the film becomes dense and grows perpendicularly towards the organic phase; finally the increase in thickness and density of the film inhibits the diffusion of the monomers and the reaction. The advantage of interfacial reaction is the fast reaction kinetics and self-inhibiting, resulting in an extremely thin film. Heat treatment is often applied to complete the interfacial reaction and further cross-linking [19]. The type and concentration of monomers/solvents/additives, polymerization condition, and curing process can affect the performance of the PA layer and these conditions need to be optimized for fabricating high-performance TFC membranes [27, 28].

    Fig. 1.4 (A) Schematic of the formation of a thin-film composite membrane via interfacial polymerization; (B) reaction between MPD and TMC. Adapted from Weng X, Ji Y, Zhao F, An Q, Gao C. Tailoring the structure of polyamide thin film composite membrane with zwitterions to achieve high water permeability and antifouling property. RSC Adv 2015;5(120):98730-9

    In addition to the control of the reaction parameters of interfacial polymerization, the surface properties of the support such as pore structure, surface morphology and chemistry are also important for the PA layer formation [27, 29]. For example, a relatively porous and hydrophobic support could result in the formation of a more permeable and rougher skin layer; whereas a smoother PA layer with lower permeability is formed on a support of hydrophilic surface [30]. The increasing surface pore size was also found to increase the RO membrane water permeability [31]. Although it has been found that the support surface properties play an important role in forming PA layer, the mechanisms are yet to be systematically investigated. Besides the surface properties, an ideal support of a RO membrane shall possess good mechanical, chemical, thermal, and biological stabilities and relatively low hydraulic resistance.

    In the aspect of RO support preparation, phase inversion is the most extensively used method, with which an asymmetric UF membrane is generally formed from polymers PSf or PES (Table 1.2). Similar to the CA RO membrane fabrication and other UF membrane preparation, factors such as the composition of polymer solution (e.g., the type and concentration of polymer/additive), and coagulant bath, etc. have been shown to influence the properties of resultant supports [15, 27]. For the commercially available TFC membranes, the TFC RO supports are produced in flat sheets; hollow fiber TFC RO membranes are only prepared in lab scale currently due to the challenges of large-scale production. Nevertheless, the high performance hollow fiber TFC RO membranes have been recently developed in a few research labs with promising applications in brackish water desalination and surface water treatment [32].

    On the other hand, the recently emerged nanofibrous membranes produced by electrospinning have attracted many interests in preparing membranes or membrane supports [29, 33]. Electrospinning involves a process during which polymer is stretched into nanofibers from its solution under an applied electrostatic force. This technique is so versatile that the produced nanofibrous membranes can be tailored with desired pore size, porosity, and thickness by varying the polymer solution and electrospinning conditions [29, 34]. A typical nanofibrous membrane has interconnected pores and relatively high porosity, which is favorable for making RO/NF membranes with high water permeability [33]. Nevertheless, the surface pore size of the nanofibrous membrane support needs to be carefully controlled within a certain range for the successful formation of PA selective layer. The nanofibrous support with high porosity may also require other treatment to enhance its stability against high pressure, such as heat pressing [33].

    1.2.3 Membrane With a Polyelectrolyte Multilayer Film

    RO/NF membranes can be prepared by the layer-by-layer assembly of polyelectrolytes onto a charged support (UF or NF supports), where a polycation (e.g., polyallylamine hydrochloride, PAH) and a polyanion (e.g., poly(sodium 4-styrene-sulfonate), PSS) are deposited alternately [35–37]. Fig. 1.5 presents a schematic of layer-by-layer assembly cycle. The rejection property of the membrane usually depends on several factors, such as the type and concentration of polyelectrolytes, ionic condition of the polyelectrolyte solution, contact time, and the number of bilayers (one positively charged layer and one negatively charged layer) [37, 38]. These conditions may influence both the tightness of the pores in the membrane skin layer as well as the surface charge, thereby affecting the passage of solutes, especially charged solutes. This method is more widely used for NF membrane preparation, as the monovalent rejection of the polyelectrolyte film is not very high while the divalent ions are much better retained. To enhance the solute rejection and the film stability, cross-linking via heat treatment and/or the use of cross-linking agents are generally adopted [37–40].

    Fig. 1.5 Schematic of layer-by-layer assembly for preparing polyelectrolyte multilayer film.

