You are on page 1of 10

Desalination 386 (2016) 6776

Contents lists available at ScienceDirect

Desalination
journal homepage: www.elsevier.com/locate/desal

Thin lm composite membranes embedded with graphene oxide for


water desalination
Mohamed E.A. Ali a,c, Leyi Wang b, Xinyan Wang b, Xianshe Feng a,
a
b
c

Department of Chemical Engineering, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada
Zhaojin Motian Co., Ltd., Zhaoyuan, Shandong, China
Water Desalination & Treatment unit, Hydrogeochemistry Dept., Desert Research Center, Cairo, Egypt

H I G H L I G H T S

Thin lm composite membranes embedded with graphene oxide were prepared.


Membranes exhibited improved water ux, mechanical strength.
Membranes were stable in acidic and alkaline solutions.
The presence of graphene oxide increased chlorine and fouling resistances.

a r t i c l e

i n f o

Article history:
Received 29 September 2015
Received in revised form 24 February 2016
Accepted 25 February 2016
Available online xxxx
Keywords:
Membranes
Thin lm composite
Graphene oxide
Chlorine tolerance
Fouling resistance

a b s t r a c t
This work deals with thin lm composite membranes prepared from m-phenylenediamine and 1,3,5benzenetricarbonyl chloride by interfacial polymerization on the surface of a polysulfone substrate, and graphene
oxide was embedded into the membrane during membrane formation to improve the membrane performance.
The desalination performance of the membranes was evaluated in terms of water ux and salt rejection, along
with a baseline membrane containing no graphene oxide. The membrane morphology and surface properties
were also studied using contact angle measurements, FT-IR, XRD and SEM. Incorporating a small amount of
graphene oxide into the membrane was shown to improve the water ux, mechanical stability, and chlorine
and fouling resistances of the membrane. At 15 bar, a water ux of 29.6 L/m2h and a salt rejection of 97%
were obtained for a saline solution (2000 ppm of NaCl) when the amine reactant contained 100 ppm of graphene
oxide during membrane fabrication. The membranes were found to be stable in acidic and alkaline solutions.
2016 Elsevier B.V. All rights reserved.

1. Introduction
The shortage of clean and fresh water has become one of the critical
problems for sustainable development to meet the growing social, economic and environmental needs of the society [1]. The desalination of
brackish and sea water based on thermal processes and membrane
technologies holds great promises to address these issues [2,3]. Especially, the development of thin-lm-composite (TFC) membranes comprising of a substrate and an interfacially polymerized polyamide (PA)
skin layer has signicantly advanced the membrane technology for
water desalination [4]. In the thin lm composite membranes, the substrate provides mechanical strength to the membrane against the operating pressure applied across the membrane, whereas the PA active

Corresponding author. Tel.: +1 519 888 4567.


E-mail address: xfeng@uwaterloo.ca (X. Feng).

http://dx.doi.org/10.1016/j.desal.2016.02.034
0011-9164/ 2016 Elsevier B.V. All rights reserved.

layer is responsible for rejecting salt while allowing water to pass. The
PA skin layer also determines the membrane resistances to fouling
and chlorine [5]. Ideally, the membranes should be chemically and mechanically stable over a long period of operation at high pressures, while
maintaining their desired water ux and salt rejection characteristics.
Unfortunately, the current generation of membranes faces two main
challenges: chlorine sensitivity and fouling propensity [6]. The amide
groups in PA skin layer are vulnerable to chlorine attack even at a low
chlorine dosage in the feed water [7], and membrane chlorination normally leads to reduced salt rejection that compromises the quality of
the permeated water. In addition, the surface fouling of the thin lm
composite membranes is often a serious problem because frequent
cleaning will not only increase the operating cost but the service
life of the membrane will also be shortened if harsh cleaning agents
are needed [8].
Therefore, many efforts have been made to modify the membrane
surface in order to improve water ux, salt rejection, and fouling and

68

M.E.A. Ali et al. / Desalination 386 (2016) 6776

chlorine resistances of the membrane. One common approach is the


development of hybrid organicinorganic membranes with tailored
surface properties. Over the last two decades, various hydrophilic
nanomaterials (including alumina, silica, titanium dioxide, zeolites,
and carbon nanotubes) have been used to improve the performance of
TFC membranes [9]. In general, a small amount of such nanollers is
often insufcient to alter the membrane permselectivity signicantly,
while the uniform dispersion of nanollers in the membrane will be affected because of nanoparticle aggregation if too much nanollers are
used. Nanoparticles with unique structures and functional groups
(e.g., hydrophilicized graphene oxide nanosheets) that enhance their
compatibilities with the polymer matrix of the membrane are of particular interest. Because of the oxygen-containing functional groups in
graphene oxide (GO) nanosheets (e.g., hydroxyl and epoxy groups on
the basal plane and carboxyl groups at the edge), they generally have a
better dispersibility in water or polar solvents than other nanoparticles
[10,11]. It may be mentioned that a great deal of work has been done
on the synthesis of various polymer-carbon based hybrid materials
through different approaches to enhance mechanical, electrical, catalytic
or other properties (see, for example [1221]). GO nanosheets have recently attracted signicant attention in membrane development because
of their unique nanostructures and physical and mechanical properties
[2224]. Because of the different functional groups (e.g., hydroxyl, carboxyl, and epoxide) in graphene oxide, there is a good compatibility between the nanosheets with the host (polymer) materials through
covalent or non-covalent attachments [25]. When graphene oxide
nanosheets are embedded into a membrane matrix, the surface hydrophilicity of the membrane will be enhanced, which are helpful
to enhance water permeability and fouling resistance. In addition,
because of the intermolecular hydrogen bonding between the
amide groups of PA and the functional groups of graphene oxide,
the active amide groups in PA vulnerable to chlorine attack are
shielded by the nanosheets, resulting in improved membrane resistance to chlorine [26]. Moreover, the embedded GO nanosheets in
the membrane matrix also increases the mechanical strength and
enhances the membrane stability against high transmembrane pressures [27].
The main objective of the present work was to prepare hybrid
organicinorganic thin lm composite membranes by incorporating
graphene oxide nanosheets into the interfacially polymerized polyamide
skin layer. The effects of the addition of GO nanollers in the membrane
on the desalination performance of the resulting membrane were investigated. It was shown that by properly controlling the membrane formation
conditions, thin lm composite membranes with signicantly improved
performance (i.e., mechanical stability against high pressure, tolerance
for high chlorine dosages, resistance to strong acidic and alkaline solutions) were produced. These membranes compared very favorably
with those GO-modied membranes recently reported in the literature
[26,2830]. It may be mentioned that unlike GO-modied membranes
where GO was incorporated into the membrane by coating or layer-bylayer deposition [26,30], the composite membranes developed in this
work comprised of GO that was embedded in the polyamide skin layer
during the interfacial polymerization.
It may be noted that prior work on modication of TFC membranes with GO involved surface coating or layer-by-layer assembly.
In one approach, GO was deposited on top of a polyamide active layer
by electrostatic self assembly, and the GO coating layer led to an
additional resistance to water permeability [14,16]. In another approach, oppositely charged GOs were deposited on a porous substrate by electrostatic self assembly, and the polyamide layer was
formed on top of the GO layer by conventional interfacial polymerization [30]. In both cases, the GO layer and the polyamide layer
were not interpenetrated, and the stability of the electrostatically assembled GO layer may be a potential concern when used in water desalination because of the ionic solutes involved. Therefore, it is
desirable to incorporate GO into the polyamide layer during the

