You are on page 1of 58

Accepted Manuscript

A Review on Polyamide Thin Film Nanocomposite (TFN) Membranes: History,


Applications, Challenges and Approaches

W.J. Lau, Stephen Gray, T. Matsuura, D. Emadzadeh, J.Paul Chen, A.F. Ismail

PII: S0043-1354(15)00274-2
DOI: 10.1016/j.watres.2015.04.037
Reference: WR 11266

To appear in: Water Research

Received Date: 8 January 2015


Revised Date: 23 April 2015
Accepted Date: 24 April 2015

Please cite this article as: Lau, W.J., Gray, S., Matsuura, T., Emadzadeh, D., Chen, J.P., Ismail, A.F., A
Review on Polyamide Thin Film Nanocomposite (TFN) Membranes: History, Applications, Challenges
and Approaches, Water Research (2015), doi: 10.1016/j.watres.2015.04.037.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Graphical Abstract

PT
RI
U SC
AN
M
D
TE
C EP
AC
1
ACCEPTED MANUSCRIPT
1 A Review on Polyamide Thin Film Nanocomposite (TFN) Membranes: History, Applications,

2 Challenges and Approaches

4 W.J. Laua*, Stephen Grayb, T. Matsuuraa,c, D. Emadzadeha,d, J. Paul Chene, A.F. Ismaila
a
5 Advanced Membrane Technology Research Centre (AMTEC), Universiti Teknologi Malaysia, 81310

PT
6 Skudai, Johor, Malaysia
b
7 Institute for Sustainability and Innovation (ISI), College of Engineering and Science, Victoria

RI
8 University, Werribee Campus, PO Box 14428, Melbourne, VIC 8001, Australia

SC
c
9 Industrial Membrane Research Laboratory, Department of Chemical and Biological Engineering,

10 University of Ottawa, 161 Louis Pasteur St, Ottawa, ON K1N 6N5, Canada

U
d
11 Department of Chemical Engineering, Gachsaran Branch, Islamic Azad University, Gachsaran, Iran
AN
e
12 Department of Civil and Environmental Engineering, National University of Singapore, 10 Kent

13 Ridget, 129791, Singapore


M

14 *Corresponding author. Tel.:+60 75535926


D

15 E-mail addresses: lwoeijye@utm.my, lau_woeijye@yahoo.com; lau.woeijye09@gmail.com (W.J. Lau)


TE

16

17 Abstract
EP

18 This review focuses on the development of polyamide (PA) thin film nanocomposite (TFN)

19 membranes for various aqueous media-based separation processes such as nanofiltration, reverse
C

20 osmosis and forward osmosis since the concept of TFN was introduced in year 2007. Although the
AC

21 total number of published TFN articles falls far short of the articles of the well-known thin film

22 composite (TFC) membranes, its growth rate is significant, particularly since 2012. Generally, by

23 incorporating an appropriate amount of nanofiller into a thin selective PA layer of a composite

24 membrane, one could produce TFN membranes with enhanced separation characteristics as

25 compared to the conventional TFC membrane. For certain cases, the resulting TFN membranes

26 demonstrate not only excellent antifouling resistance and/or greater antibacterial effect, but also
2
ACCEPTED MANUSCRIPT
27 possibly overcome the trade-off effect between water permeability and solute selectivity.

28 Furthermore, this review attempts to give the readers insights into the difficulties of incorporating

29 inorganic nanomaterials into the organic PA layer whose thickness usually falls in a range of several-

30 hundred nanometers. It is also intended to show new possible approaches to overcome these

31 challenges in TFN membrane fabrication.

PT
32

33 Keywords: thin film nanocomposite; membrane; nanomaterials; polyamide; thin film composite

RI
34

SC
35

36

U
37
AN
38

39
M

40
D

41
TE

42

43
EP

44

45
C

46
AC

47

48

49

50

51

52
3
ACCEPTED MANUSCRIPT
53 1. Introduction

54 Thin film nanocomposite (TFN) membrane is a new type of composite membranes prepared via

55 interfacial polymerization (IP) process. Nanoparticles are incorporated within the thin polyamide

56 (PA) dense layer of the thin film composite (TFC) membrane with the aim of improving the

57 characteristics of the interfacially polymerized layer. For example, to increase the hydrophilicity

PT
58 and/or surface charge density, without sacrificing the separation efficiency of the TFC membrane.

59 The term TFN membrane was used by Hoek and his co-workers (Jeong et al., 2007) in their

RI
60 pioneering research work published in Journal of Membrane Science in early 2007. In this work, they

SC
61 reported a new concept for the fabrication of reverse osmosis membrane via the IP process by

62 embedding zeolite NaA nanoparticles (in the range of 0.0040.4% (w/v)) within the PA layer (made

U
63 of m-phenylediamine (MPD) and trimesoyl chloride (TMC)) of the composite membrane. A
AN
64 remarkable membrane flux enhancement was achieved upon zeolite nanoparticle incorporation with

65 solute rejection remaining comparable to the typically prepared TFC membrane. It was suggested
M

66 that water molecules tend to flow preferentially through the super-hydrophilic and molecular sieve
D

67 nanoparticle pores. A similar membrane type was actually presented by the same research group in
TE

68 2005 during an AIChE Annual Meeting and Fall Showcase in the United States, but the membrane

69 incorporated with nanoparticles modified with a biocidal agent was referred to as Tfnc at that time,
EP

70 rather than TFN (Jeong et al., 2005).

71 Industrially, NanoH2O Inc successfully conducted a field test of a TFN element at Port
C

72 Hueneme United States Navy Facility in September 2008 and reported that the water flux of the TFN
AC

73 membrane was twice the flux of the conventional PA membrane with NaCl rejection maintained at

74 >99.7% (Kurth et al., 2011; LG NanoH2O Inc., 2015). Two years later in September 2010, the

75 company established its first commercial pilot seawater desalination installation using TFN

76 membrane modules under the QuantumFlux brand followed by official launching into the RO

77 membrane desalination market in April 2011 (LG NanoH2O Inc., 2015).

78
4
ACCEPTED MANUSCRIPT
79 Research interest in TFN membranes has subsequently increased since the first reports from

80 Hoeks research group (Jeong et al., 2007; Jeong et al., 2005). As of first quarter of 2015 when this

81 review article is prepared, we have seen that more than 90 relevant articles reporting on the use of

82 nanomaterials in preparing PA TFN membranes for various separation processes. As shown in Table

83 1, the potential of the TFN membrane is not only limited to aqueous-based media like nanofiltration

PT
84 (NF), reverse osmosis (RO) and forward osmosis (FO) separation processes, but it is also used for the

85 treatment of organic solvents as well as pervaporation processes. Although the history of the TFN

RI
86 membrane is still relatively short in comparison to the TFC membrane, which has been known since

SC
87 1960s, the growing interest on TFN membrane development among scientists is obvious over the

88 past 12 years. This is reflected by the increasing number of articles/technical papers published in

U
89 the open literature in 2013-2014, as shown in Figure 1.
AN
90 The present review focuses on the works related to the PA TFN membrane fabrication and

91 its applications since 2007. The main highlight, hence the most significant contribution of this review
M

92 article, is the focus and emphasis on methods to incorporate different types of nanomaterials into
D

93 PA for achieving better results than that of conventionally prepared TFC membrane. The key
TE

94 technical challenges that one might encounter in the fabrication process of the TFN membrane and

95 the possible approaches that could be employed to overcome these limitations are also highlighted.
EP

96

97 *****Figure 1. Number of relevant research documents published in refereed journals between


C

98 2007 and 2015 (Source: Scopus, Date of access: 31 March 2015)*****


AC
ACCEPTED MANUSCRIPT

Table 1. The potential of using PA TFN membranes in different separation processes


a b c d e
Membrane Nanofillers Active Substrate Optimized TFN Separation performance of TFN Ref.
type embedded in/on monomers surface

PT
PA layer (Ave. for PA characteristic
particle size) layer
Aminosilanized MPD-TMC PES CA: 75.8o The trade-off effect between permeability and rejection (Rajaeian

RI
TiO2 NPs Ra: 79.2 nm was only overcome when the PA layer of membrane was et al.,
(n/a) incorporated with small loading of NPs (i.e. 0.005 wt%). 2013)

SC
By incorporating higher loadings of NPs (0.05 and 0.1
NF wt%) into PA layer, NaCl rejection was reported to
decrease and flux increase.

U
Modified SiO2 NPs PIP-TMC PSF CA: n/a Both pure water flux and Na2SO4 rejection of membrane (Wu et al.,

AN
(100 nm) Ra: n/a was influenced upon addition of modified SiO2 (ranging 2013)
between 0 and 0.07 wt%) into PA layer of membrane.
Optimum flux was achieved at 0.03 wt% NPs but salt

M
rejection was reported to decrease gradually with an
increase in NPs loading.
SRNF MOF NPs MIL- MPD-TMC Cross- CA: 50o The incorporation of only 0.2% (w/v) MIL-101 (Cr) into (Sorribas et

D
101 (Cr) linked PI Ra: n/a the PA layer of TFN membrane was able to result in an al., 2013)

TE
(47 nm) exceptional increase in solvent flux, from 1.5 to 3.9 and
from 1.7 to 11.1 L m-2 h-1 bar-1 for styrene oligomers that
contained methanol and tetrahydrofuran, respectively.
EP
NaA zeolite MPD-TMC PSF CA: ~38o Pure water flux of membrane was increased from (Jeong et
(100 nm) Ra: ~57 nm 2.110-12 mPa-1s-1 (as in TFC) to 3.810-12 mPa-1s-1 while al., 2007)
rejection for NaCl remained at 94% with increasing
C

zeolite loading from zero to 0.4% (w/v). The expected


AC

permeability-selectivity trade-off was overcome.


Ag NPs MPD-TMC PSF CA: 32.9o TFN membrane embedded with Ag NPs demonstrated (Yin et al.,
RO (15.3 nm) Ra: n/a water permeability of 69.4 L m-2 h-1 and NaCl rejection of 2013)
93.6% when tested at ~20.7 bar (300 psi). Compared to
the control TFC membrane (49.8 L m-2 h-1 and 95.9%), it
was found that the TFN membrane flux was able to
ACCEPTED MANUSCRIPT

improve with only a slight drop in rejection upon


addition of Ag NPs.
Aluminosilicate MPD-TMC PSF CA: 41.6o When tested at 16 bar, pure water flux of membrane (Baroa et
SWNTs Ra: 50.8 nm was reported to increase gradually from 10.5 L m-2 h-1 al., 2013)

PT
(OD ID L : 2.7 (typical TFC) to >24 L m-2 h-1 (TFN) by gradually increasing
nm n/a 150 nm) SWNTs loadings from zero to 0.2 % (w/v). Rejections
against NaCl and Na2SO4 remained in the range 95.5-

RI
96.0% and 97-97.5%, respectively, regardless of the
SWNTs loading.

SC
FO Functionalized MPD-TMC PSF CA: n/a Performance of TFN FO membrane with respect to water (Amini et
MWCNTs Ra: 97.2 nm flux and reverse solute flux improved compared to al., 2013)
(OD ID L : 5 nm typical TFC membrane. The most permeable TFN

U
1.3-2 nm 50 membrane (embedded with 0.1 wt% NPs) had a water

AN
m) flux of 95.7 L m-2 h-1 that was nearly 160% higher than
TFC when tested under same conditions. However, this
membrane showed poor NaCl rejection (73%) in RO

M
experiments.
Pervaporation NaX zeolite MPD-TMC Modified CA: 32o Water selectivity of membrane improved significantly (Fathizadeh
(105 nm) PAN Ra: 29.8 nm from around 180 to >300 with an increase of zeolite et al.,

D
loading from zero to 0.15% (w/v) when tested with 90 2013)

TE
wt% aqueous isobutanol. The improvement might be
due to superior hydrophilicity of zeolite coupled with its
molecular sieving effect.
EP
a
NF: Nanofiltration; SRNF: Solvent resistant nanofiltration; RO: Reverse osmosis and FO: Forward osmosis
b
TiO2: Titanium dioxide, NPs: Nanoparticles; SiO2: Silica nanoparticles; MOF: Metal organic framework; Ag: Silver; SWNTs: Single-walled nanotubes and MWCNTs:
Multi-walled carbon nanotubes; OD: Outer diameter; ID: Inner diameter; L: Length
C

c
MPD: m-phenylediamine; TMC: Trimesoyl chloride and PIP: Piperazine
d
PES: Polyethersulfone; PSF: Polysulfone; PI: Polyimide and PAN: Polyacrylonitrile
AC

e
CA: Contact angle and Ra: Average roughness value
7
ACCEPTED MANUSCRIPT
99 2. Progress of TFN membrane

100 In this section, a comprehensive review on the properties and separation performances of TFN

101 membranes made of various inorganic nanomaterials for different types of applications will be

102 provided. Focus is placed on how the presence of nanomaterials with different surface properties

103 and different loadings could alter the characteristics of PA layer and/or supporting substrate of the

PT
104 TFN membrane and improve membrane separation performance. Comments are also provided on

105 some experimental results that might be less convincing from a technical point of view.