    1.2.4 Recent Advances in Membranes

    1.2.4.1 Mixed Matrix Membranes

    Mixed matrix membranes (MMMs) for RO applications are fabricated by incorporating inorganic nanoparticles into membrane matrix via blending the materials into the polymer solution for the support or the monomer solution for the PA layer [27, 41]. The membrane with nanoparticles embedded PA layer is also named thin-film nanocomposite (TFN) membrane. The pioneering study by Hoek's group [12] reported the fabrication of a TFN RO membrane via immobilizing zeolite NaA nanoparticles in the PA layer at loadings of 0.004%–0.4% (w/v). The results demonstrated dramatic increase in water permeability upon zeolite incorporation, while solute rejection remained comparable to the pure TFC membrane (Fig. 1.6). The improved intrinsic properties of the TFN membranes could be attributed to the unique pore structure of zeolite molecular sieve particles, which provide a preferential flow path for water molecules through its internal pore structure. In addition, the hydrophilicity and negative charges of zeolite nanoparticles could result in greater affinity for water molecules and thus increase the repulsion of anions [10]. Similar to zeolites that provide size and shape selectivities, silica (SiO2) are also available in various forms and sizes, such as nonporous silica and mesoporous silica. The latter has pore size range of ~ 2–50 nm, which could make even shorter water flow path in TFN membranes [42]. However, increasing mesopore size of silica may cause sacrificed salt retention [43]. Besides the selection of nanoparticles, the loading of nanoparticles shall also be carefully controlled to minimally affect the crosslinking degree of PA layer [27]. Other promising nanoparticles for MMM preparation include titanium nanotube, silver (Ag), carbon nanotube (CNT), and grapheme oxide (GO), etc. In addition to the enhanced separation properties, TFN membranes may exhibit improved chemical and thermal stability, improved reaction and sorption capacity [27]. For example, the TFN membrane with Ag incorporated in the selective layer exhibited better anti-biofouling property [44], and CNT embedded TFN membranes showed improved resistance to chlorine and fouling [45].

    Fig. 1.6 Zeolite nanoparticle-embedded TFN membranes. (A) Schematic of TFN membranes; (B) SEM image of zeolite nanoparticles; (C) TEM image of PA layer with embedded zeolite nanopartilces; and (D) water permeability and solute rejection performance upon zeolite loading. Reproduced from Jeong BH, Hoek EMV, Yan Y, Subramani A, Huang X, Hurwitz G, et al. Interfacial polymerization of thin film nanocomposites: a new concept for reverse osmosis membranes. J Membr Sci 2007;294(1-2):1-7.

    On the other hand, MMMs with fillers in the supports have also been extensively investigated [27]. Although TFN approach seems to be more effective to improve RO membrane separation property, the nanoparticle incorporation in the supports may benefit in a different way. Similar to the effect of changing polymer solution composition (see Section 1.3.2), a small amount of fillers in the support can significantly change the support's properties, which may subsequently affect the selective layer formation [27]. The hydrophilic inorganic particles such as silica, zeolites, TiO2, and CNTs could increase the hydrophilicity of the support layer. Meanwhile, the changed pore structure (especially the surface characteristics of the support) may contribute to the enhanced water permeability [46]. This approach is very helpful in decreasing the structural parameter of the support layer of FO membranes to reduce the internal concentration polarization effect, thereby increasing FO water flux [46]. Moreover, the fillers with superior mechanical property such as CNT can significantly increase the tensile strength of the MMM membranes [47]. However, there are still some challenges encountered during TFN membrane/MMM fabrication, such as poor dispersion of nanoparticles in the casting solution, agglomeration of nanoparticles in the PA layer, lack of interaction between nanoparticles and membrane matrix (leach out easily), and unknown water and ion transport mechanisms especially for nonporous nanoparticles [10].

    1.2.4.2 Biomimetic Membranes

    Different from the MMMs, the fillers used in biomimetic membranes are aquaporin (AQP) laden vesicles. AQPs are known as water channels, which selectively conduct water molecules while preventing the passage of ions and other solutes (Fig. 1.7A). These characteristics make them perfect for desalination. The single channel permeability of AQP is ~ 2–14 × 10− 14 cm³ water molecules/s (CNT is on the orders of 10− 14–10− 12 cm³/s but without perfect solute rejection) [48–50]. AQP laden vesicles have been demonstrated to possess high water productivity as compared to that of commercial RO membranes [51]. Among different methods of making biomimetic membranes, TFC membrane is still the most practical and promising approach [48]. Conceptually, the way of immobilizing biomolecules into a PA layer is similar to the TFN membrane fabrication. The fabrication of biomimetic TFC RO membrane has been firstly successfully demonstrated by Zhao et al. [52], and the resultant membrane showed 25% increase in water flux without compromising the NaCl rejection. Later Li et al. showed that the water permeability of AQP incorporated TFC hollow fiber membrane reached ~ 8 L/m² h/bar, which was twice as high as those of pristine TFC membrane and commercial BWRO membranes [32]. A cross-sectional SEM image of biomimetic RO membrane is shown in Fig. 1.7B, where the AQP laden vesicles can be observed within the PA layer. Despite the superior separation performances, other issues shall be taken into consideration, such as membrane stability (very few studies looked into the long term stability [53]), scalability, and production cost. A comprehensive review on biomimetic membranes can be found elsewhere [48]. On the other hand, the high performance of natural water channel has inspired researchers to synthesize artificial water channels with nanotubular structures (Fig. 1.7A), which have potential to overcome the challenges faced by AQP based biomimetic membranes [54].