course of interfacial polymerization so as to form a nanostructured


hybrid membrane. In the present work, GO nanollers were dispersed in the aqueous amine reactant phase, which was allowed to
react with the organic phase of acyl chloride reactant on a substrate
surface to achieve interfacial polymerization, thereby incorporating
the GO in situ into the thin polyamide layer during the course of interfacial polymerization.

2. Experimental
2.1. Materials
Graphene oxide was prepared from graphite powder, potassium
permanganate, sulfuric acid, nitric acid and hydrogen peroxide; all the
chemicals were supplied by Fisher Scientic, except for graphite powder which was supplied by Acros Organics. Microporous polysulfone
ultraltration membranes supplied by Sepro Membranes were used as
the substrate. They had a molecular weight cut-off of 10,000, and a
pure water permeability of approximately 90 L/m2hbar. 1,3,5benzenetricarbonyl chloride (i.e., trimesoyl chloride, TMC) (N98%), mphenylenediamine (MPD) (N 99%), and camphor sulphonic acid (CSA)
were supplied by Fischer Scientic. n-Hexane was purchased from Caledon Laboratories. Sodium lauryl sulfate (SLS) was purchased from
Matheson Coleman & Bell Chemical. NaCl (EMD Chemical) was used
to characterize the salt rejection of the TFC membranes. The chlorine solution was prepared from a commercially available sodium hypochlorite
solution (NaClO, 14.5% available chlorine, Alfa Aesar). When needed, the
feed solution pH was adjusted to desired values using hydrochloric acid
(37 wt%, Sigma-Aldrich) or sodium hydroxide (Caledon Laboratories).
Bovine serum albumin (BSA) supplied by Sigma-Aldrich was used as a
model foulant in membrane fouling experiments.

2.2. Preparation of graphene oxide


GO was prepared following a modied Hummers' method [31].
Briey, in an ice bath, 1 g of graphite powder was dispersed in a mixture
of 25 ml of cold concentrated sulfuric acid and 1 g of sodium nitrate.
Then, 3 g of potassium permanganate was slowly added under vigorous
stirring and cooling conditions to keep the temperature below 20 C.
Thereafter, the reaction mixture was placed in a water bath at 35 C
with continuous stirring for 1 h. Then, 50 ml of deionized water was
slowly added, and the reaction mixture was allowed to stay at 98 C
for 12 h, producing a bright-yellow suspension. Finally, the reaction
was terminated by sequentially adding 140 ml of deionized water and
3 ml of hydrogen peroxide solution (30%). The solid product was ltered
out of the solution, rinsed with a dilute hydrochloric acid (3.4 wt.%)
until a pH of 7, and vacuum-dried at room temperature.

Fig. 1. Schematic diagram of the cross-ow membrane testing unit.

M.E.A. Ali et al. / Desalination 386 (2016) 6776

69

Fig. 2. Schematic representation of chemical interactions between MPD and TMC with and without GO relevant to membrane formation by interfacial polymerization. (For interpretation
of the references to color in this gure legend, the reader is referred to the web version of this article.)

2.3. Preparation of thin lm composite membranes


The thin lm composite membranes with and without graphene
oxide were prepared on the top surface of the polysulfone substrate

using the interfacial polymerization technique. In brief, the substrate


membrane was rinsed with deionized water for at least 2 h to remove
any preservatives. After that, 10 ml of the aqueous solution of MPD
(2 wt.%) containing additives (camphor sulphonic acid 1 wt.%, and