RI
106

SC
107 2.1 Reverse osmosis

108 The earliest mention of the TFN membrane which was disclosed by Hoek and his co-workers (Jeong

U
109 et al., 2007) in 2007 was for RO applications. In this pioneering work, the authors introduced a new
AN
110 concept in preparing composite RO membrane by adding zeolite nanoparticles (NaA) into the matrix

111 of thin PA selective layer. The results showed that compared with the control TFC membrane, the
M

112 newly developed TFN membranes could exhibit improved water permeability without compromising
D

113 NaCl rejection when both types of membranes were tested under the same conditions. Later, Lind et
TE

114 al. (2009) and Fathizadeh et al. (2011) used different types of zeolite nanoparticles for the TFN

115 membranes.
EP

116 Table 2 compares the performances of the TFN membranes made by these three research

117 groups with respect to water flux and NaCl rejection. As can be seen, the separate research groups
C

118 showed that the incorporation of a small quantity of zeolite nanoparticles, in the range of 0.040.4
AC

119 w/v%, into PA layer could promote water permeability with minimal/no changes in salt rejection.

120 However, the degree of flux enhancement varied depending on the zeolite type and loading as well

121 as the conditions of the IP process. For instance, it was reported by Lind et al. (2009) that smaller

122 zeolites (~100 nm) tended to produce greater water flux due to larger characteristic pores while

123 larger zeolite particles (~300 nm) produced more favourable surface properties by increasing the

124 negative charge and lowering the contact angle, owing to its exposure at the PA surface. Exposure of
8
ACCEPTED MANUSCRIPT
125 nanofillers at the PA surface is highly possible to occur especially for the nanoparticles with large

126 particle size. This is because of the irregular morphology of PA layer that is very often reported to be

127 100500 nm in thickness (Lau et al., 2015). Fathizadeh et al. (2011) on the other hand claimed that

128 for the TFN membrane with same zeolite loading (0.2%), PA layer made from a lower concentration

129 of MPD (2 w/v%) and TMC (0.1 w/v%) could produce a better combination of flux and salt rejection

PT
130 for seawater desalination than the PA layer made of higher monomer concentration, i.e. 3 and

131 0.15%, respectively. With respect to the dispersion quality of nanoparticles in organic solvent (during

RI
132 IP process), none of these research works studied in detail the possible PA surface defects due to the

SC
133 agglomeration/non-homogenous distribution of nanoparticles in the PA layer. Although several

134 recent findings indicated that the use of high amounts of nanoparticles (> 1 wt%) during PA layer

U
135 synthesis could lead to significant particle agglomeration in selective layer which causes surface
AN
136 defects and reduced salt removal efficiency (Huang et al., 2013a; Xu et al., 2013; Emadzadeh et al.,

137 2014; Ghanbari et al., 2015a), one cannot rule out other possibilities that might lead to poor
M

138 separation performance. These include PA matrix not binding with the nanoparticles firmly (which
D

139 increases system free volume) and/or reduced PA cross-linking degree when nanoparticles
TE

140 concentrations exceed an upper limit.

141 The improvement in the water flux of the TFN membrane incorporated with zeolite was
EP

142 attributed to the unique crystal structure of zeolite molecular sieve particles, which could provide a

143 preferential flow path for water molecules through its internal pore structure (Lind et al., 2009).
C

144 Besides offering pores that can accommodate water molecules due to their relatively large sizes, it
AC

145 has also been claimed that the hydrophilicity and negative charge of zeolite nanoparticles could

146 result in greater affinity for water molecules and increases repulsion of anions due to Coulombic

147 effects (Barrer and James, 1960; Jeong et al., 2007). Peeters et al. (1998) also reported that a

148 membrane with higher negative surface charge tends to have greater retention for bivalent anions

149 (e.g. SO42-) than that of monovalent anions (e.g. Cl-) owing to the Donnan exclusion that takes place

150 during separation process, in addition to the mechanisms of sieving and diffusion-solubility.
9
ACCEPTED MANUSCRIPT
151 Nevertheless, the use of charged zeolite nanoparticles is associated with a drawback, i.e.

152 poor rejection of NaCl under certain conditions. As reported by Lind et al. (2009), increasing MPD

153 and TMC concentrations from 2.3 to 3.2% (w/v) and from 0.1 to 0.13% (w/v), respectively during TFN

154 membrane preparation, resulted in a membrane with very high flux but poor NaCl rejection. Lind et

155 al. (2009) suggested that this might be caused by the interaction between the zeolites and the active

PT
156 monomers which altered the pore structure of PA film, or nanofillers did not form strong bond with

157 the polymer matrix resulting in the formation of defects. Similarly, Fathizadeh et al. (2011) found

RI
158 that the increases in both MPD and TMC concentration from 2 to 3 % w/v and from 0.1 to 0.15 %

SC
159 w/v, respectively, for TFN membrane preparation produced membranes with greater flux but with

160 much lower NaCl rejection. They attributed the results to the poor interaction between zeolite and

U
161 PA layer of higher molecular weight polymer which leads to the formation of PA surface defects,
AN
162 although the defects were not verified by any characterisation measurements. The findings of Lind

163 et al (2009) and Fathizadeh et al. (2011) are important because they demonstrate that changes to
M

164 the IP process are required once nanoparticles are added to the formulation if good performing
D

165 membranes are to be obtained. Addition of nanoparticles directly to an existing IP process without
TE

166 altering those conditions did not produce membranes of improved performance. Restricting the

167 MPD and TMC concentration to around 2 and 0.1%, respectively was required to produce a PA TFN
EP

168 membrane with a good combination of water flux and NaCl rejection. Similar monomer

169 concentrations have also been adopted by other researchers for TFN RO membrane preparation
C

170 (Jadav and Singh, 2009; Kong et al., 2010; Huang et al., 2013b; Pendergast et al., 2013; Yin et al.,
AC

171 2012).

172 Besides zeolite nanoparticles, other types of nanomaterials such as aluminosilicate single-

173 walled nanotubes (SWNTs) (Baroa et al., 2013), silica (SiO2) (Jadav and Singh, 2009; Yin et al., 2012),

174 titanium nanotubes (TNTs) (Emadzadeh et al., 2014d), halloysite nanotubes (HNTs) (Ghanbari et al.,

175 2015a), silver (Ag) (Kim et al., 2012; Ben-Sasson et al., 2014) and metal alkoxides (Kong et al., 2011)

176 have been reported as components of TFN RO membranes. These studies all showed that the
10
ACCEPTED MANUSCRIPT
177 resultant TFN membranes exhibited better performance, in particular much higher water flux,

178 provided an appropriate amount of nanomaterials was used in the preparation process. For

179 instance, Baroa et al. (2013) reported that the water flux of the TFN membrane incorporated with

180 0.59 wt% aluminosilicate SWNTs increased to as high as 24.6 L m-2 h-1 at 16 bar, i.e. a 1.5-fold

181 increase in comparison to the typical TFN membrane with little variation in NaCl rejection (see Table

PT
182 2). This significant surge in water flux can be attributed to the increased preferential flow of water

183 molecules through the hydrophilic inner walls (with diameter approximately 1 nm) of the

RI
184 aluminosilicate nanotubes embedded within the PA matrix. Remarkable water enhancement in the

SC
185 TFN membrane was also reported by Yin et al. (2012) upon addition of 0.1 wt% porous MCM-41 SiO2

186 nanoparticles into the PA selective layer without compromising NaCl separation efficiency (see Table

U
187 2). The increase in water flux from 28.5 L m-2 h-1 of TFC membrane to > 46 L m-2 h-1 of TFN membrane
AN
188 at an operating pressure of ~20.7 bar (equivalent to 300 psi) is attributed to the internal pores of

189 MCM-41 (about 3.85 nm) which create shorter flow paths for water molecules to pass through. In
M

190 addition to the preferential flow mechanism, Baroa et al. (2013) did not rule out that the reduction
D

191 in membrane surface contact angle upon addition of nanotubes might partly contribute to better
TE

192 water permeability. Compared to solid nanomaterials such as TiO2 and metal alkoxides, the

193 existence of pore channels in mesoporous nanomaterials, coupled with unique surface
EP

194 characteristics could potentially achieve synergistic effect for rapid diffusion of water molecules,

195 resulting in much greater water permeability. However, the preferential flow channels created by
C

196 mesoporous nanomaterials embedded in TFN membrane for water molecules deserves more
AC

197 research to validate this proposition, as CNTs are known to require a functionalised entrance

198 (Bakajin et al.,2009) and MFI-type zeolite membranes require high pressures to achieve even low

199 flux (Zhu et al., 2015).

200 Other progress has also been made on the synthesis of advanced TFN RO membrane that

201 incorporates Ag nanoparticles. Ag nanoparticles can act as antibacterial agents to deactivate

202 microorganisms during the filtration process, mitigating membrane biofouling. Research has
11
ACCEPTED MANUSCRIPT
203 revealed that TFN membranes containing Ag nanoparticles in the thin layer improved water

204 permeability and also demonstrated antibacterial effects on the growth of P. aeruginosa PAO1 (Kim

205 et al., 2012). The increased permeability arising from the incorporation of Ag nanoparticles is

206 interesting, as the Ag nanoparticles do not contain any potential water channels. This suggests that

207 another water transport mechanisms may be active in achieving higher membrane permeability

PT
208 apart from the possibility of highly conductive water channels within the added nanoparticles. With

209 respect to salt separation performance, the TFN membrane achieved very similar rejection rate

RI
210 against NaCl and Na2SO4 in comparison to the TFC membrane.

SC
211 The incorporation of antibacterial effects within membranes by adding Ag nanoparticles has

212 also been a significant topic for research (Yin and Deng, 2015). Unlike Lee et al. (2007) and Kim et al.

U
213 (2012) who fabricated TFN RO membrane by introducing Ag nanoparticles into the organic phase
AN
214 during interfacial polymerization, Yin et al. (2013) grafted the TFC PA surface with Ag nanoparticles

215 in order to reduce the potential risks of releasing Ag nanoparticles and Ag+ ions to the environment.
M

216 This approach chemically immobilizes Ag nanoparticles on the TFC membrane surface by reacting
D

217 the PA layer with NH2-(CH2)2-SH ethanol solution (20 mM) followed by incubating with an Ag
TE

218 nanoparticles suspension (0.1 mM) for 12 h, as illustrated in Figure 2. Besides showing enhanced

219 water flux with slightly lower NaCl rejection (compared with TFC membrane), the resultant TFN
EP

220 membrane also demonstrated excellent antibacterial ability to inhibit E. coli growth. The authors

221 claimed that this grafting approach was very effective at reducing Ag nanoparticle leaching during
C

222 filtration due to the strong covalent bonding formed between thiol groups and Ag nanoparticles as
AC

223 confirmed by Raman spectra. With respect to the Ag particle size, it is found that the TFN membrane

224 prepared via covalent bonding showed much smaller particles size (1020 nm) on membrane surface

225 (Yin et al., 2013) than that of physically embedding Ag nanoparticles in membranes (50100 nm)

226 (Lee et al., 2007) and has about the same size with that of Ag nanoparticles (average 15.3 nm).

227 Furthermore, the grafting approach offers distinct advantages over physically embedding technique
12
ACCEPTED MANUSCRIPT
228 as the original PA structure is maintained with only a very minor change in thickness and is likely to

229 reduce Ag leaching during preparation.

230 The use of nanoparticles to improve TFN membrane resistances against fouling and chlorine

231 attack is also an interesting work in the development of practical and sustainable membrane

232 materials for industrial applications (Misdan et al., 2012). In the work conducted by Zhao et al.