    Fig. 1.7 Water channels and biomimetic membranes. (A) Aquaporin and synthetic water channels (self-assembled artificial water channel and unimolecular artificial water channel); (B) biomimetic TFC RO membranes with incorporated aquaporins (spanned in lipid vesicles) in the PA layer. Adapted from Li X, Chou S, Wang R, Shi L, Fang W, Chaitra G, et al. Nature gives the best solution for desalination: aquaporin-based hollow fiber composite membrane with superior performance. J Membr Sci 2015;494:68–77; Barboiu M. Artificial water channels—incipient innovative developments. Chem Commun 2016;52(33):5657–65.

    1.3 Membrane Properties and Characterizations

    Membrane separation properties generally consist of water permeability and solute rejection, which directly determines membrane productivity. In addition, other important characteristics include hydrophilicity, surface roughness and charge, and membrane stability, which are not only related to the separation properties, but also influence membrane fouling significantly. Correspondingly, membrane characterization (Table 1.4) provides the information about the important membrane properties. It can guide the optimization of membrane fabrication to achieve excellent separation efficiency and fouling resistance.

    Table 1.4

    Adapted from Fane AG, Wang R, Tang C. Membrane technology for water: microfiltration, ultrafiltration, nanofiltration, and reverse osmosis. Treatise on water science, vol. 4. Amsterdam; Hackensack, NJ: Elsevier Science; 2011. p. 301.

    1.3.1 Membrane Properties

    1.3.1.1 Water Permeability and Solute Permeability

    The solution-diffusion model assumes that the solute and solvent are absorbed into and diffuse through the membrane selective layer, independent of each other, under the respective chemical potential gradient. According to the model, the water flux J and the solute flux Js are proportional to the net pressure difference and the concentration difference, respectively, across the membrane [9]:

       (1.1)

       (1.2)

    where A is defined as water permeability coefficient and B is the solute permeability coefficient. Δ P and Δ π are hydraulic pressure difference and osmotic pressure difference across the membrane, respectively. Δ c is the concentration difference across the membrane. The transport coefficients in this model can be related to membrane properties as shown below:

       (1.3)

       (1.4)

    where Dwm and Dsm are water diffusion coefficient and solute diffusion coefficient inside the selective layer, respectively. Csm is the concentration of water inside the selective layer, Vw is the molar volume of water, Ksm is the solute partitioning coefficient in the selective layer, lm is the thickness of the selective layer, Rg is universal gas constant, and T is absolute temperature.

    The intrinsic rejection of a RO membrane Rint is defined as [55]:

       (1.5)

    where cp and cm are the solute concentration in permeate and at the feed solution/membrane interface, respectively, and Rint can be determined using the following equation:

       (1.6)

    According to the above Eqs. (1.3), (1.4), and (1.6), the intrinsic rejection can be improved by increasing the preferential absorption of water molecules to the solutes, enhancing the diffusion coefficient of water molecules inside the selective layer compared with that of solutes, and increasing the applied pressure. This intrinsic rejection is different from apparent rejection R, which is directly measured from experiments and can be greatly affected by the concentration polarization (CP) (as discussed in Section 1.5):

       (1.7)

    where cf is the solute concentration in the feed. In general, apparent rejection R has smaller value than intrinsic rejection Rint due to the CP effect.

    In addition to the solution diffusion model, permeate water flux J can be also correlated to the driving force by following Darcy's law and is commonly presented in terms of membrane hydraulic resistance Rm and dynamic viscosity of the permeating water η[9]:

       (1.8)