70

M.E.A. Ali et al. / Desalination 386 (2016) 6776

sodium lauryl sulfate 0.2 wt.%) was allowed to sit on the top surface of
the substrate for 1 min. These additives were used to enhance MPD
sorption onto the substrate. After draining off any droplets of the
amine solution from the substrate surface, 10 ml of the TMC in hexane solution (0.1 wt.%) was brought into contact with substrate surface for 50 s to induce interfacial polymerization between TMC and
the MPD deposited. The excess organic solution on the membrane
surface was removed, and the resulting thin lm composite membrane was subjected to heat treatment at 65 C for 5 min, thereby
completing the interfacial polymerization to attain the desired stability of the TFC membrane against high pressure in desalination
processes. The membranes were stored in water after thorough
washing with water. To prepare TFC membranes incorporated
with graphene oxide, the same membrane fabrication procedure
was followed, except that the aqueous reactant solution of MPD
contained different amounts of graphene oxide (0300 ppm). The
MPD/GO solutions were sonicated in an ultrasonication bath for
30 min before being deposited on the substrate surface for interfacial polymerization with TMC. For convenience, thin lm composite
membranes with and without GO are designated as TFC and TFC/GO
membranes, respectively.
2.4. Membrane characterization
The membranes were characterized with infrared spectroscopy
using a Bruker Vector 22 FT-IR spectrophotometer. The X- ray diffraction patterns of the membranes were recorded using a Bruker-AXS D8
Discover diffractometer (Co-K source). The surface morphology of
the membranes was characterized using a scanning electron microscopy (SEM, Hitachi S-4800). The roughness and thickness values were
estimated by using the image-J software program. The surface hydrophilicity of the membranes was evaluated using a contact angle analyzer
(Cam-plus Micro, Tantec Inc.). The membrane samples were air-dried at
ambient temperature for measuring contact angles of deionized water
on the membrane surface using the sessile droplet technique, the contact angle reported was an average of at least ve measurements on different locations of a single membrane sample.

Fig. 3. FT-IR spectra of graphene oxide (GO), and the TFC and TFC/GO membranes.

conductivity meter. The desalination experiments were conducted


in triplicates for every membrane sample, and the average values
were reported.

2.5. Membrane performance for water desalination


The performance of the membranes for water desalination was
measured in term of water ux and salt rejection using a crossow permeation apparatus (Fig. 1). It consisted of six membrane
cells, and each cell had an effective membrane area of 14.75 cm2.
This allowed six membranes to be measured under identical conditions. Unless stated otherwise, the operating conditions were: pressure 15 bar gauge, NaCl concentration in feed water 2000 ppm,
temperature 22 C, and pH 7. During the reverse osmosis runs, the
retentate solution was recirculated to the feed tank, and the permeate samples from the membrane cells were also recycled back to the
system upon measurements of their solute concentrations individually. This way, an essentially constant saline concentration in the
feed was maintained during the experiment. The salt rejection (R)
and permeation ux (J) were determined using the following two
equations,
R



CP
 100%
1
CF

V
At

where CF and CP are the salt concentrations in the feed and permeate, respectively. V is the volume of permeate collected over a time
interval t for a membrane area of A. The salt concentrations in the
permeate and feed solutions were determined using an Orion

Fig. 4. XRD patterns of GO, the TFC and TFC/GO membranes. The TFC/GO membrane was
prepared using 100 ppm of GO in the MPD solution.

M.E.A. Ali et al. / Desalination 386 (2016) 6776

71

Fig. 5. SEM images of the top surfaces and cross sections of TFC (a and d), TFC/GO with 100 ppm of GO in MPD solution (b and e), and TFC/GO with 300 ppm of GO in MPD solution (c and f).

2.6. Evaluation of membranes resistance to chlorine


To evaluate the chlorine resistance of the membranes, the water ux
and salt rejection were measured with a feed saline solution (2000 ppm
NaCl) containing 500 ppm of NaOCl at a pressure of 7 bar gauge for 10 h.
The variations in water ux and salt rejection were monitored as permeation proceeded with time.

2.7. Evaluation of membrane resistance to fouling


Membrane fouling was evaluated with ltration-cleaning cycles. The
feed saline solution contained 2000 ppm of NaCl, and BSA was used as a
model protein foulant. First, the membranes were preconditioned in the
membrane unit with a saline solution (2000 ppm NaCl) at a pressure of
15 bar gauge, and the initial water ux was determined when the water
ux reached a steady state. Then, the membranes were subjected to a
ltration test with a saline solution of 2000 ppm of NaCl (in the absence
of BSA) for 6 h, followed by a ltration test with a saline solution
(2000 ppm of NaCl) containing 100 ppm of BSA for another 6 h. After
that, the membranes were washed thoroughly with de-ionized water
for 2 h, followed by ltration tests with the two saline solutions again.

The ltration-cleaning cycles were repeated to determine the signicance of membrane fouling as reected by the water ux decline.
3. Results and discussion
3.1. Characterization of TFC and TFC/GO membranes
The TFC and TFC/GO composite membranes were prepared by the
interfacial polymerization technique. Fig. 2 is a schematic of interfacial
polymerization between MPD and TMC; the interactions between GO
and PA are also illustrated in the gure. GO can form a hydrogen bond
with terminal primary and secondary amines, as well as covalent
bond through condensation reactions with terminal carboxyl groups
of TMC in the linear portion of PA. During the experiment, we noticed
Table 1
Estimated skin layer thicknesses of the membranes.
Membrane

Thickness (m)