PT
233 (2014), it was demonstrated that the introduction of carboxyl-functionalized multi-walled carbon

234 nanotubes (MWCNTs) into PA structure could improve both membrane antifouling property and

RI
235 chlorine resistance. In comparison to the control TFC membrane, the TFN membrane embedded

SC
236 with 0.1 wt% functionalized MWCNTs always showed smaller degree of flux decline in different

237 filtration cycles using feed solution containing bovine serum albumin (BSA). The authors attributed

U
238 the better antifouling property of TFN membrane to the improved surface hydrophilicity as well as
AN
239 greater negative surface charge upon addition of MWCNTs. Furthermore, this novel TFN membrane

240 has been reported to exhibit better chlorine resistance when evaluated either in dynamic or
M

241 immersion mode, owing to the protection of amide linkage by electron-rich MWCNTs. Similar results
D

242 were also reported by Park et al. (2010) using TFN membrane incorporated with MWCNTs. The
TE

243 authors explained that the better interaction between the carboxylic group of MWCNTs and the

244 amide bond in the PA layer was the main reason for better stability of TFN membrane against
EP

245 chlorine. Chlorine is generally known as the main component attacking PA structure, forming N-

246 chlorinated amide in the initial step. The process is followed by a non-reversible reaction i.e. ring-
C

247 chlorination through intramolecular rearrangement of chlorine atom into aromatic ring of the
AC

248 diamine moiety via Orton rearrangement. Effort was also made by Kim et al. (2013) to synthesize

249 TFN membrane with enhanced chlorine resistance using hyper-branched aromatic polyamide-

250 grafted silica (HBP-g-silica) nanoparticles. Besides improving composite membrane water flux, the

251 modified silica nanoparticles were also reported to protect the active layer of PA structure from

252 chlorine attack. This enhanced chlorine resistance may be due to the following factors: (1) improved

253 intermolecular hydrogen bonding between the modified silica and PA layer structure and (2) the
13
ACCEPTED MANUSCRIPT
254 additional amino groups introduced by aminated nanoparticles. The synergistis effects of particle

255 adsorption and membrane rejection using TFN membrane incorporated with nano-adsorbents is also

256 possible to obtain, considering the successful studies on the use of nano-adsorbents in the

257 fabrication of mixed matrix ultrafiltration membranes for arsenic removal (He et al., 2014).

258 It should also be highlighted that in addition to the inorganic nanoparticles, there are several

PT
259 research works reporting the potential of using biomimetic composite membranes made of

260 aquaporin (water channel protein) for water treatment process (Zhao et al., 2012; Tang et al., 2013;

RI
261 Li et al., 2014). Conceptually, the approach of incorporating PA layer with biomolecules is similar to

SC
262 the TFN membrane fabrication as discussed earlier. The only difference is TFN membrane is

263 synthesized by introducing inorganic nanofillers into PA layer while biomimetic TFC membrane is

U
264 made of biological entities. With respect to separation performance, it was reported that the
AN
265 aquaporin-based TFC biomimetic membrane could potentially achieve 40% higher water

266 permeability than those of commercial RO membranes tested (i.e. BW30 and SW30HR from Dow
M

267 FilmTecTM) with even slightly better in NaCl rejection (Zhao et al., 2012). Despite these promising
D

268 results, this technology is not developed to the point of commercialization (Aquaporin A/S, 2015)
TE

269 and further studies are required at larger scale to determine its sustainable operation (Subramani

270 and Jacangelo, 2015).


EP

271

272 *****Figure 2. Schematic diagram of immobilization of silver (Ag) nanoparticles onto the top PA
C

273 surface of TFC membrane (Yin et al., 2013).*****


AC
ACCEPTED MANUSCRIPT

14

Table 2. Comparison between the performance of TFC and TFN membranes prepared from different types of nanofillers

PT
Zeolite Aluminosilicate SWNTs MCM-41 SiO2 NPs
a b c d e
Work of Jeong et al. Work of Lind et al. Work of Fathizadeh et al. Work of Barona et al. Work of Yin et al.

RI
Membrane (2007) (2009) (2011) (2013) (2012)
NaA LTA NaX
loading Flux RNaCl loading Flux RNaCl loading Flux RNaCl Loading Flux RNaCl Loading Flux RNaCl

SC
(w/v%) (m Pa-1 s-1) (%) (w/v%) ( s-1) (%) (w/v%) (L m-2 h-1) (%) (w/v%) (L m-2 h-1) (%) (w/v%) (L m-2 h-1) (%)
TFC 2.110-12 ~23.5 ~88 8.32 10.5 ~95.6 28.5 98.1

U
TFN (Type 1) 0.004 ~2.110-12 0.004 ~8.55

AN
TFN (Type 2) 0.01 ~2.510-12 ~93.0 0.01 ~10.9 ~94 0.05 ~14 ~96.2 0.01 ~39 ~97.8
TFN (Type 3) 0.1 ~3.1510-12 0.2 ~25.9 ~90 0.04 ~11.3 0.1 ~19 ~96.1 0.1 46.6 ~97.7
3.810-12 0.2 24.6 ~96.2

M
TFN (Type 4) 0.4 0.2 14.6
a
Solution properties during IP process= Aqueous (w/v): 2% MPD, Hexane (w/v): 0.1% TMC and NaA
Testing conditions = 12.4 bar and 2000 ppm NaCl

D
b
Solution properties during IP process= Aqueous (w/v): 2.5% MPD, 6.6% TEACSA, 0.02% SLS and 20% IPA, Isopar (w/v): 0.1% TMC and 0.2% LTA (200 nm)
Testing conditions = 15.5 bar and 2000 ppm NaCl

TE
c
Solution properties during IP process= Aqueous (w/v): 2% MPD, Hexane (w/v): 0.1% TMC and NaX
Testing conditions = 12 bar and 2000 ppm NaCl
d
Solution properties during IP process= Aqueous (w/v): 2% MPD, Hexane (w/v): 0.01% TMC and aluminosilicate SWNTs
EP
Testing conditions = 16 bar and 2000 ppm NaCl
e
Solution properties during IP process= Aqueous (w/v): 2% MPD, Hexane (w/v): 0.15% TMC and MCM-41
Testing conditions = ~20.7 bar and 2000 ppm NaCl
C
AC
15
ACCEPTED MANUSCRIPT
274 2.2 Nanofiltration

275 The distinct differences between the performances of the TFN RO and the TFN NF membranes are

276 the greater water permeability of the NF membrane and its high rejection for multivalent salt (e.g.

277 Na2SO4) compared to lower rejection to monovalent salt (e.g. NaCl). Compared to the PA layer of RO

278 membranes which is mostly prepared via IP between MPD and TMC (see Table 2), the PA film of NF

PT
279 membrane is usually synthesized using amine monomer with lower reactivity such as piperazine

280 (PIP) (Hu et al., 2012; Shen et al., 2013; Wu et al., 2013; Ong et al., 2012), causing the formation of

RI
281 more linear-like polymer structures with lower degrees of cross-linking. Nevertheless, the use of PIP

SC
282 is not a requirement for the fabrication of NF (or TFN NF) membrane, as the characteristics of PA

283 layer of NF membranes are also dependent on many other factors such as the concentration of

U
284 reactant, reaction time, type of organic solvent, additive and substrate (Kim and Deng, 2011; Lau and
AN
285 Ismail, 2009; Lau et al., 2014; Lau et al., 2012; Misdan et al., 2015, 2014, 2012; Rajaeian et al., 2013).

286 For instance, increasing TMC concentration from 0.05 to 0.30 w/v% and reaction time from 10 to 60
M

287 s during IP tended to improve separation performance of NF against Na2SO4 from ~93% to >97.5%
D

288 and from 96.5% to >97.5%, respectively, owing to the thicker and denser PA skin layer produced (Hu
TE

289 et al., 2012). Meanwhile, the addition of propanol to the aqueous amine phase could act to increase

290 aqueous/organic media miscibility during IP which increases the breadth of the reaction zone and
EP

291 produces a less dense and more permeable PA film (Lind et al., 2009). Also, to produce a more

292 permeable PA film, it is reported that a substrate of bigger pores tended to allow more PIP aqueous
C

293 solution to diffuse deep into the pore channel, leading to the possibility of PA cross-liking inside the
AC

294 pores and as a consequence, a thinner PA skin layer with greater flux being produced (Misdan et al.,

295 2013).

296 TFN NF membranes have received a similar level of attention as TFN RO membrane given

297 there are similar numbers of published articles on TFN RO and NF membranes. One of the earliest

298 mentions of the TFN NF membrane was reported by Lee et al. (2008) in 2008, one year after the

299 publication by Hoek and his co-workers (Jeong et al., 2007). Although the authors did not name the
16
ACCEPTED MANUSCRIPT
300 membranes as TFN, the concept of embedding nanomaterials, i.e. TiO2 nanoparticles into the PA

301 layer via IP was the same approach as that for TFN RO fabrication (Jeong et al., 2007). Instead of

302 hexane, which is most frequently used in the organic phase, Lee et al. (2008) used 1,1-dichloro-1-

303 fluroethane (HCFC) to better disperse high loadings of TiO2 nanoparticles (010 wt%). Results

304 showed that the membrane flux declined slightly while MgSO4 rejection increased with increasing

PT
305 TiO2 loading from zero to 5 wt%. Further increases in TiO2 loadings to 10 wt%, however, resulted in

306 an abrupt flux increase with extremely poor rejection (<5%). The authors attributed the poor

RI
307 separation performance to the weak mechanical strength of the membrane which resulted from the

SC
308 easier peeling-off of the PA-TiO2 layer from the substrate. In addition, the authors did not rule out

309 the possibility of a decrease in the degree of PA crosslinking resulting from the excessive use of

U
310 nanoparticles, which might act as end groups to interfere with the polymerization reaction.
AN
311 However, no characterisation data was provided in their work to verify this speculation. In addition

312 to this, the authors did not demonstrate how the improved dispersion quality of nanoparticles in
M

313 HCFC could lead to better distribution of nanoparticles in the PA layer synthesized.
D

314 To minimize particle agglomeration of TiO2 on the TFN NF membrane surface, Rajaeian et al.
TE

315 (2013) modified the TiO2 particle surface using an aminosilane coupling agent N-[3-

316 (trimethoxysilyl)propyl] ethylenediamine (AAPTS), before the particle was used in IP. It was assumed
EP

317 by the authors that the presence of silane functional groups on the TiO2 surface can effectively

318 minimize the probability of TiOTi oxygen bridge bonds between unmodified TiO2 nanoparticles
C

319 and consequent particle agglomeration. The modified TiO2, however, was added to the MPD
AC

320 aqueous solution instead of hexane solution. There was no discussion of the dispersion quality of

321 these modified nanoparticles in the non-polar solution. The experimental results showed that all the

322 TFN membranes made of the modified TiO2 exhibited better water flux than the control TFC

323 membrane; i.e. the pure water flux increased progressively from 11.2 L m-2 h-1 of TFC membrane to

324 ~13.2, ~20 and 27 L m-2 h-1 of TFN membranes, respectively, corresponding to the particle loading of

325 0.005, 0.05 and 0.1 wt%, when tested at ~7.5 bar (110 psi). NaCl rejection was slightly improved to
17
ACCEPTED MANUSCRIPT
326 54% when a small quantity of modified TiO2 (i.e. 0.005 wt%) was incorporated into the membrane

327 because of the increased binding affinity of particles to PA network during IP process. The

328 improvement in salt rejection could not be maintained at the highest loading of modified TiO2,

329 mainly because of the decrease of cross-linking in the PA layer. It is further elucidated by the authors

330 that the IP reaction would be impaired by the amine functional groups grafted on TiO2 surface when

PT
331 high loadings of modified TiO2 was added as the presence of amine groups on TiO2 surface tended to

332 compete with MPD for reaction with TMC.

RI
333 While Hu et al. (2012) reported the potential use of solid SiO2 nanoparticles (particle size: 11

SC
334 nm) to improve the separation characteristics of TFN NF membrane, Wu et al. (2013) explored the

335 possibility of incorporating mesoporous silica nanoparticles (MSN) having average pore size of 2.2

U
336 nm into polypiperazine-amide layer. In order to enhance the compatibility between MSN and
AN
337 polymer matrix, 3-aminopropyltriethoxysilane (APS) was used to functionalize MSN with amino

338 groups, as illustrated in Figure 3. For detailed instructions on nanoparticles synthesis, the readers are
M

339 referred to the corresponding literature. Findings of this work showed that remarkable flux
D

340 improvement with only a slight drop in Na2SO4 rejection was achieved when only 0.03 wt% modified
TE

341 MSN (mMSN) was incorporated within the PA layer of TFN membrane. The increase in water flux can

342 be mainly attributed to the additional pathway created by mMSN for water molecules. HR-TEM
EP

343 image further confirmed the ordered mesoporous network of mMSN which was in a hexagonal

344 array. The increase in system free volume resulted from the disruption of polymer chain packing
C

345 upon nanoparticles addition is also considered as one of the factors leading to the flux increment
AC

346 (Hu et al., 2012). Similar to most of the TFN membranes discussed earlier, the excessive

347 incorporation of mMSN (>0.03 wt%) within the TFN membrane would negatively affect salt rejection

348 rate.

