Next Article in Journal
Composite Gel Polymer Electrolytes Based on Organo-Modified Nanoclays: Investigation on Lithium-Ion Transport and Mechanical Properties
Previous Article in Journal
Chemical Crosslinking of 6FDA-ODA and 6FDA-ODA:DABA for Improved CO2/CH4 Separation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Important Approaches to Enhance Reverse Osmosis (RO) Thin Film Composite (TFC) Membranes Performance

Department of Chemical Engineering, University of Missouri, Columbia, MO 65211, USA
Membranes 2018, 8(3), 68; https://doi.org/10.3390/membranes8030068
Submission received: 17 July 2018 / Revised: 7 August 2018 / Accepted: 9 August 2018 / Published: 21 August 2018
(This article belongs to the Section Inorganic Membranes)

Abstract

:
Thin film composite (TFC) membrane, which consists of polyamide (PA) active film rests on porous support layer, has been the major type of reverse osmosis (RO) membrane since its development by Cadotte in the 1970s, and has been remarkably used to produce clean water for human consumption and domestic utilization. In the past 30 years, different approaches have been exploited to produce the TFC membrane with high water flux, excellent salt rejection, and better chlorine/fouling resistance. In this brief review, we classify the techniques that have been utilized to improve the RO-TFC membrane properties into four categories: (1) Using alternative monomers to prepare the active layer; (2) modification of membrane surface; (3) optimization of polymerization reactions; and (4) incorporation of nanoparticles (NPs) into the membrane PA layer. This review can provide insights to guide future research and further propel the RO TFN membrane.

1. Introduction

Because of the rapid growth of the world population and rising water needs, water shortage problems have become dominant [1,2]. In the last century, as the global population quadrupled, the world water demand has increased sevenfold. The problem of water scarcity is not only a problem of appropriate techniques, but also a social and educational problem relying on national and global endeavors as well as on technical solutions [3]. To address water shortage problems, many techniques have been developed including distillation, membrane reverse osmosis (RO), mechanical vapor pressure compression, electrodialysis and nanofiltration processes [4]. Membrane separation processes are gaining global acceptance in both water treatment and desalination due to their simplicity and relatively low cost compared to other treatment technologies [5]. RO membranes can be effectively used to remove salts and other pollutants from brackish water [6]. The water is transferred through the RO membrane by diffusion [7], while the salt is rejected by size exclusion and repulsion electrostatic force between the membrane surface and dissolved ions, which is caused by charge difference [8,9]. For efficient desalination, the membranes must be permeable to water, impermeable to solutes, and capable of tolerating high operating pressures [9].
The first work on RO membranes was initiated by Reid and co-workers [10] in the early 1950s, when they successfully fabricated an active cellulose acetate membrane to remove salt from water. The membrane exhibited efficient desalination (salt rejection: 96%), but water flux through the membrane was significantly low. Researches continued at the University of California, Los Angeles, with the concern of improving water flux without sacrificing salt rejection [11]. In 1961, Sourirajain [12] enhanced membrane flux by increasing cellulous film porosity through using pore-forming monomers. However, cellulose acetate membrane has limited applications due to its weak chemical resistance and low thermal stability [13,14]. Thus, many studies were conducted to produce a membrane with better characteristics. In 1979, Burns and coworkers [15] suggested the use of aromatic polyamide (PA) membrane, which is recognized by its cheap prize and high temperature tolerance, as an alternative [15,16]. Though water permeability by this membrane is less than that of cellulose acetate membrane, their salt rejection is higher.
A major breakthrough in the field of RO is the development of PA thin film composite (TFC) membrane. This membrane consists of two layers, the top is an active PA film prepared by the reaction between m-Phenylenediamine (MPD) and trimesoyl chloride (TMC) on a microporous polysulfone support (PSU), the bottom layer [17,18], as shown in Figure 1. Water flux through the composite membranes depends on the hydrophilicity of the membrane surface and the characteristics of the porous support layer, while salt rejection relies on the surface charges and PA structure [14].
Several state-of-the-art reviews have been published to highlight the types of RO-TFC membranes and their performance. For instance, Yin et al. [19] detailed the benefits of incorporating various nanoparticles on the membrane’s water flux, salt rejection, chlorine resistance, and antifouling properties. Another review by Xu et al. [20] discussed the influences of sub-layer adjustment on pressure gradient across the membrane and subsequent performance. Recently, Gohil et al. [21] reviewed the systematic development of TFC membranes with their structural composition and separation characteristics, including the effects of various additives and IP reaction parameters. However, until now, there has been no clear classification of the approaches that have been used to enhance RO-TFC membranes properties. Thus, the objective of this brief review is to fill this gap in literature and provide new insights for readers to improve their knowledge in this field.

2. Using Alternative Monomers to Prepare the Active Layer

Because membrane performance is substantially dependent on a thin film structure and its chemical properties, different monomers have been used to prepare the PA, as shown in Table 1. For example, Li et al. [22] used three different isomeric biphenyl acid chlorides (mm-BTCE, om-BTCE, op-PTCE) to react, separately, with m-phenylenediamine (MPD) on a porous support. Results indicated that the membrane prepared from op-PTCE exhibited higher water flux and lower salt rejection, while those prepared from mm-BTCE and om-BTCE showed lower water flux and higher salt rejection. The reason behind permeability enhancement could be due to the high density of the carboxylic acid group on the membrane prepared from op-PTCE, which led to better contact with water molecules. On the other hand, the higher salt rejection might be because of the thicker PA layer produced by using mm-BTCE and om-BTCE. Another study was reported by Wang et al. [23] in which introducing 3,5 diamino-N-(4-2-aminophenyl)-benzamide (DABA) as a monomer to react with TMC through interfacial polymerization resulted in a more hydrophilic, thinner, and smoother membrane.
One challenge facing RO TFC membrane durability is the degradation of the PA layer by chlorination [37]. Liu et al. [33] used three different polyacyl chlorides including 5-isocyanato-isophtahloyl chloride (ICIC), 5-chloroformyloxy-isophthaloyl chloride (CFIC), and TMC to prepare the TFC membrane with high chlorine tolerance. Results showed that the membrane prepared from MPD-CFIC and MPD-TMC possessed better chlorine stability when compared to MPD-ICIC membrane. It has been pointed out that the urea bond (NHCONH-) in MPD-ICIC could be easily attacked by chlorine. Recently, a composite membrane with high chlorine resistance has prepared through interfacial polymerization of hexafluoroalcohol (HFA)-aromatic diamine and trimesoyl chloride (TMC) [34]. The steric and electron withdrawing properties of HFA groups mitigated the probability of chlorine attack on the benzene rings or amide groups in the PA layer.
In term of fouling resistance, Hilal et al. [35,36] prepared composite membranes with improved antifouling properties by interfacial polymerization of bisphenol A (BPA) and trimesoyl chloride (TMC). This was attributed to the strong repulsion force between the negatively-charged bisphenol and organic foulants.

3. Modification of Membrane Surface

It was found that membrane performance is greatly affected by the treating steps that follow the synthesis process [38,39,40,41,42]. Chemical surface modifications are one of the promising post-treatment techniques that have been widely used to enhance TFC membrane surface properties. For example, Mickols and coworkers [43] used ethylenediamine and ethanolamine to increase membrane hydrophilicity. Their study showed that increasing the hydrogen bonding at the PA layer could enhance the interaction between water molecules and membrane surface, resulting in high water flux. Another study by Kuehne et al. [44] demonstrated that soaking the membrane in a solution containing glycerol promoted surface wettability and a 70% increase in water flux was obtained.
Wilf et al. [45] coated poly(vinyl alcohol) on TFC membrane surface to enhance fouling resistance and membrane durability. The modified membrane demonstrated better resistance against organic fouling when compared with the normal TFC membrane. Moreover, the membrane showed good permeability and long-term stability. The enhanced fouling resistance was ascribed to the lower rate of organics adsorption on the coated membrane. Coatings of poly(N,N-dimethylaminoethyl methacrylate) (PDMAEMA) have also exhibited enhanced chlorine resistance according to Kang et al. [46]. Additionally, Sakar et al. [47] used dendrimer-based coatings to reduce fouling effects. Recently, Ngo et al. [48] used redox initiated graft polymerization to coat TFC membrane with hydrophilic poly(acrylic acid). The coated membrane had lower roughness than the virgin one, and subsequently better fouling resistance and water flux was achieved.
Wu et al. [49,50] used gas plasma treatment to modify the TFC membrane. More carboxylic groups were introduced onto the surface by oxygen gaseous plasma treatment, which resulted in high water flux. On the other hand, argon plasma treatment improved chlorine resistance by introducing more amide groups onto the membrane surface. In addition, Lin et al. [51] demonstrated that the antifouling properties of the TFC membrane could be improved by using atmospheric gas plasma treatment. This kind of treatment created a polymeric brush at membrane surface which was capable of mitigating the attachment of organic foulant, Figure 2. However, plasma-induced grafting is a promising approach to produce a membrane with significant performance; however, it has not been thoroughly investigated and further research in this area is required.
Bing et al. [52] used redox initiation to enhance TFC membrane performance, especially chlorine resistance. Immersing the membrane in potassium persulfate (K2S2O8) solution initiated the interfacial crosslinking between the active film and PSU support, producing a thinner PA layer with more functional groups and denser cross linking. Reducing PA thickness led to enhanced water flux, as the water spent a shorter time to penetrate the membrane. On the other hand, increasing the crosslinking improved salt rejection by narrowing the passages for salt transportation. Moreover, the crosslinking reduced the N-chlorination sites on the membrane surface and hence, improved chlorine resistance.