    Although A and B value are the intrinsic properties of a membrane, the measured water flux and rejection are affected by different operating conditions, which generally include the applied pressure, cross-flow velocity of the feed stream, feed concentration (or recovery), and temperature (Fig. 1.8). At a relatively low applied pressure (low water flux level), the water flux increases linearly with applied pressure (Eq. 1.2 or Eq. 1.8), the solute rejection also increases with increasing pressure based on the solution diffusion model (Eq. 1.6). However, at a higher applied pressure, the water flux increase deviates from the linear line, due to the increased concentration polarization (i.e., increased concentration/osmotic pressure at membrane surface, see Section 1.5). The increased concentration at membrane surface also results in a lower apparent rejection. Increasing the cross-flow velocity tends to increase both water flux and solute rejection, as a result of reduced concentration polarization. Increasing the recovery or increasing the feed concentration has an effect of reducing both water flux and solute rejection, due to the increased osmotic pressure of the bulk solution. Hence, the recovery for a typical SWRO desalination plant is limited to 50%, while that for a water reclamation plant is controlled below 80% [9]. Increasing the system temperature can lead to an increase in both water flux and solute flux, due to the enhanced diffusion coefficient (Dwm and Dsm) in the membrane selective layer. The solute flux generally increases more drastically than the water flux, so the solute rejection tends to decrease at a higher temperature. It is noted that the change in temperature affects the intrinsic properties of membrane selective layer, while the other operating conditions discussed here are related to the solute concentration at membrane surface.

    Fig. 1.8 Effect of operating conditions on RO membrane performance. Adapted from Fane AG, Wang R, Tang C. Membrane technology for water: microfiltration, ultrafiltration, nanofiltration, and reverse osmosis. Treatise on water science, vol. 4. Amsterdam; Hackensack, NJ: Elsevier Science; 2011. p. 301.

    1.3.1.2 Hydrophilicity

    Hydrophilicity is a characteristic of materials exhibiting an affinity to water. The surface of a hydrophilic membrane can readily absorb water. In contrast, a hydrophobic membrane has little or no tendency to absorb water. The relative hydrophilicity/hydrophobicity can be determined using contact angle measurement (see Table 1.4). In general, a hydrophilic membrane surface is preferred in RO applications for enhancing water permeability and reducing membrane fouling propensity [56–58]. In this regard, CA membrane surface is preferred compared to a typical TFC membrane surface (a comparison of the two membranes is shown in Table 1.3).

    1.3.1.3 Surface Roughness

    Membrane surface roughness is an important parameter affecting RO membrane fouling. A rougher surface is subject to more membrane fouling due to the decreased shear force over the rougher surface and the increased membrane non-homogeneity (i.e., varied flux distribution over the membrane surface and the local high flux would promote fouling (see Section 1.6)) [9]. TFC RO membranes consisting of PA skin layer have ridge and valley surface morphology that contributes to the roughness. In contrast, CA membrane as well as semi-aromatic NF membranes possess smoother surface, which is less prone to initial foulant deposition/growth [58–60]. Nevertheless, surface modification can be applied to decrease the roughness, such as the polyvinyl alcohol (PVA) coating on BW30 membrane (Dow Filmtec) [61].

    1.3.1.4 Surface Charge

    The surface charge of a membrane is generally dependent on its charged functional groups or the preferential adsorption of some specific ionic species [9]. Most TFC RO membranes have negatively charged active surfaces at neutral pH, due to the presence of –COOH functional groups on the PA layer. The surface charge can influence solute rejection. For example, the rejection of sodium chloride by a typical RO membrane can be partially contributed by Donnan exclusion effect [13], i.e., Cl− is retained by the negatively charged surface due to electrostatic repulsion, and Na+ is also retained to maintain charge neutrality [13]. The rejection of charged trace organic solutes is also affected by the charge of membrane [62]. Moreover, the membrane water permeability may be affected by surface charge, as evidenced by the lowest water flux at pH 5 (i.e., the isoelectric point of the membrane) where membrane surface is least charged [63]. On the other hand, the membrane surface charge plays an important role in fouling. The attractive force between oppositely charged foulants and membrane surface has a strong tendency to induce foulant deposition [64]. Membrane-foulant interaction can greatly affect flux performance at initial fouling stage [65].

    1.3.1.5 Stability

    The mechanical, chemical and thermal stabilities of a membrane are important for it to be successfully applied in a membrane process. The requirement of the mechanical strength of a RO membrane depends on its application. For example, SWRO membrane shall withstand higher pressure than the BWRO membrane. The chemical and thermal stabilities are closely related to the type of materials used for membrane fabrication. The RO membranes made of CA have relatively poor stability against chemicals, bacteria, and temperature. The typical operating conditions for such membranes are in range of pH 5–7 to avoid the hydrolysis of the polymer, at temperature below 30°C to avoid the annealing effect (see Section 1.2.1). In contrast, the most widely used TFC RO membranes with PA selective layer are generally more stable, except that it has low chlorine resistance. To prevent unacceptable membrane damage, the chlorine level shall be carefully controlled at a level below 0.1 ppm (Table 1.3) [66,

    Enjoying the preview?
    Page 1 of 1