349

350 *****Figure 3. Schematic preparation process of modified mesoporous silica nanoparticles

351 (mMSN) (Wu et al., 2013)*****


18
ACCEPTED MANUSCRIPT
352 Kim and Deng (2011) synthesized hydrophilized mesoporous carbons (H-OMCs) and

353 impregnated them in a thin PA film by in-situ IP. It was hypothesized that (1) synthesized H-OMCs

354 could be better dispersed in the aqueous phase compared to unmodified OMCs resulting in stronger

355 interaction with PA, and (2) the ordered pore layer of H-OMCs (median pore diameter: 3.68 nm)

356 could provide nano-flow paths for water molecules improving membrane flux production, provided

PT
357 the amount of nano-materials used is appropriate. These two hypotheses are supported by the data

358 in Figure 4(a) that shows the stability of well-dispersed H-OMCs in deionized water with no visible

RI
359 particles observed while Figure 4(b) shows the existence of hexagonal nanopores for H-OMCs. With

SC
360 respect to permeation characteristics, the authors emphasized the importance of having an

361 appropriate amount of H-OMCs in the polymer to achieve the nano-pore effect created by

U
362 mesoporous carbon. Of the range of concentrations studied (010 wt%), the use of 5 wt% H-OMCs
AN
363 was the most ideal loading to produce the most permeable TFN membrane with NaCl and Na2SO4

364 rejection recorded at 48% and 88%, respectively.


M

365
D

366 *****Figure 4. (a) Water suspensions of H-OMCs after three days of settling following mixing with
TE

367 water and (b) TEM images of OMCs samples (scale bar: 20 nm) (Kim and Deng, 2011)*****

368
EP

369 The application of TFN NF membrane is not limited to aqueous media. There are several

370 studies reporting on the performance of the membranes for non-aqueous separation process
C

371 (Namvar-Mahboub et al., 2014; Peyravi et al., 2014). Compared to conventional separations, solvent
AC

372 filtration using solvent resistant NF is an interesting alternative with benefits in terms of economy,

373 environment and safety. The pharmaceutical industry, for instance, uses a large amount of organic

374 solvents as reaction medium during production and formulation of active pharmaceutical

375 ingredients (Geens et al., 2007).

376 Namvar-Mahboub et al. (2014) made an attempt to fabricate solvent resistant NF (SRNF)

377 membranes by incorporating functionalized UZM-5 nanoparticles in the concentration range of zero-
19
ACCEPTED MANUSCRIPT
378 0.2 w/v% into PA layer made of MPD and TMC. According to the authors, these functionalized

379 zeolites with an average particle size of 73 nm could interact well with TMC monomers via covalent

380 bonding to form a PA network as illustrated in Figure 5. However, the increase in organic phase

381 viscosity coupled with an increase in miscibility between aqueous and organic phase at higher zeolite

382 loading led to the formation of a less cross-linked PA film. This, as a consequence, resulted in a

PT
383 gradual increase in solvent flux from 10.4 to 16.9 L m-2 h-1 with increasing zeolite loading from zero

384 to 0.2 w/v% when tested at 15 bar with feed solution consisting of 20 wt% lube oil, 40 wt% MEK and

RI
385 40 wt% toluene. The increase in the solvent flux can be explained by the new transport pathways

SC
386 created by functionalized USM-5 for permeation of MEK and toluene. It was reported that the

387 modified USM-5 possessed an average pore diameter of 16.8 , i.e. larger than the kinetic diameter

U
388 of MEK (5.2 ) and toluene (6.1 ) but much smaller than those of the oil marcomolecules. However,
AN
389 the separation efficiency of SRNF membrane against oil molecules was negatively affected when

390 >0.05 w/v% nanoparticles were embedded into PA layer. This was considered to be due to significant
M

391 agglomeration of nanoparticles, which was evidenced on the top surface of TFN0.2 membrane
D

392 causing the PA structure to distort and lowering its degree of cross-linking.
TE

393 On the other hand, an extraordinarily high methanol flux was reported by Peyravi et al.

394 (2014) when they incorporated functionalized TiO2 nanoparticles into PA layer of SRNF membrane
EP

395 made of ethylenediamine and isophthaloyl chloride. By incorporating nanoparticles functionalized

396 by different compounds, i.e. thionyl chloride (TCI), monoethanolamine (MEOA) and
C

397 triethylenetetramine (TETA), they reported that the resultant SRNF membranes could achieve
AC

398 methanol flux between 120 and 140 L m-2 h-1 when the membranes were tested at only 5 bar. In

399 comparison to the performance of other SRNF membranes reported in the literature such as

400 commercial StarMem-120 and StarMem-188 (Geens et al., 2007), high flux TFC membrane (Jimenez

401 Solomon et al., 2012), in-house made TFC membrane (Kosaraju and Sirkar, 2008) and mixed matrix

402 membrane (Sani et al., 2015), the methanol flux reported by Peyravi et al. (2014) was very high for a

403 membrane used in solvent medium. The latter membranes demonstrated methanol flux at 3.65.5 L
20
ACCEPTED MANUSCRIPT
404 m-2 h-1 bar-1, 7.046.0 L m-2 h-1 (at 30 bar), ~6.41 L m-2 h-1 (at 4 bar) and ~73100 L m-2 h-1 (at 10 bar),

405 respectively. No explanation was provided by the authors as to why functionalized TiO2 could be so

406 powerful in improving methanol flux of TFN membrane by many orders of magnitude and still could

407 maintain good selectivity, and no other groups have reported flux of similar magnitude. Typically,

408 the significant low solvent flux of membrane (compared to water flux) is mainly attributed to the

PT
409 high difference in surface tension between solvent and membrane material which increases

410 transport resistance of solvent and reduces its permeation rate (Sani et al., 2015).

RI
411

SC
412 *****Figure 5. Schematic of chemical interaction between functionalized UZM-5 nanoparticles and

413 TMC monomers during interfacial polymerization (Namvar-Mahboub et al., 2014)*****

U
414
AN
415 Although Basu et al. (2009) have also incorporated nanomaterials metal-organic

416 frameworks (MOFs) to prepare SRNF membrane, their method was different from the work of
M

417 Namvar-Mahboub et al. (2014) and Peyravi et al. (2014). Instead of employing IP, the authors
D

418 developed a thin polydimethlysiloxane (PDMS) layer containing MOFs on the surface of microporous
TE

419 polyimide (PI) via typical coating method, i.e. PDMS solution (10 wt% PDMS in hexane) was coated

420 on top of PI support by keeping the support plate inclined at 60o. Of the four different types of MOFs
EP

421 studied (Cu3(BTC)2, MIL-47, MIL-53(Al) and ZIF-8), none of the SRNF membranes prepared could

422 achieve solvent flux >1.0 L m-2 h-1 bar-1 when tested using isopropanol solution containing 17.5 mol
C

423 L-1 Rose Bengal (MW: 974 g mol-1). The extremely low solvent permeability of these membranes
AC

424 could be mainly due to the formation of thick PDMS layer (3035 m) resulting from the coating

425 method.

426 Although experiences with solvent resistant TFN NF membrane are still scarce and a

427 successful application of TFN NF membrane in industry has not yet eventuated, the research of TFN

428 RO/NF membranes for aqueous medium filtration will be helpful to progress solvent resistant TFN

429 NF membrane development.


21
ACCEPTED MANUSCRIPT
430 2.3 Forward osmosis

431 In comparison to the TFN NF/RO membrane applications, the use of TFN for FO process was

432 attempted only recently. The first relevant article was published by Tang and his co-authors in 2012

433 (Ma et al., 2012). Up to date, less than 20 TFN articles have been published on FO applications with

434 the main application on brackish/seawater desalination process.

PT
435 The promising results obtained by Lind et al. (2009) by incorporating zeolite nanoparticles in

436 TFN RO membrane motivated Tang and his co-authors (Ma et al., 2012) to further explore the

RI
437 potential of zeolite NaY nanoparticles for TFN FO process. In the range of 0.020.4 wt/v% zeolite

SC
438 loading, the FO water flux of the TFN membrane tended to increase with increasing loading from

439 0.02 to 0.1% but decrease with further increasing zeolite loading to 0.4%. Under specific testing

U
440 conditions (Feed solution: 10 mM NaCl; Draw solution: 1 M NaCl), the most permeable TFN0.1
AN
441 membrane demonstrated FO flux of ~32 L m-2 h-1 (for active-layer-facing-draw-solution (AL-DS)

442 orientation) and ~15 L m-2 h-1 (for active-layer-facing-feed-solution (AL-FS) orientation) in comparison
M

443 to ~22 and ~10 L m-2 h-1 recorded for the control TFC membrane and ~24 and ~13 L m-2 h-1 for TFN0.4
D

444 membrane. The increase in water flux upon low loading of zeolite and a decrease in water flux at
TE

445 high zeolite loading are attributed to the porous nature of zeolite and formation of thicker PA layer,

446 respectively, as elucidated by the author. Similar to Lind et al. (2009), Tang and his co-authors (Ma et
EP

447 al., 2012) did not correlate the dispersion quality of nanoparticle in organic solution with its

448 distribution in thin PA layer and further FO performance.


C

449 Progress has also been made by Rahimpour and his co-authors (Amini et al., 2013; Niksefat
AC

450 et al., 2014) in an effort to study the impact of functionalized MWCNTs and SiO2 nanoparticles on

451 the morphology and performance of TFN membrane for FO application. For both studies it was

452 reported that the most permeable TFN membrane was produced when 0.1 wt/v% nanofillers were

453 added. When tested using 10 mM NaCl as feed solution and 1 M NaCl as draw solution, the TFN

454 membrane incorporated with 0.1% SiO2 demonstrated approximately 36 L m-2 h-1 (for AL-DS) and 23 L

455 m-2 h-1 (for AL-FS) while TFN membrane with 0.1% functionalized MWCNTs showed extraordinary
22
ACCEPTED MANUSCRIPT
456 high flux of around 96 and 37 L m-2 h-1, respectively. Although the authors attributed such high fluxes

457 of the TFN membrane to the presence of nano-channels (in CNTs) in the top PA layer, which created

458 a smooth pathway for water molecules, empirical evidences from the literature suggest that it is

459 difficult to fabricate a membrane of extremely high flux without sacrificing selectivity (i.e. low

460 reverse solute flux in FO application). Contradictory results were also provided by the authors in

PT
461 which the TFN membrane displayed poor NaCl rejection (73%) but was still able to keep reverse

462 solute flux at a very low level, i.e. ~5 and 3 g m-2 h-1 for AL-DS and AL-DS orientation, respectively. It

RI
463 must be kept in mind that developing TFN membrane with high FO flux/low reverse solute flux will

SC
464 necessitate a major paradigm shift in membrane development.

465 Instead of adding nanomaterials into PA layer of the composite membrane, Emadzadeh et

U
466 al. (2014b, 2014c) made the first attempt to embed commercial TiO2 nanoparticles (Degussa P25)
AN
467 into the microporous substrate (via direct blending method) of the composite membrane for FO

468 process. Compared with NF/RO process, the role of hydrophilic substrate is much more significant in
M

469 FO process as both top PA selective layer and bottom substrate are simultaneously contacted with
D

470 aqueous solutions during filtration process, but of different osmotic pressures. Specifically, two
TE

471 types of concentration polarization phenomena can take place in osmotic-drive membrane

472 processes. They are external concentration polarization which is caused by the build-up of rejected
EP

473 solute at the membrane active layer surface and internal concentration polarization which is caused

474 by the establishment of polarized layer within porous layer of composite membrane (Cath et al.,
C

475 2006). Experimental results showed that the best performing FO membrane was produced when 0.5
AC

476 wt% TiO2 nanoparticles were added to the microporous substrate made of polysulfone

477 (PSF)/polyvinylpyrrolidone (PVP) (Emadzadeh et al., 2014c). This TFN membrane demonstrated

478 water flux of ~57 L m-2 h-1 (for AL-DS) and ~30 L m-2 h-1 (for AL-FS) when tested using 10 mM NaCl as

479 feed solution and 2 M NaCl as draw solution. These values were ~90% and 71.5% higher than the

480 flux achieved by the control TFC membrane under the same test conditions. The flux improvement

481 was supported by the development of long finger-like voids extended from the top to the bottom of
23
ACCEPTED MANUSCRIPT
482 the substrate cross-section as evidenced from SEM images, resulting in smaller S values (S =

483 thickness tortuosity/porosity) and reduced internal concentration polarization. Further increase in

484 TiO2 loadings to 0.75 and 1.0 wt% was able to increase membrane water flux, but the reverse solute

485 flux of membrane was compromised. It is because excessive loading of TiO2 tends to cause

486 significant nanoparticle agglomeration on the substrate surface (as confirmed by SEM surface

PT
487 analysis), which negatively affects the degree of PA cross-linking. In addition to TiO2, the impact of

488 adding zeolite NaY nanoparticles into the substrate matrix on the FO performance of TFN membrane

RI
489 could be found elsewhere (Ma et al., 2013), where it was also reported that 0.5 wt% was the ideal

SC
490 nanofiller loading for nanocomposite substrate in order to achieve a good combination of water flux

491 and solute rejection in TFN membrane. However, the substrate surface properties of these two

U
492 studies were not the same as Emadzadeh et al. (2014c) reported S value of 0.42, contact angle 64o
AN
493 and surface roughness (Ra) 14.1 nm for the PSF membrane incorporated with 0.5 wt% TiO2, while Ma

494 et al. (2013) reported 0.34, 50o and 21.7 nm, respectively in the PSF-NaY (0.5 wt%) membrane. In
M

495 addition to the direct blending method, other fabrication methods such as coating/deposition,
D

496 chemical grafting, layer-by-layer assembly, cross-linking, etc could also be employed to prepare
TE

497 nanocomposite substrates. Detailed discussion on each of the fabrication methods can be found in

498 the recently published review article written by Yin and Deng (2015).
EP

499 Development of a substrate of higher hydrophilicity is also important in mitigating

500 membrane organic fouling, in particular for pressure retarded osmosis (PRO) applications
C

501 (Emadzadeh et al., 2014a). As shown in Figure 6, the incorporation of only 0.5 wt% TiO2 into PSF
AC

502 substrate of TFN membrane could significantly decrease BSA accumulation in the porous sub-layer,

503 making organic fouling resistance better than the typical TFC membrane. As elucidated by the

504 authors, the increase in substrate hydrophilicity upon TiO2 addition induced the formation of a water

505 layer at the surface of the substrate membrane pores, preventing BSA adsorption and thus

506 increasing the effective osmotic pressure difference (effect) across the membrane. This, as a result,

507 reduces the tendency of flux decline during operation.