4. Optimization of Polymerization Reactions

A major area of intense research is the optimization of interfacial polymerization reaction mechanisms such as kinetics, solvent solubility, reactant diffusion coefficient, reaction time, polymer molecular weight range, and characteristics of micro-porous support [53,54,55,56]. Tomaschke et al. [57] found that mixing amine salts with the casting solution formed a cross-linked membrane with an improved rejection. Chau et al. [58] added N,N-dimethyleformamide into a casting solution that introduced more carboxylic functional groups to PA layer, and eventually increased water flux. Kwak et al. [59] used dimethyl sulfoxide as an additive to modify the aromatic PA thin-film layer. The quantitative analysis of the surface morphology showed correlation between water permeability and both surface area and surface roughness; the flux improved with increasing roughness and surface area without a significant loss of salt rejection. Other researches showed that the addition of ethers, sulphur compounds, and alcohol- or water-soluble polymers to casting solution could produce high permeability without jeopardizing salt rejection [60,61,62].
Instead of modifying the casting solution, Michol et al. [63,64] succeeded in adding a complexing agent (phosphate containing compound) to the poly functional acyl halide prior to the substantial reaction between functional acyl halide and poly functional amide. It was thought that the addition of a complexing agent resulted in the formation of “association” with a polyfunctional acyl halide monomer capable of reducing the hydrolysis of acyl halide functional groups and permitting sufficient subsequent reaction with amine functional groups, thus resulting in a significant enhancement in membrane performance.
Another alternative approach for optimizing the polymerization reaction is to introduce surface-modified macromolecules (active additives) to acyl halide solution. This approach depends on the concept that the macromolecules may transfer to the PA film surface during the polymerization and change surface properties of membrane whilst maintaining bulk properties unaltered [65,66]. Arafat et al. found that by using poly(ethylene glycol) as an active additive in the interfacial polymerization, the water flux and salt rejection were significantly increased [67].

5. Incorporation of Nanoparticles (NPs) into Membrane PA Layer

A new class of membrane has been formed by the incorporation of nanoparticles (NPs) into the top layer of conventional thin film composite membrane (fabrication process Figure 3). Table 2 summarized the performance of RO thin film nanocomposite (TFN) membranes that were reported in literatures and the next section discusses the most important studies.
Jeong and Huang [60] reported that adding NaA zeolite NPs into the PA could result in an increase in water permeability without decreasing salt rejection. It was claimed that the superior hydrophilicity, high negative surface charge, and internal pores of zeolite nanomaterial facilitated water absorption and movement across the membrane, while maintaining high salt rejection via Donnan exclusion.
Lind et al. [62] studied the influence of zeolite crystal size on the apparent structure, morphology, interface, and permeability of zeolite-PA TFN membranes. The existence of zeolite NPs resulted in higher permeability, greater negative surface charge, and thicker PA when compared with the raw membrane. The smaller NPs produced greater permeability, while the larger NPs produced more favorable surface properties. This study implied that the size of NPs may be considered an additional “degree of freedom” in designing the nanostructured membranes. Recently, Mayyahi [106] used quantum dots as an ultra-small filler to modify the TFN. Both water flux and salt rejection were increased upon the addition of QDs.
Another study by Fathizadeh et al. [74] showed that increasing MPD and TMC concentrations to 3% w/v and 0.15% w/v, respectively, during TFN preparation in the presence of zeolite NPs formed a membrane with superior water flux but declined NaCl rejection. The low solute rejection was attributed to the poor dispersion of nanoparticles in the high molecular weight PA layer. The aggregation of NPs could have generated micro-holes in the PA, which allowed the brackish water to pass through. The reported results suggested that the relation between NPs and IP condition is another important factor that needs to be addressed.
In addition to zeolite, different NPs such as nano-silica [68,71,104], multiwall carbon nanotubes [69,72], zwitterion functionalized-carbon nanotubes [78], Titanium dioxide (TiO2) [101,107], and clay nano-sheets [90] have been used to modify the composite membranes. All these researches showed that imparting NPs to the PA could enhance membrane performance in terms of permeate flux, salt rejection, chlorine resistance, and antifouling properties. For instant, Barona et al. [86] found that incorporating aluminosilicate single-wall carbon nanotubes (SWNTs) into the membrane surface resulted in a significant increase in water flux without affecting salt rejection. The functional groups on SWCNTs secured excellent dispersion of fillers in the PA and, as a result, enhanced the overall performance. A remarkable enhancement in membrane performance was achieved in another study by Jun et al. [76] upon the addition of MCM-41 silica NPs without compromising salt rejection. The high water permeability was ascribed to the enhanced membrane hydrophilicity as well as the pores in the NPs that imparted extra channels for water transportation.
It is known that PA composite membranes are very sensitive to chlorine. As the PA layer touches the chlorinated water, the amine groups oxidize by chlorine and decompose in water, leading to deteriorated separation performance [108]. Park et al. [69] used acid functionalized MWCNTs to improve the chlorine resistance. When MWCNTs were incorporated into the PA active layer, the membrane showed enhanced anti-chlorine properties. This could be ascribed to the reaction between the functional groups in carbon nanotubes and the amine groups in PA structure, which as a result formed a barrier above the PA that reduced membrane chlorine exposure. Another study by Kim et al. [85] showed that attaching hyper branched polyamide modified silica NPs onto PA layer could protect the membrane from chlorine attack. The extra amino groups presented by the functionalized silica NPs were the main target for chlorine and subsequently lessened membrane surface exposure, as shown in Figure 4. It seems that all researchers followed the same strategy to produce a membrane with high chlorine resistance, which generates a protection layer on the membrane surface; however, this could not provide long term efficiency as the barrier might be finally degraded and the chlorine reaches the membrane surface [82,87,109].
Membrane fouling is generally known as the accumulation of unwanted materials on membrane surfaces [110]. Organic and micro-biological fouling of TFN membranes are among the substantial reasons that lead to membrane performance declination [111]. Hence, many studies are devoted to develop a TFC with desired fouling resistance without “trading off” any of the other properties including permeability and rejection efficiency. Kim et al. [112] showed that incorporating hydrophilic filler into the PA layer could increase membrane resistance against organic fouling. The results demonstrated that there was a reverse relationship between hydrophilicity and organic foulants accumulation. This could be ascribed to the weak interaction between organic foulants and hydrophilic surfaces. Another study by Rana et al. [75] exhibited that an increase in membrane’s negative surface charge could enhance fouling resistance, due to the strong repulsive force between the membrane and negatively charged foulants. Lee et al. [113] used Ag nanoparticles as fillers in the PA to mitigate bacterial accumulation on the surface. Results showed that the Ag-TFN membrane has better resistance against bacterial fouling. It is believed that Ag nanoparticles disturb the permeability and respiration functions of the bacterial cell, and eventually destroy the DNA [114,115,116].
Kim et al. used a new approach to prepare hybrid TFC by the self-assembly between titanium oxide NPs and PA’s carboxylic functional groups [117]. Results indicated that the UV irradiation of the membrane could reduce E-coli content on the surface and this was attributed to the ability of TiO2 to form different hydroxyl and peroxide radicals under the influence of UV light. These active radicals were capable of destroying the bacterial cells. Ben-Sasson et al. [116] used electrostatic attraction to attach copper (Cu) nanoparticles to the TFC membrane surface. Results indicated that the presence of positively charged Cu-NPs did not affect the overall hydrophilicity of the membrane, but reduced the growth of bacterial cells. The SEM images of the membrane’s surface exhibited that the bacterial cells were damaged when contacted with Cu-NPs. This could be ascribed to the high toxicity of Cu that led to bacterial DNA damage. Choi et al. [118] used “layer-by-layer assembly” to attach graphene oxide (GO) and aminated-graphene oxide (AGO) to TFC membrane surface, as shown in Figure 5. The resultant TFC showed enhanced resistance against organic fouling and chlorine attack, while preserving water flux and NaCl rejection. Hu and Mi [119] succeeded in using layer-by-layer deposition” to connect GO NPs to the PA. In this case, GO-NPs formed linkages with membrane’s functional groups. The newly developed membrane exhibited superior water flux and excellent dye rejection. The disadvantage of the “surface located nanocomposite membrane” is the loss of nanoparticles during filtration, especially those attached by electrostatic forces. The depletion of nanoparticles reduces the efficiency of the membrane and may expose nanoparticles to the permeated water causing a threat to people’s health. As a result, Yin and co-workers. [120] used cyseteamine as a “bridging agent” to attach silver (Ag) nanoparticles to the membrane surface, as shown in Figure 6. Results indicated that the modified membrane has stable Ag NPs, superior antimicrobial properties, high permeability, and good separation efficiency. Recently, Mayyahi [121] showed that UV irradiation of TFN membrane which impregnated with TiO2 could result in robust antibacterial properties.