TFC
TFC/GO 100 ppm
TFC/GO 300 ppm

4.56 1.39
1.93 0.78
4.06 0.75

72

M.E.A. Ali et al. / Desalination 386 (2016) 6776

a color change of the MPD solution from transparent to dark green then
to black in a few hours at ambient conditions, presumably due to oxidation. On the other hand, when GO (yellowish brown) was added to the
amine solution, there was little change in color of the solution over a
long period of time. This may be due to blocking the reactive amine
sites in MPD through formation of hydrogen bonds with the functional
groups of GO.
Fig. 3 shows the FT-IR spectra of GO nanosheets, polyamide active
layer in the thin lm composite membranes with and without GO.
For the GO nanosheets, the bands at 3429 cm 1, 1734 cm 1 and
1052 cm1 correspond to hydroxyl, carboxyl and epoxide functional
groups, respectively, and these results conrm the preparation of GO
by the oxidation process of graphite [29,32]. For the thin lm composite
membranes, the characteristic band at 1694 cm1 was attributed to C=
O stretching vibration (amide II) of the polyamide active layer, while the
band at 1589 cm1 was due to NH and CN stretching vibrations of the
amide group (amide II). The latter band was shown to have a lower intensity and broader width for the TFC/GO membrane, accompanied
with a band shifting to 1660 cm1, suggesting the interaction between
PA and GO nanosheets. The characteristic band at 1747 cm1 in the FTIR spectrum of the TFC/GO membrane corresponded to the stretching
vibration of C=O ester groups formed between carboxylic or hydroxyl
groups of GO and the carboxylic groups of the PA active layer. The
band at around 1113 cm1 corresponding to C-O stretching in the TFC
membrane also showed a reduction in band intensity and a shifting to
1106 cm1 for the TFC/GO membrane. The band shifting and change
in band intensity, especially from 400 to 1600 cm1, were apparent in
the FT-IR spectra when GO was incorporated into the membrane due
to the interactions between GO with the polyamide active layer in the
membrane.
Fig. 4 shows the XRD diffractogram of GO and the TFC and TFC/GO
membranes. GO showed a sharp peak at a 2 of 10.09, with a dspacing of 0.88 nm. This indicates the presence of functional groups in
the GO basal structure, which is in agreement with prior work on GOcontaining composite materials [23,33]. The PA active layer in TFC membrane was characterized by the presence of three semi-crystalline peaks
at 2 of 18.14, 23.25 and 26.55, corresponding to d-spacing of 0.48, 0.38
and 0.34 nm, respectively. A similar trend in the XRD pattern was observed for the TFC/GO composite membrane, but with a decreased
peak intensity and d-spacing. It has been reported that a small amount
of GO in the membrane can change the d-spacing between the polymer
chains [33].
Fig. 5 shows SEM images of the surfaces and the cross-sections of the
thin lm composite membranes. A comparison of the three membranes
(that is, TFC, TFC/GO with 100 ppm of GO, and TFC/GO with 300 ppm of
GO) shows that all the membrane surfaces showed a ridge and valley
morphology. Similar observations can also be found in other studies
on thin lm composite membranes containing nanollers [29,34]. As
expected, the incorporation of different amounts of GO in the membrane resulted in different surface morphologies of the membranes.
Table 1 shows the thicknesses of the active skin layers of the thin lm
composite membranes estimated from the SEM images using ImageJ
processing software. The TFC/GO membranes tended to have rougher
surface than the TFC membrane containing no graphene oxide. As far
as the effective skin layer thickness is concerned, the TFC/GO (with
100 ppm of GO) was thinnest among the three membranes considered.
Compared to the GO-free TFC membrane, there is a hindrance effect of
GO nanosheets to the diffusivity of MPD from the aqueous phase to
reach the interface for reaction with TMC in the organic phase [24],
resulting in a thinner interfacially polymerized skin layer. However, excessive loading of GO will result in aggregation of the GO nanosheets,
which will cause an increase in the active layer thickness.
Fig. 6 shows the effects of incorporating GO in the thin lm composite membranes on the surface hydrophilicity of the membranes as
measured by the water contact angle on the membrane surface. Generally speaking, a high surface hydrophilicity is advantageous to the

Fig. 6. Water contact angle and pure water permeability of the TFC and TFC/GO
membranes as a function of GO concentration in the MPD solution used in membrane
preparation. Operating conditions for pure water permeability tests: pressure, 15 bar;
temperature 20 C; feed, deionized water.

membrane performance with respect to water permeability and fouling


resistance [35]. The pure water permeability is also shown in the gure.
With an increase in the concentration of GO in the reactive amine solution from 0 to 300 ppm, the contact angle decreased from 64 for the
GO-free TFC membrane to 48 for the TFC/GO membrane. Thus, there
was a considerable improvement in the membrane surface hydrophilicity, and this was expected because of the hydrophilic functional groups
(carboxyl, hydroxyl and epoxy) of GO. A hydrophilic membrane surface
facilitates the uptake of water molecules onto the membrane surface,
which enhances the water permeability. The data in Fig. 6 shows that
the pure water permeability (PWP) was increased when a small amount
of GO was present in the membrane. However, when the GO concentration in the amine reactant was sufciently high, a further increase in the

Fig. 7. Effects of graphene oxide concentration in MPD solution used in membrane


preparation on the RO performance of resulting composite membranes. Operating
conditions: pressure 15 bar, temperature 20 C, feed NaCl concentration 2000 ppm, and
pH 7.

M.E.A. Ali et al. / Desalination 386 (2016) 6776

73

Table 2
MPD/TMC based thin lm composite membranes containing graphene oxide llers.
Monomers of TFC Additives in MPD
membrane
solution

GO conc. and addition procedure

Water contact
Pressure NaCl conc. Water ux Salt
rejection (%) angle ()
Ref.
(bar)
(ppm)
(L/m2h)

76 ppm in MPD

15

2000

55

32,000

MPD, TMC

MPD, TMC

1000 ppm, deposited onto polysulfone substrate

TEA, DMSO, EHD, and CSA 2000 ppm in MPD


TEA
20,000 ppm, deposited onto TFC membrane
CSA, and SLS
100 ppm in MPD

MPD, TMC
MPD, TMC
MPD, TMC

15
15.5
15

2000
2000
2000

16.6

99

47

[24]

28
22
14
29.6

98

55.4

[26]

88
96
98

65
26
56

[29]
[30]
This work

TEA: triethyl amine.


DMSO: dimethyl sulfoxide.
EHD: 2-ethyl-1,3-hexane diol.
CSA: camphor sulfonic acid.
SLS: sodium lauryl sulfate.