24
ACCEPTED MANUSCRIPT
508

509 *****Figure 6. Comparison between TFN and typical TFC membrane, (a) effect of nanocomposite

510 substrate on organic fouling of composite membrane in PRO mode (test conditions: feed solution:

511 10 mM NaCl with 200 mg/L bovine serum albumin, draw solution: 2.0 M NaCl, cross-flow velocity:

512 32.72 cm/s on both sides of the FO membrane and temperature: 25 C) and (b) profile of effective

PT
513 osmotic pressure difference across TFN and TFC membrane (with and without fouling) in PRO

514 mode (Emadzadeh et al., 2014a)*****

RI
515

SC
516 2.4 Other applications

517 A thorough literature search revealed that there is no PA TFN membrane reported for applications

U
518 other than NF/RO/FO processes, except only one reported by Fathizadeh et al. (2013) for
AN
519 pervaporative dehydration of aqueous alcohol solutions. Of the range of zero to 0.15 wt% NaX

520 nanoparticles studied, the authors found that the TFN membrane incorporated with 0.05 wt% NaX
M

521 could achieve the highest water flux (~2750 g m-2 h-1) with water selectivity of approximately 200
D

522 when tested with 90 wt% aqueous isobutanol solution at permeate pressure of 1 mmHg. The values
TE

523 were around 72% and 8% better than that of typical TFC membrane. It was explained that water

524 molecules tend to flow preferentially through hydrophilic molecular sieve of NaX which having pore
EP

525 size (7.4) between the molecular size of water (2.65) and isobutanol (7.7). Membrane for gas

526 separation is also of great importance, but the little attention paid on the use of conventional TFC
C

527 membrane for gas separation whether in the past or present has led to research on TFN membranes
AC

528 not being attractive.

529

530 3. Challenges of TFN membrane fabrication

531 Although the potential of PA TFN membrane for water treatment process in particular has been

532 demonstrated by many researchers over the past 78 years, there remain several key challenges
25
ACCEPTED MANUSCRIPT
533 related to TFN membrane making. Addressing these challenges is the key for further development of

534 TFN membrane for industrial applications.

535 One of the main problems encountered during TFN membrane fabrication is the

536 agglomeration of nanoparticles in the PA layer. The TEM images in Figure 7 show particle

537 agglomeration in the thin PA layer, which is likely to reduce the active surface area of nanoparticles

PT
538 and/or even result in the formation of defects (holes) in the PA structure. The main reason

539 attributed to this phenomenon is the low dispersion rate of nanomaterials in the solutions (either

RI
540 aqueous or organic) used in the IP process. The non-uniform dispersion of particles, particularly in

SC
541 non-polar organic solvent, makes them very easy to aggregate and unable to uniformly spread out

542 on the top microporous substrate surface. As a result, a certain part of the PA dense layer contains

U
543 no nanomaterials at all.
AN
544

545 *****Figure 7.TEM images of the cross-section of (a) TFN membrane with silica nanoparticles
M

546 (MCM-41) embedded (Yin et al., 2012) and (b) TFN membrane with NaA zeolite embedded (Huang
D

547 et al., 2013a).*****


TE

548

549 It is generally known that hydrophilic nanomaterials can be better dispersed in the aqueous
EP

550 phase than in the organic phase, but when excessive aqueous phase is removed by applying a soft

551 rubber roller on the substrate surface (during IP process), a large amount of nanoparticles are
C

552 removed together with the amine solution, leaving only a small amount of nanoparticles in the
AC

553 substrate pores. Therefore, nanoparticles, which are commonly hydrophilic, should be surface-

554 modified in order to make them more compatible with the organic phase. However, a literature

555 search revealed that most research has preferred to disperse nanomaterials in amine aqueous

556 solution and this could be mainly due to the difficulties in producing a homogenous nanomaterial-

557 organic mixture. The high surface energy of inorganic nanomaterials (5002000 mJ cm-2) coupled
26
ACCEPTED MANUSCRIPT
558 with high inter-particle interactions are the main reasons causing them to aggregate easily in nature

559 (Caseri, 2000).

560 It has been previously reported that instead of using hydrophilic nanoparticles, embedding

561 PA layers with hydrophilic nanotubes (e.g. aluminosilicate SWNTs and MWCNTs) could create a

562 preferential pathway for water molecules, leading to an improvement in water flux (Amini et al.,

PT
563 2013; Baroa et al., 2013; Ghanbari et al., 2015b). Nanotubes, however, have lengths between 10

564 and 50 m (equivalent to 10,000 and 50,000 nm) (Amini et al., 2013; Kim et al., 2012; Shen et al.,

RI
565 2013; Wu et al., 2010). They cannot be accommodated in a thin PA layer of 100500 nm (Lau et al.,

SC
566 2014) unless all of them are horizontally oriented, which is highly unlikely for IP processes because

567 of the large negative entropy that cannot be compensated by the enthalpy term due to the poor

U
568 interaction between nanoparticles and PA. If this orientation indeed occurs, it is still contradictory to
AN
569 the claim of many researchers that water flows through vertically aligned channels. Nevertheless,

570 randomly arranged nanotubes have to certain extent created a relatively less resistant pathway for
M

571 water molecules to pass through, leading to flux improvement with minimum change in salt
D

572 rejection as evidenced in previous works. Additional research needs to be carried out to verify the
TE

573 water and ion transport mechanism for these membranes, as improved flux has also been

574 demonstrated for non-porous nanoparticles such as Ag. Changes in free volume may be more
EP

575 relevant to the transport mechanism, and Xie et al. (2014) have shown free volume changes in

576 poly(vinyl alcohol) (PVA) upon the addition of silica nanoparticles and related this to changes in
C

577 permeability and selectivity. However, few studies that have examined such effects as gaining access
AC

578 to instruments capable of characterising thin films in this manner is difficult.

579 Lastly, the lack of chemical interaction between nanomaterials and PA matrix is likely to

580 cause nanomaterials to easily leach out during IP and/or filtration process, causing the prepared TFN

581 membrane to be less effective and the efficiency of nanoparticle use during production to be low.

582 Research findings from previous works have always showed little problem with the incompatibility

583 between inorganic nanomaterials and organic PA layer. However, no thorough study has reportedly
27
ACCEPTED MANUSCRIPT
584 been carried out so far to evaluate nanoparticle leaching during a long periods of filtration. Is the

585 physical interaction sufficient to hold the nanomaterials in the PA layer especially under the high-

586 pressure operation? What will be the outcome if the surface of nanomaterials is chemically bound to

587 the PA structure? Will the chemical interaction be better for improving TFN performance stability

588 over long term operation? Several researchers have recently made efforts to synthesize

PT
589 nanomaterials surface functionalized with NH2 groups in an attempt to improve PA-nanomaterials

590 interaction (Amini et al., 2013; Wu et al., 2013; Emadzadeh et al., 2014d; Peyravi et al., 2014), but in-

RI
591 depth analysis on the PA layer chemistry upon addition of functionalized nanomaterials has never

SC
592 been reported. Detailed study on the molecular interaction between PA and functionalized

593 nanomaterials is worthy of special attention.

U
594
AN
595 4. Possible approaches to overcome TFN fabrication problems

596 Undoubtedly, the challenges in TFN membrane fabrication as mentioned in the previous section
M

597 have motivated many dedicated scientific investigations in which some of the innovative approaches
D

598 can be potentially employed to address these challenges.


TE

599

600 4.1 Surface modification of nanomaterials


EP

601 One of the simple yet effective approaches to improve the dispersion quality of nanomaterials in

602 non-polar organic solvent is through surface modification of nanoparticles. The modified
C

603 nanoparticles usually contain specific functional groups so as they can disperse homogenously in the
AC

604 non-polar organic solution during IP process to reduce the extent of particle agglomeration in PA

605 layer or to have chemical bonding with the PA network.

606 As reported by Shen et al. (2013), surface modification of MWCNTs with acid solution (HNO3

607 and H2SO4 mixture) followed by microemulsion polymerization of methyl methacrylate (MMA)

608 monomer resulted in highly soluble nanomaterials in organic solvents. As shown in Figure 8, the

609 synthesized polymethyl methacrylate (PMMA)-MWCNTs formed a very stable, uniform dark
28
ACCEPTED MANUSCRIPT
610 dispersion in toluene even after more than 1 month of standing. With this promising dispersion

611 quality of nanomaterials, it is likely to minimize the precipitation of nanomaterials in organic solvent

612 during IP and further to reduce the extent of particle agglomeration in the PA layer.

613

614 *****Figure 8.Direct observation on the dispersion rate of PMMAMWNTs in toluenewater after

PT
615 1 month (Shen et al., 2013).*****

616

RI
617 To assist proper dispersion of UZM-5 zeolite nanoparticles in organic (hexane) solvent,

SC
618 Namvar-Mahboub et al. (2014) functionalized the nanoparticles using amino silane coupling agent

619 3-aminopropyldiethoxymethylsilane (APDEMS). According to the authors, the presence of amino

U
620 functional group (NH2) on the nanozeolite surface is not only enables nanoparticles to disperse
AN
621 better in organic solvent but also possibly form a covalent bond with TMC molecules, improving the

622 interaction of inorganic nanoparticles with the organic PA structure. The use of amino silane
M

623 coupling agent for modification of nanoparticles was also reported by Rajaeian et al. (2013) in which
D

624 they modified the surface of TiO2 nanoparticles using N-[3-(trimethoxysilyl)propyl]ethylenediamine


TE

625 (AAPTS). The reduced oxygen bridge between TiO2 nanoparticles by AAPTS modification is said to be

626 effective for minimization of the negative impact of particle agglomeration in the PA layer. However,
EP

627 in this work, the authors only demonstrated good dispersion of AAPTS-TiO2 in the aqueous phase

628 instead of the organic phase. Surface modification of mesoporous silica nanoparticles was also
C

629 conducted by Wu et al. (2013) using 3-aminopropyltriethoxysilane (APS) to prepare TFN membrane
AC

630 for NF application. Similar to Rajaeian et al. (2013), the authors did not report the dispersion of

631 nanoparticles in the organic solvent. Explanation was given only on the possibility of chemical

632 interaction between functionalized silica nanoparticles and PA structure.