6. Conclusions

A tremendous development in TFN membranes for water purification has been achieved including producing a membrane with super water flux, high salt rejection, and excellent fouling and chorine resistance via using innovative approaches such as imparting the favored properties of nanoparticles to the membrane surface, optimizing the membrane fabrication process, modifying the materials that are required to synthesis the membrane, and changing membrane surface properties by post-treatment. However, researchers have failed to find an alternative to the PA barrier layer or to suggest a new support layer. We do agree that PA atop PSU/PES showed robust efficiency in RO and other water treatments applications, but these membranes have been used since 1970 and scholars have successfully addressed almost all the challenges facing the progress of such membranes. Forthcoming researches should be dedicated to suggest a new reverse osmosis membrane rather than developing the existing one.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

AGOAminated-graphene oxide
AgSilver
Al-ZnOAluminum doped zinc oxide
BDSA2,2′-benzidinedisulfonic acid
BHAC2,2′,4,4′,6,6′-biphenyl hexaacyl chloride
BHDTBis-2,6-N,N-(2-hydroxyethyl) diaminotoluene
BPABisphenol
CFIC5-chloroformyloxy-isophthaloyl chloride
CNTCarbon nanotube
CQDsCarbon quantum dots
CuCopper
DABATriamine 3,5-diamino-N-(4-aminophenyl)-benzamide
DMFN,N-dimethylformamide
DMSO2,4,6-pyridinetricarboxylic acid chloride
DETADiethylenetriamine
DPADopamine
F-MWCNTsFunctionalized multi wall carbon nanotubes
F-silicaFunctionalized silica
GOGraphene oxide
HBP-g-silicaHyper-branched aromatic polyamide-grafted silica
HNTsHalloysite nanotubes
ICIC5-isocyanato-isophtahloyl chloride
IPInterfacial polymerization
iLSMMIn-situ hydrophilic surface modifying macromolecules
MOFsMetal–organic frameworks
MPDm-phenylenediamine
mm-BTEC3,3′,5,5′-biphenyl tetraacyl chloride
MWCNTsMultiwall carbon nanotubes
NPsNanoparticles
OA-SiO2Oleic acid modified silica
om-BTEC2,2′,4,4′-biphenyl tetraacyl chloride
MOFsMetal–organic framework
op-BTEC2,2′,5,5′-biphenyl tetraacyl chloride
PAPolyamide
PAMAMEthylenediamine cored poly(amidoamine)
PDMAEMAPoly(N,N-dimethylaminoethyl methacrylate)
PSUPolysulfone
ROReverse osmosis
SiO2Silicon dioxide
SMPDSulfonated m-phenylenediamine
SWCNTsSingle-wall carbon nanotubes
TFCThin film composite
TFNThin film nanocomposite
TiO2Titanium dioxide
TMCTrimesoyl chloride
TETATriethylenetetramine
TEPATetraethylenepentamine
TNTsTitanate nanotubes
UVUltraviolet
ZIFZeolitic imidazolate framework
ZnOZinc oxide