GO content led to a decrease in the water permeability probably because aggregated GO nanosheets would increase the length and tortuosity of the passageways for the penetration of water molecules
through the membrane. Nonetheless, it was shown that over the GO
concentration range (0300 ppm) studied, all the TFC/GO membranes
had a pure water permeability higher than that of the GO-free TFC
membrane.
3.2. Desalination performance
The effects of GO concentrations in the reactive amine solution during membrane formation on the water ux and salt rejection of the
resulting membranes are shown in Fig. 7. At low GO concentrations,
the water ux of the membrane increased with an increase in the GO
concentration, whereas the salt rejection decreased. These trends
began to change at relatively high GO concentrations. There was a decrease in water ux and a slight increase in the salt rejection when the
GO concentration was above 150 ppm. Compared to the GO-free TFC
membrane which had a water ux of 21.4 L/m2.h and a salt rejection
of 98.5%, the TFC/GO membrane showed a 39% increase in water ux
and a slight reduction (1%) in salt rejection at a GO concentration of
100 ppm. The presence of GO in the membrane enhances water transport through the membrane matrix [36]. On the one hand, water molecules can transport through the channels of GO nanosheets, and on the
other hand, the hydrophilic functional groups (hydroxyl, carboxyl and
epoxy) in GO facilitate the adsorption of water molecules on the membrane surface. Both effects favor water transport. However, water ux is
also inuenced by the thickness of the skin layer [37]. On the one hand,
the presence of GO in the diamine reactant tends to result in a thinner
interfacially formed polyamide layer. On the other hand, the GO embedded in the skin layer also contributes to the skin layer thickness, which
increases the membrane resistance to water transport [38]. Because of
these two opposing effects, a maximum water ux was observed for
TFC/GO membrane at a GO concentration of 100 ppm among the membranes studied here. This membrane was used for further tests in subsequent studies of mechanical stability, chlorine resistance and antifouling
properties. In general, there was a slight decrease in the salt rejection
when GO was incorporated into the membrane, presumably due to
the nanosheet structure of the GO randomly positioned in the membrane skin layer. The GO nanosheets positioned perpendicularly to the
membrane surface are not effective to block the passage of salt while
allowing water to permeate. In addition, too much GO in the membrane
will aggregate, which will also affect the formation of the polyamide
skin layer by interfacial reaction between MPD and TMC because the
GO aggregates dispersed in the polyamide layer can lead to localized
defects in the aggregates, which tends to lower salt rejection.
It may be pointed out that GO-modication of MPD/TMC-based
polyamide TFC membranes has been done by coating or layer-by-layer
deposition of GO onto a substrate prior to polyamide layer formation
by interfacial polymerization [26], or simply onto the surface of a TFC
membrane after skin layer formation [30]. Not until earlier this year

was the concept of embedding GO into the polyamide layer during interfacial polymerization attempted [24,29]. Chae et al. [24] dispersed
GO in the aqueous solution of MPD for interracial polymerization,
while Bano et al. [29] used several additives (including triethyl amine,
dimethyl sulfoxide, 2-ethyl-1,3-hexane diol, and camphor sulfonic
acid) in the MPD solution. In this work, we chose to use camphor sulfonic acid and sodium lauryl sulfate as additives in the aqueous MPD solution in order to help adjust the solution pH and dispersity, respectively,
thereby facilitating deposition of MPD and GO onto the substrate surface for subsequent interfacial polymerization with TMC. The membranes so formed compared well with other GO-modied membranes
for water desalination, as illustrated by the water ux and salt rejection
shown in Table 2.
The operating pressure is an important parameter for water desalination by membranes. Fig. 8 shows the water ux and salt rejection of
TFC/GO membrane at different transmembrane pressures. To assess
the effect of addition of GO on mechanical stability of resulting membranes, the performance of the GO-free TFC membrane was plotted in
Fig. 8 as well. As expected, the water ux for both TFC and TFC/GO membranes increased as the operating pressure increased from 7 to 35 bar.
Unlike the TFC/GO membrane, which showed an almost constant salt
rejection, the TFC membrane experienced a sharp decline in the salt

Fig. 8. Salt rejection and water ux of TFC and TFC/GO membranes at different feed
pressures. The TFC/GO membrane was prepared using 100 ppm of GO in the MPD
solution. Other RO operating conditions were the same as in Fig. 7.

74

M.E.A. Ali et al. / Desalination 386 (2016) 6776

Fig. 9. Salt rejection and water ux of TFC and TFC/GO membranes at different feed
solution pH. The TFC/GO membrane was prepared using 100 ppm of GO in the MPD
solution. Other RO operating conditions same as in Fig. 7.

rejection when the operating pressure was above 25 bar. This was presumably caused by membrane deformation at high pressures. It has
been reported that the mechanical stress on the membrane may lead
to membrane deformation as a result of compaction, changes in the
pore size and shape, or rupture of the skin layer under extreme conditions. The data in Fig. 8 show that incorporating GO into the membrane
enhanced the mechanical stability of the membrane, which is understandable because of reinforcement of skin layer by the presence of intercalation GO nanosheets [39,40].
It is preferable to operate thin lm composite membranes with
aqueous solutions at pH ranges between 2 and 10 [41]. The effects of
GO on the membrane performance over a wide range of pH (2 to 12)
were tested, with the pH of the feed solution being adjusted with HCl
or NaOH accordingly. The results are shown in Fig. 9. Under alkaline solution conditions, an increase in the solution pH increased the water
ux and decreased the salt rejection for both TFC and TFC/GO membranes. This can be explained in light of de-protonation of carboxyl
groups of the linear portion of the polyamide in the membrane
skin layer, as well as the carboxylic groups of GO in case of TFC/GO
membranes [42,43], making the membrane surfaces more negatively
charged. Consequently, water sorption into the membrane skin was
enhanced, leading to an increased water ux and reduced salt rejection
because of the increased membrane swelling. On the other hand, compared to neutral feed solutions at pH of 7, the salt rejection was lower
for both membranes under acidic conditions, which was accompanied
with an increase in the water ux of the TFC membrane and a slight
decrease in the water ux of the TFC/GO membrane. Under acidic conditions, the terminal acidic and amine groups in the interfacially polymerized PA skin layer will be protonated, and this will cause the
membrane surface to be more positively charged. As a result, water
sorption in the membrane will the enhanced, leading to increased
membrane swelling, which tends to improve water permeability in
the membrane while lowering the salt rejection. When GO is embedded
in the membrane, the membrane swelling is constrained, and the
hydrophlic GO nanosheets also function as passageway to water permeation. Consequently, the water permeability did not change signicantly. On the other hand, GO is known to have a strong adsorption
properties with respect to salts in aqueous solutions [44], and this favors
salt rejection of the TFC/GO membrane. However, the ion sponge