633 Recently in 2014, Peyravi et al. (2014) made an attempt to functionalize the surface of TiO2

634 nanoparticles using two amine reagents, i.e. monoethanolamine (MEOA) and triethylenetetramine

635 (TETA). The aminated TiO2 nanoparticles were claimed to provide a suitable surface for dispersion of
29
ACCEPTED MANUSCRIPT
636 nanoparticles in acyl chloride solution during IP process. However, details were not provided for the

637 dispersion quality of the functionalized nanoparticles in aqueous/organic solution. The effectiveness

638 of this approach for TFN membrane fabrication was also not very convincing as the results from AFM

639 and SEM analysis revealed significant particle agglomeration on the PA top surface. On the other

640 hand, a review article by Faure et al. (2013) had shown other potential chemical additives for

PT
641 improving the dispersion stability of TiO2 nanoparticles in organic media, but these approaches have

642 not been used for TFN membrane fabrication. For instance, polybutenesuccinimide pentamine

RI
643 (OLOA 370) was proposed by Erdem et al. (2000) as a stabilizer to improve dispersion quality of

SC
644 hydrophilic TiO2 nanoparticles in cyclohexane. The amine end group of OLOA 370 is able to accept

645 protons or donates an electron pair which results in a concentration-dependent negative zeta

U
646 potential on the TiO2 surface in low dieletric organic media. In other words, the amine end group on
AN
647 this polymeric stabilizer can readily interact with reactive hydroxyl group on TiO2 surface, leading to

648 highly dispersed nanoparticles in organic solvent.


M

649 While researchers have proposed methods to modify the surface of nanoparticles in an
D

650 effort to minimize the negative impacts of nanomaterial agglomeration in PA layer of TFN
TE

651 membrane, there is still a lack of detailed understanding as to how such modifications lead to

652 improved membrane morphologies and current results are inconsistent. There are many cases
EP

653 where scientists did not really demonstrate the dispersion quality of modified nanomaterials in non-

654 polar solution and/or perform in-depth analysis on the characteristics of nanomaterial-PA layer (
C

655 Rajaeian et al., 2013; Wu et al., 2013; Namvar-Mahboub et al., 2014; Peyravi et al., 2014). Perhaps,
AC

656 future research should focus on the development of advanced/novel nanomaterials that could

657 readily disperse in non-polar solvent to achieve a homogenous distribution of the materials in PA

658 layer. Furthermore, it would also be interesting if quantitative modelling of separation could be

659 established for TFN membranes, particularly those incorporated with nanotubes/mesoporous

660 nanoparticles. Such models may confirm or identify mechanisms for water and ion transport in TFN

661 membranes, and thereby assist in informed design of better performing membranes.
30
ACCEPTED MANUSCRIPT
662 4.2 Use of metal alkoxides

663 As solid hydrophilic nanoparticles are readily aggregated when mixed with non-polar organic

664 solvent, good dispersion of them in non-polar solvent is difficult to achieve. To solve this problem,

665 Kong et al. (2011) used metal alkoxides tetra isopropoxide (TTIP) which are readily dissolved in

666 organic solvent to replace raw TiO2 nanoparticles in preparing TFN membrane for RO application.

PT
667 Besides showing good dispersion in hexane, metal alkoxides could be hydrolysed to produce smaller

668 inorganic nanoparticles (and organic alcohol) either during or after the IP process. Results showed

RI
669 that in comparison to typical TFC membrane, the water flux of TFN membrane was improved with

SC
670 only a slight decrease in salt rejection when a small amount of TTIP solution (~0.1 wt%) was added to

671 the hexane solvent. However, when the performance of TTIP-TFN membrane was further compared

U
672 with TFN membrane prepared from other metal alkoxides phenyl triethoxysilane (PhTES), it was
AN
673 found that PhTES-TFN membrane was much better as the latter membrane could overcome the

674 trade-off effect between water permeability and solute rejection. The reasons for these variations in
M

675 performance were not identified by the authors. The potential of using metal alkoxides in TFN
D

676 membrane fabrication is still not very clear as there is no other relevant article from open literature.
TE

677

678 4.3 Modified/novel IP techniques


EP

679 In terms of IP process, Lau et al. (2012) revealed the continuous concerted efforts made by

680 researchers in preparing TFC membranes with improved interfacial properties via modified IP
C

681 procedure. Some of these modifications are introduction of secondary amine solution to react again
AC

682 with unreacted acyl chloride groups by placing the substrate membrane in an aqueous solution for a

683 second time (Zou et al., 2010) or employment of intermediate organic solvent between the aqueous

684 amine solution and the organic acid chloride solution so as to reduce possible pin-hole formation on

685 PA layer (Verssimo et al., 2005). A further review will be conducted in this work to explore how a

686 modified IP technique could be possibly adopted for composite membrane, in particular PA TFN

687 membrane fabrication process.


31
ACCEPTED MANUSCRIPT
688 As reported in the work of Kong et al. (2010), an intermediate pre-seeding hexane solution

689 containing low concentration of TMC (0.020.05 wt% TMC, 0.050.6 wt% zeolite and 515 wt%

690 ethanol) was introduced between 2 wt% MPD aqueous solution and 0.1 wt% TMC organic solution

691 to assist the dispersion of zeolite nanoparticles on the top surface of PSF substrate, as illustrated in

692 Figure 9(a). The FESEM images (see Figure 9(b) and (c)) clearly showed well-dispersed zeolite

PT
693 nanoparticles throughout the top PSF support by using pre-seeding solution containing 0.4 wt%

694 zeolite. With respect to RO performance, the TFN membrane fabricated by a pre-seeding method

RI
695 with 0.2 wt% zeolite loading showed greater water permeability of 4.210-12 m Pa-1 s-1 in comparison

SC
696 to 2.9 and 1.910-12 m Pa-1 s-1 recorded in TFN membrane prepared by adding zeolite into TMC-

697 hexane and MPD aqueous solution. Even though the water permeability of TFN membrane prepared

U
698 by the modified IP was remarkably improved, high NaCl rejection (97.4%) was not compromised. The
AN
699 authors attributed the promising results to the well-defined zeolite pores which serve as a short cut

700 for water permeation. Rejection experiments conducted using neutral organic solutes of different
M

701 molecular sizes showed that the TFN membrane prepared exhibited very similar pore size (~0.79
D

702 nm) in comparison to the pore size of incorporated zeolite Y (~0.74 nm), suggesting the pore
TE

703 channels of nanoparticles were still accessible for transport following the IP process.

704
EP

705 *****Figure 9. (a) Schematic representation of the TFN membrane fabrication via pre-seeding

706 method and FESEM images of PSF support surface before (b) and after (c) pre-seeding process
C

707 (Kong et al., 2010)*****


AC

708

709 A dynamic IP process illustrated in Figure 10 was used by An et al. (2012) to form PA layer on

710 a microporous substrate. This dynamic IP was employed to fabricate TFC membrane for

711 pervaporation process, and the results were quite interesting with respect to surface pattern of PA

712 layer. The powerful centrifugal force created during spinning process tends to align molecular chains

713 in the growing interfacial PA film in the horizontal direction, leading to the formation of thinner and
32
ACCEPTED MANUSCRIPT
714 smaller free volume PA structure. The uniquely formed PA structure was verified by positron

715 annihilation spectroscopy (PAS) in which lower values of S parameter was obtained with increasing

716 spin coating rate. In terms of performance, the PA film formed from applying dynamic IP process

717 showed simultaneous improvement in flux and selectivity, overcoming the trade-off phenomenon

718 which is commonly known in conventional IP process (static). Although this novel IP process is only

PT
719 attempted for TFC membrane making, it could possibly be used for TFN membrane preparation as

720 well (Fu et al., 2014). It is because the centrifugal force induced by a spin coater is likely to spread

RI
721 the organic solution (which contains nanoparticles) from the axis of rotation towards the outer

SC
722 substrate edge, leading to good dispersion of the nanoparticles over the entire substrate surface. In

723 addition, the advantage of spin coating in forming ultrathin PA selective layer (around 200 nm) might

U
724 potentially reduce water transport resistance during the filtration process and enhance membrane
AN
725 water productivity. This fabrication approach is one of the subjects deserving greater research

726 attention.
M

727
D

728 *****Figure 10. Establishing PA layer on top of microporous substrate through static and dynamic
TE

729 interfacial polymerization (An et al., 2012).*****

730
EP

731 4.4 Alignment of nanotubes/fillers

732 Although there is not any research work conducted so far on alignment of nanotubes in composite
C

733 membranes prepared via IP process, one can still find several innovative approaches that have been
AC

734 recently employed for asymmetric mixed matrix membrane fabrication to control nanofiller

735 orientation either vertically or horizontally in the membrane structure. Alignment of nanofillers, for

736 example CNTs, in a membrane matrix could reduce interfacial voids between the CNTs and the

737 polymer, minimizing discontinuous and tortuous paths for water molecules to be transported.

738 Wu et al. (2014) reported a new method to align MWCNTs in polystyrene (PS) membrane by

739 means of an alternating electric field to achieve an even dispersion of MWCNTs. Figures 11 (a) and
33
ACCEPTED MANUSCRIPT
740 (b) show a schematic diagram of the electrical field alignment apparatus used for MWCNT/PS

741 membrane preparation and the micrographs of the membranes prepared from 3 wt% MWCNTs at

742 different alternating electric field frequencies. As can be seen, the extent of MWCNTs aggregation

743 and maldistribution is significantly reduced in the electro-cast membranes. Increasing the frequency

744 of the electric field from 1 Hz to 100 Hz was also found to further improve dispersion of CNTs in the

PT
745 membrane matrix. These results can be explained by the electric field which aligns the conductive

746 MWCNTs in the direction of the field and causes them to exclude each other perpendicularly via

RI
747 dipoledipole interactions. Because of this, the electro-cast membranes showed much improved

SC
748 oxygen permeabilities over the control membranes.

749 Kim et al. (2014) on the other hand developed a novel in-situ bulk polymerization method to

U
750 prepare vertically aligned carbon nanotube (VACNT)/polymer composite membrane for both gas and
AN
751 water separation processes. In order to prevent CNT condensation that could disturb CNT

752 orientation during liquid phase processing, VACNT array was infiltrated with styrene monomer with
M

753 a certain amount of polystyrene-polybutadiene block copolymer that acts as a plasticizer. SEM
D

754 images identified that well-aligned CNTs were embedded in a high-density polymer matrix free of
TE

755 any macroscopic voids or structural defects.

756 In addition to these two possible alignment approaches, Goh et al. (2014) in their recent
EP

757 review article have summarized various contemporary approaches that could be employed to align

758 CNTs either horizontally or vertically in polymer matrix. However, more research is needed to
C

759 determine if these approaches could be used for thin PA layer synthesis process or it is only limited
AC

760 to asymmetric membrane structure.

761

762 *****Figure 11. (a) Schematic of an apparatus for electrical alignment of MWCNTs in polymer

763 matrix (b) Micrographs of 0.3 wt% MWCNTs dispersed in PS membrane with and without applying

764 electric field (Wu et al., 2014).*****

765
34
ACCEPTED MANUSCRIPT
766 5. Conclusions and future perspectives

767 The development of TFC membrane is an important research application in NF/RO technology for

768 water and wastewater treatment and was attracted wide attention since 1960s. In this paper, the

769 progress of new generation composite membrane TFN was reviewed for various aqueous media-

770 based separation processes, classified according to RO, NF and FO. Compared to typical TFC

PT
771 membrane, it is widely reported that TFN membrane has great potential in overcoming the problem

772 of the trade-off curve between permeability and selectivity. In addition, the introduction of inorganic

RI
773 fillers into PA layer show attractive characteristics in improving TFN membrane durability associated

SC
774 with bacterial, fouling and chlorine resistance. Despite the significantly greater achievements, there

775 are still some challenges encountered during TFN membrane fabrication. These include poor

U
776 dispersion quality of hydrophilic nanomaterials in non-polar organic solvent and their agglomeration
AN
777 within the PA layer, lack of chemical interaction between nanomaterials and organic PA layer as well

778 as alignment control of nanotubes in PA layer. Dedicated scientific investigations in recent years,
M

779 however, have offered several innovative approaches, e.g. modification of hydrophilic nanofillers
D

780 and conventional IP process that can be potentially employed to address the quest for defect-free
TE

781 organic PA/inorganic nanofillers layer fabrication. More research in this area is still needed to

782 develop TFN membrane with greater performance efficiency, reliability and stability for industrial
EP

783 implementation. Among them are to develop advanced/novel nanomaterials with specific pore

784 structure/charge properties that could readily disperse well in non-polar solvent and/or to improve
C

785 IP technique to achieve homogenous distribution of the materials in PA layer. In terms of


AC

786 characterization techniques, it is highly recommended to use the methods that can give resolutions

787 of the observation in the angstrom scale. This is of significant importance to understand the

788 interaction between PA and nanomaterials (either physically or chemically) as well as actual

789 transport mechanism of solutes in the TFN membranes, particularly those incorporated with

790 nanotubes/mesoporous nanoparticles. Developing quantitative separation modelling for TFN

791 membranes made of different types of nanomaterials is also an important subject as it could provide
35
ACCEPTED MANUSCRIPT
792 more fundamental knowledge on the ions transport in the PA layer of TFN membrane. Last but not

793 least, the selection and use of nanofillers for TFN membrane fabrication should depend on the

794 compounds (such as type of ions, bacteria, chlorine, etc) to be removed from the feed solution. It is

795 because there is no TFN membrane that is universally applicable for any types of water applications.