References

  1. Marry, P.; Hoek, E. A review of water treatment membrane nanotechnologies. Energy Environ. Sci. 2011, 4, 1946–1971. [Google Scholar]
  2. Elimelech, M.; Phillip, W.A. The future of seawater desalination: energy, technology, and the environment. Science 2011, 333, 712–717. [Google Scholar] [CrossRef] [PubMed]
  3. Li, N.; Fane, A.; Hu, W.; Matsuura, T. Advanced Membrane Technology and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2008. [Google Scholar]
  4. Shenvi, S.; Isloor, A.; Ismail, A. A review on RO membrane technology: Developments and challenges. Desalination 2015, 368, 10–26. [Google Scholar] [CrossRef]
  5. Malaeb, L.; Ayoub, G. Reverse osmosis technology for water treatment: State of the art review. Desalination 2011, 267, 1–8. [Google Scholar] [CrossRef]
  6. Mai, Z. Membrane Processes for Water and Wastewater Treatment: Study and Modeling of Interactions between Membrane and Organic Matter. Ph.D. Thesis, LGPM—Laboratoire de Génie des Procédés et Matériaux, Paris, France, 2013. [Google Scholar]
  7. Wijmans, J.G.; Baker, R.W. The solution-diffusion model: A review. J. Membr. Sci. 1995, 107, 1–21. [Google Scholar] [CrossRef]
  8. Soltanieh, M.; Gill, W.N. Review of reverse osmosis membrane and transport models. Chem. Eng. Commun. 1981, 12, 279–363. [Google Scholar] [CrossRef]
  9. Lee, C.H. Theory of reverse osmosis and some other membrane permeation operations. J. Appl. Polym. Sci. 1975, 19, 83–95. [Google Scholar] [CrossRef]
  10. Reid, C.E.; Breton, E.J. Water and ion flow across cellulosic membranes. J. Appl. Polym. Sci. 1959, 1, 133–143. [Google Scholar] [CrossRef]
  11. Sourirajan, S. Separation of hydrocarbon liquids by flow under pressure through porous membranes. Nature 1964, 203, 1348–1349. [Google Scholar] [CrossRef]
  12. Loeb, S. Sea Water Demineralization by Means of a Semipermeable Membrane: Progress Report July 1, 1962-December 31; University of California: Berkeley, CA, USA, 1963. [Google Scholar]
  13. Porter, M.C. What, when and why of membranes MF, UF, and RO. ALChEsymp. Ser. 1977, 73, 83–103. [Google Scholar]
  14. Burns and Roe Industrial Services Corporation. Reverse Osmosis Technical Manual; United States Office of Water Research and Technology: Washington, DC, USA, 1979. [Google Scholar]
  15. Belfort, G. Chapter 6: Pressure driven membrane processes and wastewater renovation in water renovation and reuse. In Water Renovation and Reuse, 1st ed.; Shuval, H., Ed.; Academic Press: Cambridge, MA, USA, 1977. [Google Scholar]
  16. Shields, C.P. Five years’ experience with reverse osmosis systems using DU PONT “Permasp” permeators. Desalination 1979, 28, 157–179. [Google Scholar] [CrossRef]
  17. Cadotte, J.E. Interfacially Synthesized Reverse Osmosis Membrane. U.S. Patent 4,277,344, 7 July 1979. [Google Scholar]
  18. Khorshidi, B.; Thundat, T.; Fleck, B.A.; Sadrzadeh, M. A novel approach toward fabrication of high performance thin film composite polyamide membranes. Sci. Rep. 2016, 6, 22029. [Google Scholar] [CrossRef] [PubMed]
  19. Yin, J.; Deng, B. Polymer-matrix nanocomposite membranes for water treatment. J. Membr. Sci. 2015, 479, 256–275. [Google Scholar] [CrossRef]
  20. Xu, G.R.; Xu, J.M.; Feng, H.J.; Zhao, H.L.; Wu, S.B. Tailoring structures and performance of polyamide thin film composite (PA-TFC) desalination membranes via sublayers adjustment—A review. Desalination 2017, 417, 19–35. [Google Scholar] [CrossRef]
  21. Gohil, J.M.; Ray, P. A review on semi-aromatic polyamide TFC membranes prepared by interfacial polymerization: Potential for water treatment and desalination. Sep. Purif. Technol. 2017, 181, 159–182. [Google Scholar] [CrossRef]
  22. Li, L.; Zhang, S.; Zhang, X.; Zheng, G. Polyamide thin film composite membranes prepared from isomeric biphenyl tetraacyl chloride and m-phenylenediamine. J. Membr. Sci. 2008, 315, 20–27. [Google Scholar] [CrossRef]
  23. Wang, H.; Li, L.; Zhang, X.; Zhang, S. Polyamide thin-film composite membranes prepared from a novel triamine 3,5-diamino-N-(4-aminophenyl)-benzamide monomer and m-phenylenediamin. J. Membr. Sci. 2010, 353, 78–84. [Google Scholar] [CrossRef]
  24. Nathaniel, G.; Lim, J.; Jung, B. High performance thin film composite polyamide reverse osmosis membrane prepared via m-phenylenediamine and 2,2-benzidinedisulfonic acid. Desalination 2012, 291, 69–77. [Google Scholar]
  25. Xie, W.; Geise, G.; Freeman, B.; Lee, H.; Byun, G.; McGrath, J. Polyamide interfacial composite membranes prepared from m-phenylene diamine, trimesoyl chloride and a new disulfonated diamine. J. Membr. Sci. 2012, 404, 152–161. [Google Scholar] [CrossRef]
  26. Zhang, Z.; Wang, S.; Chen, H.; Wang, T. Preparation of polyamide membranes with improved chlorine resistance by bis-2,6-N,N-(2-hydroxyethyl) diaminotoluene and trimesoyl chloride. Desalination 2013, 331, 16–25. [Google Scholar] [CrossRef]
  27. Sum, J.Y.; Ahmad, A.; Ooi, B.S. Synthesis of thin film composite membrane using mixed dendritic poly(amidoamine) and void filling piperazine monomers. J. Membr. Sci. 2014, 466, 183–191. [Google Scholar] [CrossRef]
  28. Li, Y.; Su, Y.; Dong, Y.; Zhao, X.; Jiang, Z.; Zhang, R.; Zhao, J. Separation performance of thin-film composite nanofiltration membrane through interfacial polymerization using different amine monomers. Desalination 2014, 333, 59–65. [Google Scholar] [CrossRef]
  29. Zhao, J.; Su, Y.; He, X.; Zhao, X.; Li, Y.; Zhang, R.; Jiang, Z. Dopamine composite nanofiltration membranes prepared by self-polymerization and interfacial polymerization. J. Membr. Sci. 2014, 465, 41–48. [Google Scholar] [CrossRef]
  30. Jewrajka, S.; Reddy, A.V.R.; Rana, H.; Mandal, S.; Khullar, S.; Haldar, S.; Joshi, N.; Ghosh, P. Use of 2,4,6-pyridinetricarboxylic acid chloride as a novel co-monomer for the preparation of thin film composite polyamide membrane with improved bacterial resistance. J. Membr. Sci. 2013, 439, 87–95. [Google Scholar] [CrossRef]
  31. Wang, T.; Dai, L.; Zhang, Q.; Li, A.; Zhang, S. Effects of acyl chloride monomer functionality on the properties of polyamide reverse osmosis (RO) membrane. J. Membr. Sci. 2013, 440, 48–57. [Google Scholar] [CrossRef]
  32. Yong, Z.; Sanchuan, Y.; Meihong, L.; Congjie, G. Polyamide thin film composite membrane prepared from m-phenylenediamine and m-pheneylenediamine-5-sulfonic acid. J. Membr. Sci. 2006, 270, 162–168. [Google Scholar] [CrossRef]
  33. Liu, M.; Wu, D.; Yu, S.; Gao, C. Influence of the polyacyl chloride structure on the reverse osmosis performance, surface properties and chlorine stability of the thin-film composite polyamide membranes. J. Membr. Sci. 2009, 326, 205–214. [Google Scholar] [CrossRef]
  34. La, Y.; Sooriyakum, R.; Miller, D.; Fujiwara, M.; Freeman, B.; Allen, R. Novel thin film composite membrane containing ionizable hydrophobes: pH-dependent reverse osmosis behavior and improved chlorine resistance. J. Mater. Chem. 2010, 20, 4615–4620. [Google Scholar] [CrossRef]
  35. Abu Seman, M.; Khayet, M.; Hilal, N. Nanofiltration thin-film composite polyester polyethersulfone-based membranes prepared by interfacial polymerization. J. Membr. Sci. 2010, 348, 109–116. [Google Scholar] [CrossRef]
  36. Abu Seman, M.; Khayet, M.; Hilal, N. Development of antifouling properties and performance of nanofiltration membranes modified by interfacial polymerization. Desalination 2011, 273, 36–47. [Google Scholar] [CrossRef]
  37. Koo, J.; Petersen, R.; Cadotte, J. ESCA characterization of chlorine-damage polyamide reverse osmosis membrane. ACS Polym. Prepr. 1986, 27, 391–392. [Google Scholar]
  38. Liu, M.; Zhou, C.; Dong, B.; Wu, Z.; Wang, L.; Yu, S.; Gao, C. Enhancing the permselectivity of thin-film composite poly(vinyl alcohol) (PVA) nanofiltration membrane by incorporating poly(sodium-p-styrene-sulfonate) (PSSNa). J. Membr. Sci. 2014, 463, 173–182. [Google Scholar] [CrossRef]
  39. Schafer, A.L.; Fane, A.G.; Waite, T.D. Nanofiltration: Principles and Application, 1st ed.; Elservier: New York, NY, USA, 2003. [Google Scholar]
  40. Petersen, R.J. Composite reverse osmosis and nanofiltration membrane. J. Membr. Sci. 1993, 38, 81–150. [Google Scholar] [CrossRef]