Fig. 10. Effects of chlorine exposure time on salt rejection and water ux of the TFC and
TFC/GO membranes (The TFC/GO membrane was prepared using 100 ppm of GO in the
MPD solution). Chlorine dosage, 500 ppm. RO operating conditions: pressure 7 bar,
temperature 20 C, feed NaCl concentration 2000 ppm, and pH 7.

effect of GO appears to be inuenced by the solution pH. For instance,


Wang et al. [45] used a GO/amidoxime hydrogel to capture uranium
by adsorption, and the sorption capacity was shown to decrease significantly when the pH decreased in acidic solutions. If the salt sponge effect of GO to NaCl is affected by pH in a similar fashion, then the salt
rejection of TFC/GO membrane will decrease when the feed solution
pH changes from 7 to 2, as observed in Fig. 9. The study of Ganesh
et al. [46] on GO/polysulfone mixed matrix membrane also exhibited a
reduced salt rejection under acidic conditions when the solution pH decreased. A further study of NaCl sorption uptake in GO at different solution pH, which is beyond the scope of the present work, will be needed
to have a better understanding about the pH effect on salt rejection of
the TFC/GO membrane.
Fig. 10 shows the changes in water ux and salt rejection for both
TFC and TFC/GO membranes after exposure to 500 ppm of chlorine solution over different periods of time. For the sake of illustration, the
water ux and salt rejection were normalized by the performance
data of the membrane without chlorine exposure. As expected, there
was a signicant increase in water ux and a slight decrease in salt rejection after exposure of the membrane to chlorine. The TFC/GO membrane was found to be more stable against chlorine attack than the
GO-free TFC membrane. Chlorine is considered to degrade polyamide
macromolecules via a nucleophilic substitution reaction between chlorine and the hydrogen of the secondary amide group (NH) in polyamide [47]. The improved chlorine resistance of the TFC/GO membrane is
believed to result from hydrogen bonding between the secondary
amide groups in polyamide and the functional groups of GO nanosheets
[28] that functioned as a protection against chlorine.
3.3. Anti-fouling property of the membrane
Surface fouling is an important issue that affects the separation performance of membranes. Using BSA as a model foulant, the water uxes
of the TFC and TFC/GO membranes were investigated. Fig. 11 shows the
normalized water uxes (Jr) through the membranes relative to their
initial water uxes. A commercial membrane was also tested as a reference under the same conditions for comparison (the supplier of this

M.E.A. Ali et al. / Desalination 386 (2016) 6776

75

Acknowledgments
Research support from the Natural Sciences and Engineering
Research Council (NSERC) of Canada and Zhaojin Motian Co. is acknowledged. We are also grateful to the Science and Technology Development
Fund (STDF), Ministry of Scientic Research of Egypt, for awarding a fellowship to one of the authors (M.E.A.A.) to carry out the research at University of Waterloo.
References

Fig. 11. A comparison of ux recoveries for the TFC and TFC/GO membranes prepared in
this study (The GO content in MPD for preparing the TFC/GO membrane was 100 ppm).
A commercial TFC membrane was also tested as a reference. The membrane fouling
tests were carried out with feed solutions containing 2000 ppm of NaCl and 2000 ppm
of NaCl + 100 ppm of BSA, respectively, in the ltration cycle.

membrane is not disclosed per nondisclosure agreement). It was shown


that all membranes were fouled by BSA (100 ppm) when present in the
saline feed solution (2000 ppm of NaCl), as shown by the decreased
water ux over time. Although membrane cleaning with water could remediate the membranes to some extent, the membrane performance
could not be fully recovered by periodically cleaning the membranes
with deionized water. Compared to the TFC membrane, which was
fouled by BSA to a similar extent as the commercial TFC membrane,
the TFC/GO membrane was shown to be much more resistant to BSA
fouling. In addition, a water ux recovery as high as 85% was achieved
with the TFC/GO membrane after membrane cleaning, while the other
two membranes showed a water ux recovery of around 50%. The improvement on the anti-fouling property of the TFC/GO membrane is
considered to derive from the increased surface hydrophilicity of the
membrane and the more negatively charged surface when GO was embedded in the active membrane layer. As BSA is negatively charged at
pH 7, the membrane fouling is less signicant due to the favorable electrostatic interactions between the negatively charged membrane and
the foulant.

4. Conclusions
Thin lm composite membranes were prepared via the interfacial
polymerization technique, and graphene oxide was used as a modier
to improve the membrane properties. It was shown that incorporating
GO into the membrane improved the hydrophilicity, water permeability, chlorine resistance and antifouling properties of the membrane. The
membrane surface and morphology were also studied using contact
angle measurements, FT-IR, XRD and SEM. Compared to a GO-free TFC
membrane, a 39% increase in water ux was achieved when the
amine reactant contained 100 ppm of GO during membrane formation.
The thin lm composite membranes embedded with GO were shown to
be more resistant to chlorine attack and surface fouling than the membranes without GO. A higher water ux recovery was also achieved with
the TFC/GO membrane after membrane cleaning. At 15 bar, a water ux
of 29.6 L/m2h and a salt rejection of 97% were obtained for a saline solution (2000 ppm of NaCl) when the amine reactant contained 100 ppm
of graphene oxide during membrane fabrication. The membranes were
found to be stable in acidic and alkaline solutions.