796

PT
797 Acknowledgements

798 The corresponding author wishes to acknowledge the Australian Government Department of

RI
799 Education for providing the 2015 Endeavour Research Fellowship which initiates the research

SC
800 collaboration between Universiti Teknologi Malaysia (UTM) and Victoria University (VU).

801

U
802 References
AN
803 Amini, M., Jahanshahi, M., Rahimpour, A., 2013. Synthesis of novel thin film nanocomposite (TFN)

804 forward osmosis membranes using functionalized multi-walled carbon nanotubes. Journal of
M

805 Membrane Science 435, 233241.


D

806 An, Q., Hung, W., Lo, S., Li, Y., Guzman, M. De, 2012. Comparison between Free Volume

807 Characteristics of Composite Membranes Fabricated through Static and Dynamic Interfacial
TE

808 Polymerization Processes. Macromolecules 45, 34283435.


EP

809 Aquaporin A/S, http://www.aquaporin.dk/1/home.aspx (assessed on 20 April 2015)

810 Bakajin, O., Noy, A., Fornasiero, F., Park, H.-G., Holt, J., Kim, S., 2009. Membranes with
C

811 functionalized carbon nanotube pores for selective transport. Patent no: WO2009148959 A2.
AC

812 Publication date: 10 December 2009.

813 Baroa, G.N.B., Lim, J., Choi, M., Jung, B., 2013. Interfacial polymerization of polyamide-

814 aluminosilicate SWNT nanocomposite membranes for reverse osmosis. Desalination 325, 138

815 147.

816 Barrer,R.M., James, S.D., 1960. Electrochemistry of crystal-polymer membrane. 1. Resistance

817 measurements. Journal of Physical Chemistry 64, 417421.


36
ACCEPTED MANUSCRIPT
818 Basu, S., Maes, M., Cano-Odena, A., Alaerts, L., De Vos, D.E., Vankelecom, I.F.J., 2009. Solvent

819 resistant nanofiltration (SRNF) membranes based on metal-organic frameworks. Journal of

820 Membrane Science 344, 190198.

821 Ben-Sasson, M., Lu, X., Bar-Zeev, E., Zodrow. K.R., Nejati, S., Qi, G., Giannelis, E.P., Elimelech, E.,

822 2014. In situ formation of silver nanoparticles on thin-film composite reverse osmosis

PT
823 membranes for biofouling mitigation. Water Research 62, 260270.

824 Caseri, W., 2000. Nanocomposites of polymers and metals or semiconductors: Historical

RI
825 background and optical properties. Macromol. Rapid Commun 21, 705722.

SC
826 Cath, T., Childress, a, Elimelech, M., 2006. Forward osmosis: Principles, applications, and recent

827 developments. Journal of Membrane Science 281, 7087.

U
828 Emadzadeh, D., Lau, W.J., Matsuura, T., Hilal, N., Ismail, A.F., 2014a. The potential of thin film
AN
829 nanocomposite membrane in reducing organic fouling in forward osmosis process. Desalination

830 348, 8288.


M

831 Emadzadeh, D., Lau, W.J., Matsuura, T., Ismail, A.F., Rahbari-Sisakht, M., 2014b. Synthesis and
D

832 characterization of thin film nanocomposite forward osmosis membrane with hydrophilic
TE

833 nanocomposite support to reduce internal concentration polarization. Journal of Membrane

834 Science 449, 7485.


EP

835 Emadzadeh, D., Lau, W.J., Matsuura, T., Rahbari-Sisakht, M., Ismail, A.F., 2014c. A novel thin film

836 composite forward osmosis membrane prepared from PSfTiO2 nanocomposite substrate for
C

837 water desalination. Chemical Engineering Journal 237, 7080.


AC

838 Emadzadeh, D., Lau, W.J., Rahbari-Sisakht, M., Daneshfar, A., Ghanbari, M., Mayahi, A., Matsuura, T.,

839 Ismail, A.F., 2014d. A novel thin film nanocomposite reverse osmosis membrane with superior

840 anti-organic fouling affinity for water desalination. Desalination

841 http://dx.doi.org/10.1016/j.desal.2014.11.019.
37
ACCEPTED MANUSCRIPT
842 Erdem, B., Sudol, E.D., Dimonie, V.L., El-aasser, M.S., 2000. Encapsulation of inorganic particles via

843 miniemulsion polymerization . I . Dispersion of titanium dioxide particles in organic media using

844 OLOA 370 as stabilizer. Journal of Polymer Science Part A: Polymer Chemistry 38, 44194430.

845 Fathizadeh, M., Aroujalian, A., Raisi, A., 2011. Effect of added NaX nano-zeolite into polyamide as a

846 top thin layer of membrane on water flux and salt rejection in a reverse osmosis process.

PT
847 Journal of Membrane Science 375, 8895.

848 Fathizadeh, M., Aroujalian, A., Raisi, A., Fotouhi, M., 2013. Preparation and characterization of thin

RI
849 film nanocomposite membrane for pervaporative dehydration of aqueous alcohol solutions.

SC
850 Desalination 314, 2027.

851 Faure, B., Salazar-Alvarez, G., Ahniyaz, A., Villaluenga, I., Berriozabal, G., De Miguel, Y.R., Bergstrm,

U
852 L., 2013. Dispersion and surface functionalization of oxide nanoparticles for transparent
AN
853 photocatalytic and UV-protecting coatings and sunscreens. Science and Technology of

854 Advanced Materials 14, 123.


M

855 Fu, Q., Wong, E.H.H., Kim, J., Scofield, J.M.P., Gurr, P. a., Kentish, S.E., Qiao, G.G., 2014. The effect of
D

856 soft nanoparticles morphologies on thin film composite membrane performance. Journal of
TE

857 Material Chemistry A 2, 1775117756.

858 Geens, J., De Witte, B., Van der Bruggen, B., 2007. Removal of APIs (active pharmaceutical
EP

859 ingredients) from organic solvents by nanofiltration. Separation Science and Technology 42,

860 24352449.
C

861 Ghanbari, M., Emadzadeh, D., Lau, W.J., Matsuura, T., Ismail, A.F., 2015a. Synthesis and
AC

862 characterization of novel thin film nanocomposite reverse osmosis membranes with improved

863 organic fouling properties for water desalination, RSC Advances 5, 2126821276.

864 Ghanbari, M., Emadzadeh, D., Lau, W.J., Lai, S.O., Matsuura, T., Ismail, A.F., 2015b. Synthesis and

865 characterization of novel thin film nanocomposite (TFN) membranes embedded with halloysite

866 nanotubes (HNTs) for water desalination, Desalination 358, 3341.


38
ACCEPTED MANUSCRIPT
867 Goh, P.S., Ismail, A.F., Ng, B.C., 2014. Directional alignment of carbon nanotubes in polymer

868 matrices: Contemporary approaches and future advances. Composites Part A: Applied Science

869 and Manufacturing 56, 103126.

870 He, J. , Matsuura, T., Paul Chen, J., 2014. A novel Zr-based nanoparticle-embedded PSF blend hollow

871 fiber membrane for treatment of arsenate contaminated water: Material development,

PT
872 adsorption and filtration studies, and characterization. Journal of Membrane Science 452, 433

873 445.

RI
874 Hu, D., Xu, Z.-L., Chen, C., 2012. Polypiperazine-amide nanofiltration membrane containing silica

SC
875 nanoparticles prepared by interfacial polymerization. Desalination 301, 7581.

876 Huang, H., Qu, X., Dong, H., Zhang, L., Chen, H., 2013a. Role of NaA zeolites in the interfacial

U
877 polymerization process towards a polyamide nanocomposite reverse osmosis membrane. RSC
AN
878 Advances 3, 82038207.

879 Huang, H., Qu, X., Ji, X., Gao, X., Zhang, L., Chen, H., Hou, L., 2013b. Acid and multivalent ion
M

880 resistance of thin film nanocomposite RO membranes loaded with silicalite-1 nanozeolites.
D

881 Journal of Materials Chemistry A 1, 1134311349.


TE

882 Jadav, G.L., Singh, P.S., 2009. Synthesis of novel silica-polyamide nanocomposite membrane with

883 enhanced properties. Journal of Membrane Science 328, 257267.


EP

884 Jeong, B.-H., Hoek, Eric M.V., Yan, Y., Subramani, A., Huang, X., Hurwitz, G., Ghosh, A.K., Jawor, A.,

885 2007. Interfacial polymerization of thin film nanocomposites: A new concept for reverse
C

886 osmosis membranes. Journal of Membrane Science 294, 17.


AC

887 Jeong, B.-H., Subramani, A., Yan, Y., Hoek, E.M.V., 2005. Antifouling thin film nanocomposite (Tfnc)

888 membranes for desalination and water reclamation, in: 2005 AIChE Annual Meeting and Fall

889 Showcase; Cincinnati, OH; United States. 30 October 20054 November 2005; Code 6692.

890 Jimenez Solomon, M.F., Bhole, Y., Livingston, Andrew Guy, 2012. High flux membranes for organic

891 solvent nanofiltration (OSN)Interfacial polymerization with solvent activation. Journal of

892 Membrane Science 423-424, 371382.


39
ACCEPTED MANUSCRIPT
893 Kim, E., Hwang, G., El-din, M.G., Liu, Y., 2012. Development of nanosilver and multi-walled carbon

894 nanotubes thin-film nanocomposite membrane for enhanced water treatment. Journal of

895 Membrane Science 394-395, 3748.

896 Kim, E.-S., Deng, B., 2011. Fabrication of polyamide thin-film nano-composite (PA-TFN) membrane

897 with hydrophilized ordered mesoporous carbon (H-OMC) for water purifications. Journal of

PT
898 Membrane Science 375, 4654.

899 Kim, S., Fornasiero, F., Park, H.G., In, J. Bin, Meshot, E., Giraldo, G., Stadermann, M., Fireman, M.,

RI
900 Shan, J., Grigoropoulos, C.P., Bakajin, O., 2014. Fabrication of flexible, aligned carbon

SC
901 nanotube/polymer composite membranes by in-situ polymerization. Journal of Membrane

902 Science 460, 9198.

U
903 Kim, S.G., Chun, J.H., Chun, B.-H., Kim, S.H., 2013. Preparation, characterization and performance of
AN
904 poly(aylene ether sulfone)/modified silica nanocomposite reverse osmosis membrane for

905 seawater desalination. Desalination 325, 7683.


M

906 Kong, C., Koushima, A., Kamada, T., Shintani, T., Kanezashi, M., Yoshioka, T., Tsuru, T., 2011.
D

907 Enhanced performance of inorganic-polyamide nanocomposite membranes prepared by metal-


TE

908 alkoxide-assisted interfacial polymerization. Journal of Membrane Science 366, 382388.

909 Kong, C., Shintani, T., Tsuru, T., 2010. Pre-seeding-assisted synthesis of a high performance
EP

910 polyamide-zeolite nanocomposite membrane for water purification. New Journal of Chemistry

911 34, 21012104.


C

912 Kosaraju, P.B., Sirkar, K.K., 2008. Interfacially polymerized thin film composite membranes on
AC

913 microporous polypropylene supports for solvent-resistant nanofiltration. Journal of Membrane

914 Science 321, 155161.

915 Kurth, C.J., Burk, R., Green, J., 2011. Utilizing nanotechnology to enhance RO membrane

916 performance for seawater desalination, in: IDA World Congress Perth Convention and

917 Exhibition Centre (PCEC), Perth, Western Australia.49 September, 2011.


40
ACCEPTED MANUSCRIPT
918 Lau, W.J., Ismail, A.F., Goh, P.S., Hilal, N., Ooi, B.S., 2014. Characterization methods of thin film

919 composite nanofiltration membranes. Separation & Purification Reviews 44, 135156.

920 Lau, W.-J., Ismail, A.F., 2009. Polymeric nanofiltration membranes for textile dye wastewater

921 treatment: Preparation, performance evaluation, transport modelling, and fouling control a

922 review. Desalination 245, 321348.

PT
923 Lau, W.J., Ismail, A.F., Misdan, N., Kassim, M.A., 2012. A recent progress in thin film composite

924 membrane: A review. Desalination 287, 190199.

RI
925 Lee, H.S., Im, S.J., Kim, J.H., Kim, H.J., Kim, J.P., Min, B.R., 2008. Polyamide thin-film nanofiltration

SC
926 membranes containing TiO2 nanoparticles. Desalination 219, 4856.