  41. Hilal, N.; Al-Zoubi, H.; Darwish, N.A.; Mohamma, A.W.; Arabi, M.A. A comprehensive review of nanofiltration membranes: Treatment, pretreatment, modeling and atomic force microscopy. Desalination 2004, 170, 281–308. [Google Scholar] [CrossRef]
  42. Lau, W.J.; Ismail, A.F. Polymeric nanofiltration membrane for textile dyeing waste treatment: preparation, performance evaluation, transport modeling, and fouling control—A review. Desalination 2010, 245, 4551–4566. [Google Scholar]
  43. Mickols, W.E. Method of Treating Polyamide Membranes to Increase Flux. U.S. Patent 5,755,964, 26 May 1998. [Google Scholar]
  44. Kuehne, M.A.; Song, R.Q.; Li, N.N.; Petersen, R.J. Flux enhancement in TFC RO membranes. Environ. Prog. 2001, 20, 23–26. [Google Scholar] [CrossRef]
  45. Wilf, M.; Alt, S. Application of low fouling RO membrane elements for reclamation of municipal wastewater. Desalination 2000, 132, 11–19. [Google Scholar] [CrossRef]
  46. Kang, G.D.; Gao, C.J.; Chen, W.D.; Jie, X.M.; Cao, Y.M. Study on hypochlorite degradation of aromatic polyamide reverse osmosis membrane. J. Membr. Sci. 2007, 300, 165–171. [Google Scholar] [CrossRef]
  47. Sarkar, A.; Carver, P.I.; Zhang, T.; Merrington, A.; Bruza, K.J.; Rousseau, J.L.; Keinath, S.E.; Dvornic, P.R. Dendrimer-based coatings for surface modification of polyamide reverse osmosis membranes. J. Membr. Sci. 2010, 349, 165–171. [Google Scholar] [CrossRef]
  48. Hong Anh Ngo, T.; Dinh Do, K.; Thi Tran, D. Surface modification of polyamide TFC membranes via redox-initiated graft polymerization of acrylic acid. J. Appl. Polym. Sci. 2017, 134, 45110. [Google Scholar] [CrossRef]
  49. Wu, S.; Xing, J.; Zheng, C.; Xu, G.; Zheng, G.; Xu, J. Plasma modification of aromatic polyamide reverse osmosis composite membrane surface. J. Appl. Polym. Sci. 1997, 764, 1923–1926. [Google Scholar] [CrossRef]
  50. Gilman, A.B. Low temperature plasma treatment as an effective method for surface modification of polymeric materials. High Energy Chem. 2003, 37, 17–23. [Google Scholar] [CrossRef]
  51. Lin, N.H.; Kim, M.M.; Lewis, G.T.; Cohen, Y. Polymer surface nano-structuring of reverse osmosis membrane for fouling resistance and improved flux performance. J. Membr. Sci. 2010, 20, 4642–4652. [Google Scholar] [CrossRef]
  52. Bing, S.; Wang, J.; Xu, H.; Zhao, Y.; Zhou, Y.; Zhang, L.; Gao, C.; Hou, L.A. Polyamide thin-film composite membrane modified with persulfate for improvement of perm-selectivity and chlorine-resistance. J. Membr. Sci. 2018, 555, 318–326. [Google Scholar] [CrossRef]
  53. Song, Y.; Sun, P.; Henry, L.L.; Sun, B. Mechanism of structure and performance controlled thin film composite membrane formation via interfacial polymerization process. J. Membr. Sci. 2005, 251, 67–79. [Google Scholar] [CrossRef]
  54. Karode, S.K.; Kulkarni, S.S.; Suresh, A.K.; Mashelkar, R.A. New insight into kinetics and thermodynamics of interfacial polymerization. Chem. Eng. Sci. 1998, 53, 2649–2663. [Google Scholar] [CrossRef]
  55. Dhumal, S.S.; Wagh, S.J.; Suresh, A.K. Interfacial polymerization-modeling of kinetics and film properties. J. Membr. Sci. 2008, 352, 758–771. [Google Scholar] [CrossRef]
  56. Ghosh, A.K.; Hoek, E.M.V. Impact of reaction and curing condition on polyamide composite reverse osmosis membrane properties. J. Membr. Sci. 2008, 311, 34–45. [Google Scholar] [CrossRef]
  57. Tomaschke, J.E. Interfacially Synthesized Reverse Osmosis Membrane Containing an Amine Salt and Processes for Preparing the Same. U.S. Patent 4,948,507, 14 August 1990. [Google Scholar]
  58. Chau, M.M.; Light, W.G.; Chu, H.C. Dry High Flux Semipermeable Membrane. U.S. Patent 4,983,291, 18 January 1991. [Google Scholar]
  59. Kwak, S.Y.; Jung, S.G.; Kim, S.H. Structure-motion-performance relationship of flux enhanced reverse osmosis (RO) membranes composed of aromatic polyamide thin films. Environ. Sci. Technol. 2001, 35, 4334–4340. [Google Scholar] [CrossRef] [PubMed]
  60. Jeong, B.-H.; Hoek, E.M.V.; Yan, Y.; Subramani, A.; Huang, X.; Hurwitz, G.; Ghosh, A.K.; Jawor, A. Interfacial polymerization of thin film nanocomposites: A new concept for reverse osmosis membranes. J. Membr. Sci. 2007, 294, 1–7. [Google Scholar] [CrossRef]
  61. Lind, M.L.; Jeong, B.H.; Subramani, A.; Huang, X.; Hoek, E.M.V. Effect of mobile cation on zeolite-polyamide thin film nanocomposite membranes. J. Mater. Res. 2009, 24, 1624–1631. [Google Scholar] [CrossRef]
  62. Lind, M.L.; Ghosh, A.K.; Jawor, A.; Huang, X.; Hou, W.; Yang, Y.; Hoek, E.M.V. Influence of zeolite crystal size on zeolite-polyamide thin film nanocomposite membranes. Langmuir 2009, 25, 10139–10145. [Google Scholar] [CrossRef] [PubMed]
  63. Mickols, W.E. Composite Membrane and Method for Making the Same. U.S. Patent 6,878,278, 12 April 2005. [Google Scholar]
  64. Mickols, W.E. Composite Membrane and Method for Making the Same. U.S. Patent 6,337,018, 8 January 2002. [Google Scholar]
  65. Khayet, M. Membrane surface modification and characterization by X-ray Photo electron spectroscopy, atomic force microscopy and contact angle measurement. Appl. Surf. Sci. 2004, 238, 269–272. [Google Scholar] [CrossRef]
  66. Khayet, M.; Suk, D.E.; Narbaitz, R.M.; Santerre, J.P. study on surface modification by surface modifying macromolecules and its application in membrane separation process. Appl. Polym. 2003, 89, 2902–2961. [Google Scholar] [CrossRef]
  67. Tarboush, B.J.A.; Ranan, D.; Matsuura, T.; Nabratiz, H.A. Preparation of thin film composite polyamide membrane for water desalination using novel hydrophilic surface modifying macromolecules. J. Membr. Sci. 2008, 325, 166–175. [Google Scholar] [CrossRef]
  68. Jadav, G.L.; Singh, P.S. Synthesis of novel silica-polyamide nanocomposite membrane with enhanced properties. J. Membr. Sci. 2009, 328, 257–267. [Google Scholar] [CrossRef]
  69. Park, J.; Choi, W.; Kim, S.H.; Chun, B.H.; Bang, J.; Lee, K.B. Enhancement of chlorine resistance in carbon nanotube-based nanocomposite reverse osmosis membranes. Desalin. Water Treat. 2010, 15, 198–204. [Google Scholar] [CrossRef]
  70. Lind, M.L.; Suk, D.E.; Nguyen, T.V.; Hoek, E.M.V. Tailoring the structure of thin film nanocomposite membranes to achieve seawater RO membrane performance. Environ. Sci. Technol. 2010, 44, 8230–8235. [Google Scholar] [CrossRef] [PubMed]
  71. Jadav, G.L.; Aswal, V.K.; Singh, P.S. SANS study to probe nanoparticle dispersion in nanocomposite membranes of aromatic polyamide and functionalized silica nanoparticles. J. Colloid. Interface Sci. 2010, 351, 304–314. [Google Scholar] [CrossRef] [PubMed]
  72. Roy, S.; Ntim, S.A.; Mitra, S.; Sirkar, K.K. Facile fabrication of superior nanofiltration membranes from interfacially polymerized CNT-polymer composites. J. Membr. Sci. 2011, 375, 81–87. [Google Scholar] [CrossRef]
  73. Kong, C.; Koushima, A.; Kamada, T.; Shintani, T.; Kanezashi, M.; Yoshioka, T.; Tsuru, T. Enhanced performance of inorganic-polyamide nanocomposite membranes prepared by metal-alkoxide-assisted interfacial polymerization. J. Membr. Sci. 2011, 366, 382–388. [Google Scholar] [CrossRef]
  74. Fathizadeh, M.; Aroujalian, A.; Raisi, A. Effect of added NaX nano-zeolite into polyamide as a top thin layer of membrane onwater flux and salt rejection in a reverse osmosis process. J. Membr. Sci. 2011, 375, 88–95. [Google Scholar] [CrossRef]
  75. Rana, D.; Kim, Y.; Matsuura, T.; Arafat, H.A. Development of antifouling thin film-composite membranes for seawater desalination. J. Membr. Sci. 2011, 367, 110–118. [Google Scholar] [CrossRef]
  76. Yin, J.; Kim, E.S.; Yang, J.; Deng, B. Fabrication of a novel thin-film nanocomposite (TFN) membrane containing MCM-41 silica nanoparticles (NPs) for water purification. J. Membr. Sci. 2012, 423, 238–246. [Google Scholar] [CrossRef]
  77. Zhao, Y.; Qiu, C.; Li, X.; Vararattanavech, A.; Shen, W.; Torres, J.; Hélix-Nielsen, C.; Wang, R.; Hu, X.; Fane, A.G.; et al. Synthesis of robust and high performance aquaporin-based biomimetic membranes by interfacial polymerization-membrane preparation and RO performance characterization. J. Membr. Sci. 2012, 423, 422–428. [Google Scholar] [CrossRef]