[1] J.K. Kaldellis, E.M. Kondili, The water shortage problem in the Aegean archipelago
islands: cost-effective desalination prospects, Desalination 216 (2007) 123138.
[2] A. Subramani, J.G. Jacangelo, Emerging desalination technologies for water treatment: a critical review, Water Res. 75 (2015) 164187.
[3] E. Curcio, G.D. Proo, E. Fontananova, E. Drioli, Membrane technologies for seawater
desalination and brackish water treatment, in: A. Basile, A.C.K. Rastogi (Eds.), Advances in Membrane Technologies for Water Treatment, Woodhead Publishing, Oxford 2015, pp. 411441.
[4] J.E. Cadotte, R.J. Petersen, R.E. Larson, E.E. Erickson, A new thin-lm composite seawater reverse osmosis membrane, Desalination 32 (1980) 2531.
[5] A.F. Ismail, M. Padaki, N. Hilal, T. Matsuura, W.J. Lau, Thin lm composite membrane
recent development and future potential, Desalination 356 (2015) 140148.
[6] B. Mi, M. Elimelech, Silica scaling and scaling reversibility in forward osmosis, Desalination 312 (2013) 7581.
[7] D. Manish, P.R. Buch, P. Rao, J.J. Trivedi, A.V.R. Reddy, Chlorine stability of fully aromatic and mixed aromatic/aliphatic polyamide thin lm composite membranes, Int.
J. Nucl. Desalin. 3 (2008) 175185.
[8] G.-R. Xu, J.-N. Wang, C.-J. Li, Strategies for improving the performance of the
polyamide thin lm composite (PA-TFC) reverse osmosis (RO) membranes: surface
modications and nanoparticles incorporations, Desalination 328 (2013) 83100.
[9] H. Wu, B. Tang, P. Wu, Optimizing polyamide thin lm composite membrane covalently bonded with modied mesoporous silica nanoparticles, J. Membr. Sci. 428
(2013) 341348.
[10] L. Wu, L. Liu, B. Gao, R. Muoz-Carpena, M. Zhang, H. Chen, Z. Zhou, H. Wang, Aggregation kinetics of graphene oxides in aqueous solutions: experiments, mechanisms,
and modeling, Langmuir 29 (2013) 1517415181.
[11] S. Kim, J.B. Yoo, G.-R. Yi, Y. Lee, H.R. Choi, J.C. Koo, J.-.S. Oh, J.-D. Nam, An aggregationmediated assembly of graphene oxide on amine-functionalized poly(glycidyl methacrylate) microspheres for core-shell structures with controlled electrical conductivity, J. Mater. Chem. C 2 (2014) 64626466.
[12] M. Hassan, K.R. Reddy, E. Haque, S.N. Faisal, S. Ghasemi, A.I. Minett, V.G. Gomes, Hierarchical assembly of graphene/polyaniline nanostructures to synthesize freestanding supercapacitor electrode, Compos. Sci. Technol. 98 (2014) 18.
[13] S.J. Han, H.-I. Lee, H.M. Jeong, B.K. Kim, A.V. Raghu, K.R. Reddy, Graphene modied
lipophilically by stearic acid and its composite with low density polyethylene, J.
Macromol. Sci. B 53 (2014) 11931204.
[14] H. Choi, D.H. Kim, A.V. Raghu, K.R. Reddy, H.-I. Lee, K.S. Yoon, H.M. Jeong, B.K. Kim,
Properties of graphene/waterborne polyurethane nanocomposites cast from colloidal dispersion mixtures, J. Macromol. Sci. B Phys. 51 (1) (2012) 197207.
[15] M. Hassan, K.R. Reddy, E. Haque, A.I. Minett, V.G. Gomes, High-yield aqueous phase
exfoliation of graphene for facile nanocomposite synthesis via emulsion polymerization, J. Colloid Interface Sci. 410 (2013) 4351.
[16] K.R. Reddy, B.C. Sin, C.H. Yoo, D. Sohn, Y. Lee, Coating of multiwalled carbon nanotubes with polymer nanosphere through microemulsion polymerization, J. Colloid
Interface Sci. 340 (2009) 160165.
[17] A. Nol, J. Faucheu, J.-M. Chenal, J.-P. Viricelle, E. Bourgeat-Lami, Electrical and mechanical percolation in graphene-latex nanocomposites, Polymer 55 (2014)
51405145.
[18] K.R. Reddy, M. Hassan, V.G. Gomes, Hybrid nanostructures based on titanium dioxide for enhanced photocatalysis, Appl. Catal. A Gen. 489 (2015) 116.
[19] K.R. Reddy, B.C. Sin, K.S. Ryu, J.C. Kim, H. Chung, Y. Lee, Conducting polymer functionalized multi-walled carbon nanotubes with noble metal nanoparticles: synthesis, morphological characteristics and electrical properties, Synth. Met. 159 (2009)
595603.
[20] Y.R. Lee, S.C. Kim, H.-i. Lee, H.M. Jeong, A.A.V. Raghu, K.R. Reddy, B.K. Kim,
Graphite oxides as effective re retardants of epoxy resin, Macromol. Res. 19
(2011) 6671.
[21] K.R. Reddy, K.P. Lee, A.I. Gopalan, M.S. Kim, A.M. Showkat, Y.C. Nho, Synthesis of
metal (Fe or Pd)/Alloy (FePd)-nanoparticles-embedded multiwall carbon nanotube/sulfonated polyaniline composites by irradiation, J. Polym. Sci. A Polym.
Chem. 44 (2006) 33553364.
[22] Y.M. Lee, B. Jung, Y.H. Kim, A.R. Park, S. Han, W.S. Choe, P.J. Yoo, Nanomeshstructured ultrathin membranes harnessing the unidirectional alignment of viruses
on a graphene-oxide lm, Adv. Mater. 26 (2014) 38993904.
[23] L. He, L.F. Dumee, C. Feng, L. Velleman, R. Reis, F. She, W. Gao, L. Kong, Promoted
water transport across graphene oxide-polyamide thin lm composite membranes
and their antibacterial activity, Desalination 365 (2015) 126135.
[24] H.-R. Chae, J. Lee, C.-H. Lee, I.-C. Kim, P.-K. Park, Graphene oxide-embedded thin-lm
composite reverse osmosis membrane with high ux, anti-biofouling, and chlorine
resistance, J. Membr. Sci. 483 (2015) 128135.
[25] J. Liang, Y. Huang, F. Zhang, Y. Zhang, N. Li, Y. Chen, The use of graphene oxide membranes for the softening of hard water, Sci. China Technol. Sci. 57 (2014) 284287.