927 Lee, S.Y., Kim, H.J., Patel, R., Im, S.J., Kim, J.H., Min, B.R., 2007. Silver nanoparticles immobilized on

U
928 thin film composite polyamide membrane: characterization, nanofiltration, antifouling
AN
929 properties. Polymers for Advanced Technologies 18, 562568.

930 LG NanoH2O Inc. http://www.nanoh2o.com/ (assessed on 5 Jan 2015)


M

931 Li, X., Wang, R., Wicaksana, F., Tang, C.Y., Torres, J., Fane, A.G., 2014. Preparation of high
D

932 performance nanofiltration (NF) membranes incorporated with aquaporin Z. Journal of


TE

933 Membrane Science 450, 181188.

934 Lind, M.L., Ghosh, A.K., Jawor, A., Huang, X., Hou, W., Yang, Yang, Hoek, E.M. V, 2009. Influence of
EP

935 zeolite crystal size on zeolite-polyamide thin film nanocomposite membranes. Langmuir 25,

936 1013910145.
C

937 Ma, N., Wei, J., Liao, R., Tang, C.Y., 2012. Zeolite-polyamide thin film nanocomposite membranes:
AC

938 Towards enhanced performance for forward osmosis. Journal of Membrane Science 405-406,

939 149157.

940 Ma, N., Wei, J., Qi, S., Zhao, Y., Gao, Y., Tang, C.Y., 2013. Nanocomposite substrates for controlling

941 internal concentration polarization in forward osmosis membranes. Journal of Membrane

942 Science 441, 5462.


41
ACCEPTED MANUSCRIPT
943 Misdan, N., Lau, W.J., Ismail, A.F., Matsuura, T., 2013. Formation of thin film composite

944 nanofiltration membrane: Effect of polysulfone substrate characteristics. Desalination 329, 9

945 18.

946 Misdan, N., Lau, W.J., Ismail, A.F., 2012. Seawater Reverse Osmosis (SWRO) desalination by thin-film

947 composite membraneCurrent development, challenges and future prospects. Desalination

PT
948 287, 228237.

949 Misdan, N., Lau, W.J., Ismail, A.F., Matsuura, T., Rana, D., 2014. Study on the thin film composite

RI
950 poly(piperazine-amide) nanofiltration membrane: Impacts of physicochemical properties of

SC
951 substrate on interfacial polymerization formation. Desalination 344, 198205.

952 Misdan, N., Lau, W.J., Ong, C.S., Ismail, A.F., Matsuura, T., 2015. Study on the thin film composite

U
953 poly(piperazine-amide) nanofiltration membranes made of different polymeric substrates:
AN
954 Effect of operationg conditions. Korean Journal of Chemical Engineeiring 32 (4), 753760.

955 Namvar-Mahboub, M., Pakizeh, M., Davari, S., 2014. Preparation and characterization of UZM-
M

956 5/polyamide thin film nanocomposite membrane for dewaxing solvent recovery. Journal of
D

957 Membrane Science 459, 2232.


TE

958 Niksefat, N., Jahanshahi, M., Rahimpour, A., 2014. The effect of SiO2 nanoparticles on morphology

959 and performance of thin film composite membranes for forward osmosis application.
EP

960 Desalination 343, 140146.

961 Ong, C.S., Lau, W.J., Ismail, A.F. 2012. Treatment of dyeing solution by NF membrane for
C

962 decolorization and salt reduction. Desalination and Water Treatment 50, 245253.
AC

963 Park, J., Choi, W., Kim, S.H., Chun, B.H., Bang, J., Lee, K.B., 2010. Enhancement of chlorine resistance

964 in carbon nanotube-based nanocomposite reverse osmosis membrane. Desalination and Water

965 Treatment, 15, 198204.

966 Peeters, J.M.M., Boom, J.P., Mulder, M.H.V., Strathmann, H., 1998. Retention measurements of

967 nanofiltration membranes with electrolyte solutions. Journal of Membrane Science 145, 199

968 209.
42
ACCEPTED MANUSCRIPT
969 Pendergast, M.M., Ghosh, A.K., Hoek, E.M.V., 2013. Separation performance and interfacial

970 properties of nanocomposite reverse osmosis membranes. Desalination 308, 180185.

971 Peyravi, M., Jahanshahi, M., Rahimpour, A., Javadi, A., Hajavi, S., 2014. Novel thin film

972 nanocomposite membranes incorporated with functionalized TiO2 nanoparticles for organic

973 solvent nanofiltration. Chemical Engineering Journal 241, 155166.

PT
974 Rajaeian, B., Rahimpour, A., Tade, M.O., Liu, S., 2013. Fabrication and characterization of polyamide

975 thin film nanocomposite ( TFN ) nanofiltration membrane impregnated with TiO2 nanoparticles.

RI
976 Desalination 313, 176188.

SC
977 Sani, N.A.A., Lau, W.J., Ismail, A.F., 2015. Polyphenylsulfone-based solvent resistant nanofiltration

978 (SRNF) membrane incorporated with copper-1,3,5-benzenetricarboxylate (Cu-BTC)

U
979 nanoparticles for methanol separation. RSC Advances 5, 1300013010
AN
980 Shen, J.N., Yu, C.C., Ruan, H.M., Gao, C.J., Van der Bruggen, B., 2013. Preparation and

981 characterization of thin-film nanocomposite membranes embedded with poly(methyl


M

982 methacrylate) hydrophobic modified multiwalled carbon nanotubes by interfacial


D

983 polymerization. Journal of Membrane Science 442, 1826.


TE

984 Sorribas, S., Gorgojo, P., Carlos, T., Joaquin, C., Livingston, Andrew G., 2013. High flux thin film

985 nanocomposite membranes based on metalorganic frameworks for organic solvent


EP

986 nanofiltration. Journal of The American Chemical Society 135, 1520115208.

987 Subramani, A., Jacangelo, J.G., 2015. Emerging desalination technologies for water treatment: A
C

988 critical review. Water Research 75, 164187.


AC

989 Tang, C.Y., Zhao, Y., Wang, R., Helix-Nielsen, C., Fane, A.G., 2013. Desalination by biomimetic

990 aquaporin membranes: Review of status and prospects. Desalination 308, 3440.

991 Verssimo, S., Peinemann, K.-V., Bordado, J., 2005. New composite hollow fiber membrane for

992 nanofiltration. Desalination 184, 111.


43
ACCEPTED MANUSCRIPT
993 Wu, B., Li, X., An, D., Zhao, S., Wang, Y., 2014. Electro-casting aligned MWCNTs/polystyrene

994 composite membranes for enhanced gas separation performance. Journal of Membrane

995 Science 462, 6268.

996 Wu, H., Tang, B., Wu, P., 2010. MWNTs/Polyester Thin film nanocomposite membrane: An approach

997 to overcome the trade-off effect between permeability and selectivity. The Journal of Physical

PT
998 Chemistry C 114, 1639516400.

999 Wu, H., Tang, B., Wu, P., 2013. Optimizing polyamide thin film composite membrane covalently

RI
1000 bonded with modified mesoporous silica nanoparticles. Journal of Membrane Science 428,

SC
1001 341348.

1002 Xie, Z., Hoang, M., Ng, D., Doherty, C., Hill, A., Gray, S., 2014. Effect of heat treatment on

U
1003 pervaporation separation of aqueous salt solution using hybrid PVA/MA/TEOS membrane.
AN
1004 Separation and Purification Technology 127, 1017.

1005 Xu, G.-R., Wang, J.-N., Li, C.-L., 2013. Strategies for improving the performance of the polyamide thin
M

1006 film composite (PA-TFC) reverse osmosis (RO) membranes: Surface modifications and
D

1007 nanoparticles incorporation. Desalination 328. 83100.


TE

1008 Yin, J., Deng, B. 2014. Polymer-matrix nanocomposite membranes for water treatment, Journal of

1009 Membrane Science 479, 256275.


EP

1010 Yin, J., Kim, E.-S., Yang, John, Deng, B., 2012. Fabrication of a novel thin-film nanocomposite (TFN)

1011 membrane containing MCM-41 silica nanoparticles (NPs) for water purification. Journal of
C

1012 Membrane Science 423-424, 238246.


AC

1013 Yin, J., Yang, Yu, Hu, Z., Deng, B., 2013. Attachment of silver nanoparticles (AgNPs) onto thin-film

1014 composite (TFC) membranes through covalent bonding to reduce membrane biofouling.

1015 Journal of Membrane Science 441, 7382.

1016 Zhao, H., Qiu, S., Wu, L., Zhang, L., Chen, H., Gao, C., 2014. Improving the performance of polyamide

1017 reverse osmosis membrane by incorporation of modified multi-walled carbon nanotubes,

1018 Journal of Membrane Science 450, 249256.


44
ACCEPTED MANUSCRIPT
1019 Zhao, Y., Qiu, C., Li, X., Vararattanavech, A., Shen, W., Torres, J., Helix-Nielsen, C., Wang, R., Hu, X.,

1020 Fane, A.G., Tang, C.Y., 2012. Synthesis of robust and high-performance aquaporin-based

1021 biomimetic membranes by interfacial polymerization-membrane preparation and RO

1022 performance evaluation. Journal of Membrane Science 423424, 422428.

1023 Zhu, B., Myat, D.T., Shin, J.-W., Na, Y.-H., Moon, I.-S., Connor, G., Maeda, S., Morris, G., Gray, S.,

PT
1024 Duke, M., 2015. Application of robust MFI-type zeolite membrane for desalination of saline

1025 wastewater. Journal of Membrane Science 475, 167174.

RI
1026 Zou, H., Jin, Y., Yang, Jun, Dai, H., Yu, X., Xu, J., 2010. Synthesis and characterization of thin film

SC
1027 composite reverse osmosis membranes via novel interfacial polymerization approach.

1028 Separation and Purification Technology 72, 256262.

U
1029
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Figure and Caption

PT
RI
U SC
AN
M
D

Figure 1. Number of relevant research documents published in refereed journals between 2007
TE

and 2015 (Source: Scopus, Date of access: 31 March 2015)


C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D

Figure 2. Schematic diagram of immobilization of silver (Ag) nanoparticles onto the top PA
surface of TFC membrane (Yin et al., 2013).
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M

Figure 3. Schematic preparation process of modified mesoporous silica nanoparticles (mMSN)

(Wu et al., 2013)


D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M

20 nm
D

(a) (b)
TE

Figure 4. (a) Water suspensions of H-OMCs after three days of settling following mixing with

water and (b) TEM images of OMCs samples (scale bar: 20 nm) (Kim and Deng, 2011)
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D

Figure 5. Schematic of chemical interaction between functionalized UZM-5 nanoparticles and


TE

TMC monomers during interfacial polymerization (Namvar-Mahboub et al., 2014)


C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
(a) (b)

Figure 6. Comparison between TFN and typical TFC membrane, (a) effect of nanocomposite
M

substrate on organic fouling of composite membrane in PRO mode (test conditions: feed
D

solution: 10 mM NaCl with 200 mg/L bovine serum albumin, draw solution: 2.0 M NaCl, cross-
TE

flow velocity: 32.72 cm/s on both sides of the FO membrane and temperature: 25 C) and (b)

profile of effective osmotic pressure difference across TFN and TFC membrane (with and
EP

without fouling) in PRO mode (Emadzadeh et al., 2014a)


C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
(a) (b)
M

Figure 7.TEM images of the cross-section of (a) TFN membrane with silica nanoparticles

(MCM-41) embedded (Yin et al., 2012) and (b) TFN membrane with NaA zeolite embedded
D

(Huang et al., 2013a).


TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
Toluene

U SC
Water

AN
M

Figure 8.Direct observation on the dispersion rate of PMMAMWNTs in toluenewater after 1

month (Shen et al., 2013).


D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
(b)

RI
SC
(c)

U
AN
M

(a)
D

Figure 9. (a) Schematic representation of the TFN membrane fabrication via pre-seeding
TE

method and FESEM images of PSF support surface before (b) and after (c) pre-seeding

process (Kong et al., 2010)


C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M

Figure 10. Establishing PA layer on top of microporous substrate through static and dynamic

interfacial polymerization (An et al., 2012).


D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
(b)

(a)

RI
U SC
AN
Figure 11. (a) Schematic of an apparatus for electrical alignment of MWCNTs in polymer matrix
M

(b) Micrographs of 0.3 wt% MWCNTs dispersed in PS membrane with and without applying
D

electric field (Wu et al., 2014).


TE
C EP
AC
ACCEPTED MANUSCRIPT
Highlights

Comprehensive review on the development of TFN membrane for water treatment

Potential of TFN membrane in overcoming permeability/selectivity trade-off effect

Nanofiller characteristics/amount and IP condition that affect TFN properties

PT
Innovative approaches to overcome TFN fabrication challenges

RI
U SC
AN
M
D
TE
C EP
AC

You might also like