  78. Chan, W.F.; Chen, H.Y.; Surapathi, A.; Taylor, M.G.; Shao, X.; Marand, E.; Johnson, J.K. Zwitterion functionalized carbon nanotube/polyamide nanocomposite membranes for water desalination. ACS Nano 2013, 7, 5308–5319. [Google Scholar] [CrossRef] [PubMed]
  79. De Lannoy, C.F.; Jassby, D.; Gloe, K.; Gordon, A.D.; Wiesner, M.R. Aquatic biofouling prevention by electrically charged nanocomposite polymer thin film membranes. Environ. Sci. Technol. 2013, 47, 2760–2768. [Google Scholar] [CrossRef] [PubMed]
  80. Huang, H.; Qu, X.; Ji, X.; Gao, X.; Zhang, L.; Chen, H.; Hou, L. Acid and multivalent ion resistance of thin film nanocomposite RO membranes loaded with silicalite-1 nanozeolites. J. Mater. Chem. A 2013, 1, 11343–11349. [Google Scholar] [CrossRef]
  81. Huang, H.; Qu, X.; Dong, H.; Zhang, L.; Chen, H. Role of NaA zeolites in the interfacial polymerization process towards a polyamide nanocomposite reverse osmosis membrane. RSC Adv. 2013, 3, 8203–8207. [Google Scholar] [CrossRef]
  82. Kim, S.G.; Hyeon, D.H.; Chun, J.H.; Chun, B.H.; Kim, S.H. Nanocomposite poly (arylene ether sulfone) reverse osmosis membrane containing functional zeolite nanoparticles for seawater desalination. J. Membr. Sci. 2013, 443, 10–18. [Google Scholar] [CrossRef]
  83. Pendergast, M.M.; Ghosh, A.K.; Hoek, E.M.V. Separation performance and interfacial properties of nanocomposite reverse osmosis membranes. Desalination 2013, 308, 180–185. [Google Scholar] [CrossRef]
  84. Bao, M.; Zhu, G.; Wang, L.; Wang, M.; Gao, C. Preparation of monodispersed spherical mesoporous nanosilica-polyamide thin film composite reverse osmosis membranes via interfacial polymerization. Desalination 2013, 309, 261–266. [Google Scholar] [CrossRef]
  85. Kim, S.G.; Chun, J.H.; Chun, B.H.; Kim, S.H. Preparation, characterization and performance of poly(aylene ether sulfone)/modified silica nanocomposite reverse osmosis membrane for seawater desalination. Desalination 2013, 325, 76–83. [Google Scholar] [CrossRef]
  86. Baroña, G.N.B.; Lim, J.; Choi, M.; Jung, B. Interfacial polymerization of polyamide-aluminosilicate SWNT nanocomposite membranes for reverse osmosis. Desalination 2013, 325, 138–147. [Google Scholar] [CrossRef]
  87. Zhao, H.; Qiu, S.; Wu, L.; Zhang, L.; Chen, H.; Cao, C. improving the performance of polyamide reverse osmosis membrane by incorporating of modified multi wall carbon nanotubes. J. Membr. Sci. 2014, 450, 249–256. [Google Scholar] [CrossRef]
  88. Ghanbaria, M.; Emadzadeh, D.; Lau, W.J.; Matsuura, T.; Ismail, A.F. Synthesis and characterization of novel thin film nanocomposite reverse osmosis membranes with improved organic fouling properties for water desalination. RSC Adv. 2015, 5, 21268–21276. [Google Scholar] [CrossRef]
  89. Rakhshan, N.; Pakizeh, M. The effect of chemical modification of SiO2 nanoparticles on the nanofiltration characteristics of polyamide membrane. J. Membr. Sci. 2015, 32, 2524–2533. [Google Scholar] [CrossRef]
  90. Dong, H.; Wu, L.; Zhang, L.; Chen, H.; Cao, C. Clay nano-sheet as charged filler materials for highperformance and fouling resistance thin film nanocomposite membrane. J. Membr. Sci. 2015, 494, 92–103. [Google Scholar] [CrossRef]
  91. Safarpour, M.; Khataee, A.; Vatanpour, V. Thin film nanocomposite reverse osmosis membrane modified by reduced graphene oxide/ TiO2 with improved desalination performance. J. Membr. Sci. 2015, 489, 43–54. [Google Scholar] [CrossRef]
  92. Emadzadeh, D.; Lau, W.J.; Rahbara, R.; Danseshfar, A.; Mayahi, A.; Matsuura, T.; Ismail, A.F. A novel thin film nanocomposite reverse osmosis membrane with superior anti organic fouling affinity for water desalination. Desalination 2015, 368, 106–113. [Google Scholar] [CrossRef]
  93. Bano, S.; Mahmood, A.; Kim, S.; Lee, K. Graphene oxide modified polyamide nanofiltration membrane with improved flux and antifouling properties. J. Mater. Chem. 2015, 3, 2065–2071. [Google Scholar] [CrossRef]
  94. Al Hobibi, A.S.; Ghoul, J.; Chiloufi, I.; EL Mir, L. Synthesis and characterization of polyamide thin film nanocomposite membrane reached by aluminum doped ZnO. Mater. Sci. Semicond. Process. 2016, 42, 111–114. [Google Scholar] [CrossRef]
  95. Liu, L.; Zhu, G.; Liu, Z.; Gao, C. effect of MCM-48 nanoparticles on the performance of TFN membrane for reverse osmosis application. Desalination 2016, 394, 72–82. [Google Scholar] [CrossRef]
  96. Yin, J.; Deng, B. Graphene oxide enhanced polyamide thin-film nanocomposite membrane for water purification. Desalination 2016, 379, 93–101. [Google Scholar] [CrossRef]
  97. Mayyahi, A.A.; Deng, B. Efficient water desalination using photo-responsive ZnO polyamide thin film nanocomposite membrane. Environ. Chem. Lett. 2018, 1–7. [Google Scholar] [CrossRef]
  98. Kadhom, M.; Deng, B. Thin film nanocomposite membrane filled with metal-organic frameworks UiO-66 and MIL-125 nanoparticles for water desalination. Membrane 2017, 7, 31. [Google Scholar] [CrossRef] [PubMed]
  99. Bi, R.; Zhang, Q.; Zhang, R.; Su, Y.; Jiang, Z. Thin Film Nanocomposite Membranes Incorporated with Graphene Quantum Dots for High Flux and Antifouling Property. J. Membr. Sci. 2018, 553, 17–24. [Google Scholar] [CrossRef]
  100. Aljundi, I.H. Desalination characteristics of TFN-RO membrane incorporated with ZIF-8 nanoparticles. Desalination 2017, 430, 12–20. [Google Scholar] [CrossRef]
  101. Khorshidi, B.; Biswas, I.; Ghosh, T.; Thundat, T.; Sadrzadeh, M. Robust fabrication of thin film polyamide-TiO2 nanocomposite membranes with enhanced thermal stability and anti-biofouling propensity. Sci. Rep. 2018, 8, 784. [Google Scholar] [CrossRef] [PubMed]
  102. Sun, H.; Wu, P. Tuning the functional groups of carbon quantum dots in thin film nanocomposite membranes for nanofiltration. J. Membr. Sci. 2018, 564, 394–403. [Google Scholar] [CrossRef]
  103. He, Y.; Zhao, D.L.; Chung, T.S. Na+ functionalized carbon quantum dot incorporated thin-film nanocomposite membranes for selenium and arsenic removal. J. Membr. Sci. 2018, 564, 483–491. [Google Scholar] [CrossRef]
  104. Peyki, A.; Rahimpour, A.; Jahanshahi, M. Preparation and characterization of thin film composite reverse osmosis membranes incorporated with hydrophilic SiO2 nanoparticles. Desalination 2015, 368, 152–158. [Google Scholar] [CrossRef]
  105. Ma, D.; Peh, S.P.; Han, G.; Chen, S.B. Thin-Film Nanocomposite (TFN) Membranes Incorporated with Super-Hydrophilic Metal–Organic Framework (MOF) UiO-66: Toward Enhancement of Water Flux and Salt Rejection. ACS Appl. Mater. Interfaces 2017, 9, 7523–7534. [Google Scholar] [CrossRef] [PubMed]
  106. Mayyahi, A.A. Thin-film composite (TFC) membrane modified by hybrid ZnO-graphene nanoparticles (ZnO-Gr NPs) for water desalination. J. Environ. Chem. Eng. 2018, 6, 1109–1117. [Google Scholar] [CrossRef]
  107. Rajaeian, B.; Rahimpour, A.; Tade, M.O. Fabrication and characterization of polyamide thin film nanocomposite (TFN) nanofiltration membrane impregnated with TiO2 nanoparticles. Desalination 2013, 313, 176–188. [Google Scholar] [CrossRef]
  108. Glater, J.; Hong, S.; Elimelech, M. The search for a chlorine-resistance reverse osmosis membrane. Desalination 1994, 95, 325–345. [Google Scholar] [CrossRef]
  109. Xue, S.X.; Ji, C.H.; Xu, Z.L.; Tang, Y.J.; Li, R.H. Chlorine resistant TFN nanofiltration membrane incorporated with octadecylamine-grafted GO and fluorine-containing monomer. J. Membr. Sci. 2018, 545, 185–195. [Google Scholar] [CrossRef]
  110. Nikolova, J.D.; Islam, M.A. Contribution to adsorbed layer resistance to flux decline in ultrafiltration process. J. Membr. Sci. 1998, 146, 105–111. [Google Scholar] [CrossRef]
  111. Sadar Ghayeni, S.B.; Beatson, P.J.; Schneider, R.P.; Fane, A.G. Adhesive of water bacteria to reverse osmosis membrane. J. Membr. Sci. 1998, 138, 29–42. [Google Scholar] [CrossRef]
  112. Kim, E.-S.; Deng, B. Fabrication of polyamide thin-film-nano-composite (PA-TFN) membrane with hydrophilized ordered mesoporous carbon for water purification. J. Membr. Sci. 2011, 375, 46–54. [Google Scholar] [CrossRef]
  113. Lee, S.; Kim, H.; Patel, R.; Im, S.; Kim, J.; Min, B. Silver nanoparticles immobilized on thin film composite polyamide membrane: characterization, nanofiltration, antifouling properties. Polym. Adv. Tech. 2007, 18, 562–568. [Google Scholar] [CrossRef]