76

M.E.A. Ali et al. / Desalination 386 (2016) 6776

[26] S.G. Kim, D.H. Hyeon, J.H. Chun, B.-H. Chun, S.H. Kim, Novel thin nanocomposite RO
membranes for chlorine resistance, Desalin. Water Treat. 51 (2013) 63386345.
[27] J. Yin, B. Deng, Polymer-matrix nanocomposite membranes for water treatment, J.
Membr. Sci. 479 (2015) 256275.
[28] F. Perreault, M.E. Tousley, M. Elimelech, Thin-lm composite polyamide membranes
functionalized with biocidal graphene oxide nanosheets, Environ. Sci. Technol. Lett.
1 (2014) 7176.
[29] S. Bano, A. Mahmood, S.-J. Kim, K.-H. Lee, Graphene oxide modied polyamide
nanoltration membrane with improved ux and antifouling properties, J. Mater.
Chem. A 3 (2015) 20652071.
[30] W. Choi, J. Choi, J. Bang, J.H. Lee, Layer-by-layer assembly of graphene oxide nanosheets on polyamide membranes for durable reverse-osmosis applications, ACS
Appl. Mater. Interfaces 5 (2013) 1251012519.
[31] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. 80
(1958) 1339.
[32] H.M. Hegab, Y. Wimalasiri, M. Ginic-Markovic, L. Zou, Improving the fouling
resistance of brackish water membranes via surface modication with graphene
oxide functionalized chitosan, Desalination 365 (2015) 99107.
[33] M. Ionita, E. Vasile, L.E. Crica, S.I. Voicu, A.M. Pandele, S. Dinescu, L. Predoiu, B.
Galateanu, A. Hermenean, M. Costache, Synthesis, characterization and in vitro
studies of polysulfonegraphene oxide composite membranes, Compos. Part B 72
(2015) 108115.
[34] B.-H. Jeong, E.M.V. Hoek, Y. Yan, A. Subramani, X. Huang, G. Hurwitz, A.K. Ghosh, A.
Jawor, Interfacial polymerization of thin lm nanocomposites: a new concept for
reverse osmosis membranes, J. Membr. Sci. 294 (2007) 17.
[35] V. Nayak, M.S. Jyothi, R.G. Balakrishna, M. Padaki, A.F. Ismail, Preparation and characterization of citosan tin ms on mxed-matrix membranes for complete removal of
chromium, ChemistryOpen 4 (2015) 278287.
[36] Y. Gao, M. Hu, B. Mi, Membrane surface modication with TiO2-graphene oxide for
enhanced photocatalytic performance, J. Membr. Sci. 455 (2014) 349356.
[37] S.-T. Kao, S.-H. Huang, D.-J. Liaw, W.-C. Chao, C.-C. Hu, C.-L. Li, D.-M. Wang, K.-R. Lee,
J.-Y. Lai, Interfacially polymerized thin lm composite polyamide membrane:
positron annihilation spectroscopic study, characterization and pervaporation performance, Polym. J. 42 (2010) 242248.

[38] Z. Xie, M. Hoang, T. Duong, D. Ng, B. Dao, S. Gray, Solgel derived poly(vinyl alcohol)/maleic acid/silica hybrid membrane for desalination by pervaporation,
J. Membr. Sci. 383 (2011) 96103.
[39] J. Zhang, Z. Xu, W. Mai, C. Min, B. Zhou, M. Shan, Y. Li, C. Yang, Z. Wang, X. Qian, Improved hydrophilicity, permeability, antifouling and mechanical performance of
PVDF composite ultraltration membranes tailored by oxidized low-dimensional
carbon nanomaterials, J. Mater. Chem. A 1 (2013) 31013111.
[40] J. Lee, H.-R. Chae, Y.J. Won, K. Lee, C.-H. Lee, H.H. Lee, I.-C. Kim, J.-m. Lee, Graphene
oxide nanoplatelets composite membrane with hydrophilic and antifouling properties for wastewater treatment, J. Membr. Sci. 448 (2013) 223230.
[41] M. Dalwani, N.E. Benes, G. Bargeman, D. Stamatialis, M. Wessling, Effect of pH on the
performance of polyamide/polyacrylonitrile based thin lm composite membranes,
J. Membr. Sci. 372 (2011) 228238.
[42] R.L.D. Whitby, A. Korobeinyk, V.M. Gun'ko, R. Busquets, A.B. Cundy, K. Laszlo, J.
Skubiszewska-Zieba, R. Leboda, E. Tombacz, I.Y. Toth, K. Kovacs, S.V. Mikhalovsky,
pH-driven physicochemical conformational changes of single-layer graphene
oxide, Chem. Commun. 47 (2011) 96459647.
[43] E.I. Mouhoumed, A. Szymczyk, A. Schafer, L. Paugam, Y.-H. La, Physico-chemical
characterization of polyamide NF/RO membranes: insight from streaming current
measurements, J. Membr. Sci. 461 (2014) 130138.
[44] R.K. Joshi, P. Carbone, F.C. Wang, V.G. Kravets, Y. Su, I.V. Grigorieva, H.A. Wu, A.K.
Geim, R.R. Nair, Precise and ultrafast molecular sieving through graphene oxide
membranes, Science 343 (2014) 752754.
[45] F. Wang, H. Li, Q. Liu, Z. Li, R. Li, H. Zhang, L. Liu, G.A. Emelchenko, J. Wang, A
graphene oxide/amidoxime hydrogel for enhanced uranium capture, Sci. Rep. 6
(2016) 19367, http://dx.doi.org/10.1038/srep19367.
[46] B.M. Ganesh, A.M. Isloor, A.F. Ismail, Enhanced hydrophilicity and salt rejection study of
graphene oxide-polysulfone mixed matrix membrane, Desalination 313 (2013)
199207.
[47] J. Glater, S.-k. Hong, M. Elimelech, The search for a chlorine-resistant reverse
osmosis membrane, Desalination 95 (1994) 325345.

You might also like