  114. Sondi, I.; Sondi, B. silver nanoparticles as antimicrobial agent: A case study on E. Coil as a model for gram-negative bacteria. J. Colloid Interface Sci. 2004, 275, 177–182. [Google Scholar] [CrossRef] [PubMed]
  115. Park, M.; Neigh, A.; Vermeulen, J.; Fonteyne, L.; Verharen, H.; Briede, J.; Loveren, H.; Jong, W. The effect of particle size on the cycotoxicity, inflammation, development of toxicity and genotoxicity of silver nanoparticles. Biomaterials 2011, 36, 9810–9817. [Google Scholar] [CrossRef] [PubMed]
  116. Ben-Sasson, M.; Zodrow, K.R.; Genggeng, Q.; Kang, Y.; Giannelis, E.P.; Elimelech, M. Surface functionalization of thin-film composite membranes with copper nanoparticles for antimicrobial surface properties. Environ. Sci. Technol. 2014, 48, 384–393. [Google Scholar] [CrossRef] [PubMed]
  117. Kim, S.; Kwak, S.; Sohn, B.; Park, T. design of TIO2 nanoparticles self-assembled aromatic polyamide thin-film composite (TFC) membrane as an approach to solve biofouling problem. J. Membr. Sci. 2003, 211, 157–165. [Google Scholar] [CrossRef]
  118. Choi, W.; Choi, J.; Bang, J.; Lee, J.H. Layer-by-layer assembly of graphene oxide nanosheets on polyamide membranes for durable reverse-osmosis applications. ACS Appl. Mater. Interfaces 2013, 5, 12510–12519. [Google Scholar] [CrossRef] [PubMed]
  119. Hu, M.; Mi, B. Enabling graphene oxide nanosheets as water separation membranes. Environ. Sci. Technol. 2013, 47, 3715–3723. [Google Scholar] [CrossRef] [PubMed]
  120. Yin, J.; Yang, Y.; Hu, Z.; Deng, B. Attachment of silver nanoparticles (AgNPs) onto thin-film composite (TFC) membranes through covalent bonding to reduce membrane biofouling. J. Membr. Sci. 2013, 441, 73–82. [Google Scholar] [CrossRef]
  121. Mayyahi, A.A. TiO2 polyamide thin film nanocomposite reverses osmosis membrane for water desalination. Membranes 2018, 8, 66. [Google Scholar] [CrossRef]
Figure 1. Polyamide thin film composite membrane, reproduced with permission from Khorshidi et al. [18].
Figure 1. Polyamide thin film composite membrane, reproduced with permission from Khorshidi et al. [18].
Membranes 08 00068 g001
Figure 2. A schematic illustration of the nano-structured RO membrane, showing the antifouling polymeric brush, reproduced with permission from Lin et al. [51], with copyright permission from ©2010 Royal Society of Chemistry.
Figure 2. A schematic illustration of the nano-structured RO membrane, showing the antifouling polymeric brush, reproduced with permission from Lin et al. [51], with copyright permission from ©2010 Royal Society of Chemistry.
Membranes 08 00068 g002
Figure 3. TFN membrane fabrication by the IP process, reproduced with permission from Yin et al. [19], with copyright permission from © 2014 Elsevier.
Figure 3. TFN membrane fabrication by the IP process, reproduced with permission from Yin et al. [19], with copyright permission from © 2014 Elsevier.
Membranes 08 00068 g003
Figure 4. Protection sequence against chlorine attack, reproduced with permission from Kim et al. [85], with copyright permission from © 2013 Elsevier.
Figure 4. Protection sequence against chlorine attack, reproduced with permission from Kim et al. [85], with copyright permission from © 2013 Elsevier.
Membranes 08 00068 g004
Figure 5. Layer-by-layer deposition of positively-charged GO and aminated-GO nanosheets on the membrane surface; reproduced with permission from Choi et al. [118], with copyright permission from © 2013, American Chemical Society.
Figure 5. Layer-by-layer deposition of positively-charged GO and aminated-GO nanosheets on the membrane surface; reproduced with permission from Choi et al. [118], with copyright permission from © 2013, American Chemical Society.
Membranes 08 00068 g005
Figure 6. Schematic illustration of attaching Ag-NPs on the TFC membrane surface; reproduced with permission from Yin et al. [120], with copyright permission from © 2013 Elsevier.
Figure 6. Schematic illustration of attaching Ag-NPs on the TFC membrane surface; reproduced with permission from Yin et al. [120], with copyright permission from © 2013 Elsevier.
Membranes 08 00068 g006
Table 1. Reported monomers for synthesis of polyamide composite membranes.
Table 1. Reported monomers for synthesis of polyamide composite membranes.
AmineChemical StructureAcid ChlorideChemical StructureMembrane PerformanceRef.
MPD Membranes 08 00068 i001TMC Membranes 08 00068 i002It is well-known that the interfacial polymerization of MPD and TFC on a porous support layer results in high water flux and salt rejection[17]
BDSA Membranes 08 00068 i003TMC Membranes 08 00068 i004Water Flux increased by more than 100% by using BDSA in the interfacial polymerization. Simultaneously, salt rejection increased from 89 to 99%.[24]
S-BAPS Membranes 08 00068 i005TMC Membranes 08 00068 i006When compared to the traditional TFC membrane, this membrane showed higher water flux, but lower NaCl rejection and chlorine resistance.[25]
BHDT Membranes 08 00068 i007TMC Membranes 08 00068 i008This membrane demonstrated higher chlorine resistance when compared to the normal TFC membrane.[26]
PAMAM Membranes 08 00068 i009TMC Membranes 08 00068 i010In this study, the effects of PAMAM content on TFC membrane performance were studied. NaCl rejection was increased when PAMAM concentration was increased from 0.1% to 0.5% (w/v), while water flux was reduced.[27]
DETA, TETA, or TEPA Membranes 08 00068 i011TMC Membranes 08 00068 i012Under operating pressure of 36.52 psi, water fluxes of TEPA/TMC, TETA/TMC, and DETA/TMC were 51.1 ± 4.5, 43.5 ± 0.5, and 33.5 ± 2 L/m2·h, respectively. On the other hand, Na2SO4 rejection sequence was: DETA/TMC > TEPA/TMC > TETA/TMC.[28]
DPA Membranes 08 00068 i013TMC Membranes 08 00068 i014The polyester bonds of DPA/TMC produced TFC membrane with high chemical stability, while maintaining good performance.[29]
DABA Membranes 08 00068 i015TMC Membranes 08 00068 i016Results showed that as DABA concentration was increased, the membrane became more hydrophilic and as a result, high water flux (55.4 L/m2·h-250 psi) was achieved.[23]
MPD Membranes 08 00068 i017DMSO Membranes 08 00068 i018The developed membrane showed excellent antimicrobial efficiency and high water flux and salt rejection.[30]
MPD Membranes 08 00068 i019BTAC Membranes 08 00068 i020Membrane surface was highly negatively charged, smooth, and very thin, which in turn produced high fouling resistance.[31]
SMPD Membranes 08 00068 i021TMC Membranes 08 00068 i022When SMPS content was increased, the molecular weight of PA was decreased, and it subsequently increased water flux and decreased NaCl rejection.[32]
MPD Membranes 08 00068 i023mm-PETC Membranes 08 00068 i024Under 290 psi, water flux was 37.1 L/m2·h and NaCl rejection was 98.4% [22]
MPD Membranes 08 00068 i025om-PETC Membranes 08 00068 i026Under 290 psi, water flux was 50 L/m2·h and NaCl rejection was 97.8%
MPD Membranes 08 00068 i027op-PETC Membranes 08 00068 i028Under 290 psi, water flux was 45.2 L/m2·h and NaCl rejection was 97.2%
MPD Membranes 08 00068 i029ICIC Membranes 08 00068 i030Under operating pressure of 232 psi, water flux was 63 L/m2·h and NaCl rejection was 98.2%. In addition, the membrane showed significant resistance against chlorine.[33]
MPD Membranes 08 00068 i031CFIC Membranes 08 00068 i032Under operating pressure of 232 psi, water flux was around 43.3 L/m2·h and NaCl rejection was around 98.6%. In addition, the membrane showed significant resistance against chlorine.
HFA-MDA Membranes 08 00068 i033TMC Membranes 08 00068 i034Under operating pressure of 400 psi, NaCl rejection was 85% at low pH 4, but increased to 96.1% at pH 10. Water flux was 48 L/m2·h and 80 L/m2·h at pH 4 and pH 10, respectively. Besides, the membrane showed significant chlorine resistance.[34]
Bisphenol A Membranes 08 00068 i035TMC Membranes 08 00068 i036This membrane showed significant fouling resistance along with high water flux and salt rejection.[35]
TMBPA Membranes 08 00068 i037TMC Membranes 08 00068 i038Under operating pressure of 130 psi, water flux was 66.7 L/m2·h and the membrane showed good antifouling properties.[36]
Table 2. Summary of important fillers used to modify TFC membranes.
Table 2. Summary of important fillers used to modify TFC membranes.
NanofillerPA Layer MonomersSubstratePerformance of TFNRef.
Zeolite NaAMPD-TMCPSUWater flux was increased from 2.5 × 1012 to 3.9 × 1012 mPa−1·s−1 without compromising salt rejection (94%) by increasing the concentration of nanoparticles from 0 to 0.4 wt.%.[60]
Zeolite NaAAaAMPD:TEA-TMCPSUBoth AgA-TFN and NaA-TFN membranes exhibited higher water flux than that of TFC membrane. No change in salt rejection was observed. Both membranes showed enhanced antimicrobial properties.[61]
Different sized zeoliteMPD:TEA:SLS:IPA-TMCPSUThe membrane embedded with smaller zeolite NPs produced higher water flux than the membrane with larger zeolite NPs.[62]
SilicaMPD-TMCPSUBy increasing silica concentration, the thermal properties of the membrane were considerably enhanced.[68]
MWCNTsMPD-TMCPSUUnder filtration pressure of 225 psi, both water flux and salt rejection were decreased from 18 to 12 L/m2·h and 98 to 92.2 wt.%, respectively, by increasing the concentration of MWCNTs from 0 to 1 wt.%. On the other hand, the membrane demonstrated significant chlorine resistance.[69]
Zeolite -LTAMPD-TMC-post TreatmentPSUNaCl rejection and water flux were 99.4 wt.% and 42 L/m·h, respectively, and had a filtration pressure of 300 psi. [70]
F-SilicaMPD-TMCPSUWhen NPs concentration was 0.4 wt.%, the membrane showed high thermal stability.[71]
F-MWCNTsMPD-TMCPSUThe membrane showed high dyes and brilliant blue rejection (91%)[72]
Metal alkokxideMPD: SLS-TMCPSUWater flux was encreased by approximately 2-fold when compared with the virgin membrane. [73]
Zeolite NaXMPD-TMCPESUnder filtration pressure of 175 psi, the water flux was increased from 8.01 to 29.76 L/m2·h by increasing the content of NPs from 0 to 0.2 wt.% without jeopordizing NaCl rejection (above 90%). Also, the membrane showed good thermal stability.[74]
iLSMMMPD-TMCPSUUnder filtration presure of 300 psi, the optimized water flux was 42 L/m2·h and the NaCl rejection was 97%. Besides, the membrane showed good antifouling properties.[75]
MCM-41MPD-TMCPSUUnder filtration pressure of 300 psi, Water flux was increased from 28 to 46 L/m2·h by increasing the concentration of NPs from 0 to 0.1 wt.%, while NaCl rejection was maintained (97 wt.%). [76]
APQZMPD-TMCPSUWater flux was increased from 16 to 40 L/m2·h by increasing the concentration of NP from 0 to 0.1 wt.%. In addition, the membrane showed good mechanical stability. [77]
Zwitterion-CNTMPD-TMCPESUnder 530 psi, the optimized water flux was 48.46 L/m2·h, and NaCl rejection was 98.6%. [78]
Carboxylic MWNTsMPD-TMCPESUnder 100 psi, the optimized water flux was 40 L/m2·h. Moreover, the membrane showed good mechanical stability. [79]
Zeolite (Silicate-1)MPD-TMCPSUThe membrane showed higher chemical stability than the one with NaX-Zeolite NPs. [80]
Zeolite-NaAMPD-TMCPSUUnder 232 psi, good water flux was achieved (46.5 L/m2·h) by adding the NPs in organic phase and high salt rejection (97%) by adding the NPs in aqueous phase. [81]
Aminated ZeoliteMPD:aPES:TEA-TMCPSUUnder 797 psi, adding PES and TEA to MPD-nanoparticle solution increased water flux from 23.2 to 37.8 L/m2·h without compromising salt rejection (98%). Moreover, the membrane showed good chlorine resistance.[82]
Zeolite-AMPD-TMCPSUThe membrnae showed significant fouling resistance.[83]
Mesoporous SiO2MPD-TMCPSUUnder 232 psi, water flux was increased from 19 to 53 L/m2·h by increasing the concentration of NPs from 0 to 0.1 wt.%, while NaCl rejection remained (97%).[84]
HBP-g-silicaMPD: aPES-TMCPSUUnder 797.7 psi, the optimized water flux was 34.4 L/m2·h, while the salt rejection was 97.7%. And, the membrane showed better chlorine resistance.[85]
Aluminosilicate CNTsMPD-TMCPSUUnder 232 psi, the optimized water flux was 23 L/m2·h, while NaCl rejection was 97.5%.[86]
F-MWCNTsMPD-TMCPSUUnder 232 psi, the optimized water flux was 28.05 L/m2·h, while salt rejection was 90%. In addition, the membrane showed better antifouling and antioxidant properties.[87]
HNTsMPD-TMCPSUUnder 217.5 psi, water flux was increased from 18 to 36.1 L/m2·h by increasing the concentration of NPs from 0 to 0.1% without sacrificing NaCl rejection (93%). Besides, the membrane had enhanced fouling properties.[88]
OA-SiO2MPD-TMCPSUThe OA modified-silica PA membrane produced higher salt rejection (98%) when compared to the unmodified silica PA membrane (95%).[89]
ClayMPD-TMCPSUUnder 232 psi, water flux was increased from 36.6 to 51 L/m2·h by adding 0.1 wt.% NPs without compromising NaCl rejection (around 99%). Also, the membrane exhibited significant antifouling properties. [90]
GO-TiO2MPD-TMCPSUUnder 217.5 psi, both water flux and salt rejection were increased from 34 to 51 L/m2·h and 97 to 99%, respectively, by adding 0.02 wt.% NPs. Besides, the membrane demonstrated robust chlorine resistance. [91]
HN2-TNTsMPD-TMCPSUUnder 217.5 psi, both water flux and NaCl rejection were increased from 19 to 36 L/m2·h and 94 to 96%, respectively, by adding 0.05 wt.% NPs. Moreover, the membrane showed good fouling resistance.[92]
GOMPD-TMCPSUUnder 217 psi, the optimized water flux was 22 L/m2·h, while NaCl rejection was above 80%. Moreover, the modified membrane exhibited excellent fouling resistance against BSA and HA.[93]
Al-ZnOMPD-TMCPSUUnder 225 psi, the optimized water flux was 32 L/m2·h, while NaCl rejection was 98%.[94]
MCM-48-SiO2MPD-TMCPSUUnder 232 psi, the optimized water flux was 68 L/m2·h. And, NaCl rejection was around 97%.[95]
GOMPD-TMCPSUUnder 300 psi, water flux was increased from 39 to 60 L/m2·h by increasing NPs concentrations from 0 to 0.015 wt.%, while NaCl rejection was above 93%. [96]
ZnOMPD-TMCPSUUnder 300 psi, water flux was increased from 60 to 85 L/m2·h by increasing the concentration of ZnO from 0 to 0.1 wt.%. Under UV irradiation the membrane showed super water flux (120 L/m2·h). In addition, the membrane showed excellent fouling resistance.[97]
MOFsMPD-TMCPSUUnder operation pressure of 300 psi, water flux and NaCl rejection were 85 L/m2·h and 98.5%, respectively.[98]
Graphene quantum dotsPIP-TMCPESUnder operation pressure of 0.2 Mpa, water flux was 120 L/m2·h, 6.8-times higher than that of the virgin membrane. Moreover, the membrane showed excellent fouling resistance.[99]
ZIF-8MPD-TMCPSU53% enhancement in water flux was achieved. NaCl rejection was 99.4%.[100]
TiO2MPD-TMCPESThe addition of TiO2 resulted in higher water flux (24.3 L/m2·h) as compasred with the virgin TFC (21.5 L/m2·h), while membrane selectivity was preserved (97%). Additionally, by increasing feed solution temeprature from 25 to 65 °C, further enhancement in water flux was achieved.[101]
CQDsPIP-TMCPSUThe addition of carbon quantum dots led to significant incerease in permeate flux (from 18 to 42.1 L/m2·h) without jeopordizing Na2SO4 rejection (93%). Moreover, the fouling capacity of membrane was enhanced.[102]
Na+ functionalized CQDsMPD-TMCPESImpresive water flux (104 L/m2·h), high rejection of SeO32 (97.5%), and excellent fouling resistance were achieved when quantum dots concentration was 0.05 wt.%.[103]
SiO2MPD-TMCPSUWater flux was increased from 30 to 50 L/m2·h by increassing NPs concentration from 0 to 0.1 wt.% along with slight increase in salt rejection (from 92 to 95%).[104]
Ziconiumv (IV)-carboxylate MOFsMPD-TMCPSU-PVP-LiCl52% increase in water flux was achieved without comprimising NaCl rejection (95.5%).[105]

Share and Cite

MDPI and ACS Style

Al Mayyahi, A. Important Approaches to Enhance Reverse Osmosis (RO) Thin Film Composite (TFC) Membranes Performance. Membranes 2018, 8, 68. https://doi.org/10.3390/membranes8030068

AMA Style

Al Mayyahi A. Important Approaches to Enhance Reverse Osmosis (RO) Thin Film Composite (TFC) Membranes Performance. Membranes. 2018; 8(3):68. https://doi.org/10.3390/membranes8030068

Chicago/Turabian Style

Al Mayyahi, Ahmed. 2018. "Important Approaches to Enhance Reverse Osmosis (RO) Thin Film Composite (TFC) Membranes Performance" Membranes 8, no. 3: 68. https://doi.org/10.3390/membranes8030068

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop