You are on page 1of 11

Journal of Membrane Science 511 (2016) 29–39

Contents lists available at ScienceDirect

Journal of Membrane Science


journal homepage: www.elsevier.com/locate/memsci

Developing high throughput thin film composite polyamide


membranes for forward osmosis treatment of SAGD produced water
Behnam Khorshidi a, Amrit Bhinder a, Thomas Thundat b, David Pernitsky c,
Mohtada Sadrzadeh a,n
a
Department of Mechanical Engineering, University of Alberta, 6-074 NINT Building, Edmonton, Alberta, Canada T6G 2G8
b
Department of Chemical and Material Engineering, University of Alberta, 7-026 ECERF, Edmonton, Alberta, Canada T6G 2V4
c
Suncor Energy Inc., P.O. Box 2844, 150-6th Ave. SW, Calgary, Alberta, Canada T2P 3E3

art ic l e i nf o a b s t r a c t

Article history: Application of thin film composite (TFC) membranes for forward osmosis (FO) separation processes has
Received 15 October 2015 attracted growing attention due to their outstanding permeation properties as compared to conventional
Received in revised form asymmetric membranes. The aim of the present study is to fabricate high-performance TFC membranes
27 February 2016
by an innovative adjustment of interfacial polymerization (IP) reaction between m-phenylenediamine
Accepted 28 March 2016
Available online 29 March 2016
(MPD) and trimesoyl chloride (TMC) at the surface of a polyethersulfone (PES) microporous support. It
was found that reducing the temperature of the organic solution down to  20 °C effectively reduced the
Keywords: thickness of the PA selective layer and thus, significantly enhanced water permeation through the
Forward osmosis membranes. The water flux increased more than double from 17.6 LMH for membranes prepared at 25 °C
Thin film composite
to 38.5 LMH at  20 °C, when 3 M NaCl solution and de-ionized water were used as draw and feed
Polyamide
solutions, respectively. In addition, all the lab-made membranes showed significantly lower specific
Interfacial polymerization
Oil sands solute flux than the commercial membrane. The performance of lab-synthesized TFC membranes was
SAGD also evaluated for the treatment of boiler feed water (BFW) of steam assisted gravity drainage (SAGD)
process. The results showed superior permeation properties of lab-made membranes to commercially
available TFC-FO membranes. This was attributed to the thinner PA selective layer and lower structural
parameter (451 7 13) of the lab-made membranes compared to the commercial membrane (1770 μm)
which alleviated the effect of internal concentration polarization (ICP) remarkably. This study provides
valuable guidelines for an effective tuning of the organic solution temperature during the IP reaction to
develop high-throughput TFC FO membranes.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction solute rejection, low concentration polarization and fouling pro-


pensity as well as high chemical and mechanical stability [5,6]. The
Forward osmosis (FO) has attracted increasing interest in the majority of the recent advances in the FO process is devoted to
past decade as an alternative to conventional pressure-driven membrane materials development with the aim of fabricating
membrane processes for various applications including seawater high-performance FO membranes [7]. Commonly used commer-
desalination [1], wastewater treatment [2], food processing [3] and cial FO membranes are made from asymmetric cellulose triacetate
clean energy generation [4]. The mechanism of transport in a FO (CTA) [8–10] which became popular due to its hydrophilic nature
membrane is based on the osmotic pressure gradient between a and relatively low-cost [11]. However, the major drawbacks of the
low-concentration solution, known as “feed”, and a high con- CTA membranes are their low-permselectivity and poor stability in
centration solution which is referred to as “draw”. Applying a semi- harsh acidic and basic environment [12]. These limitations have
permeable membrane between the draw and the feed solutions diverted attention from single layer cellulose based membranes
provides pathways for water molecules while restricting the pas- toward thin film composite (TFC) polyamide (PA) based
sage of solutes from one side of the membrane to the other side. membranes.
An ideal FO membrane exhibits high water permeability and A typical TFC membrane comprises a top skin layer (∼100 to
300 nm thick, mostly made from PA) formed by an interfacial
n
Corresponding author. polymerization (IP) reaction at the surface of a microporous sub-
E-mail address: sadrzade@ualberta.ca (M. Sadrzadeh). strate [13,14]. The support substrate usually consists of polysulfone

http://dx.doi.org/10.1016/j.memsci.2016.03.052
0376-7388/& 2016 Elsevier B.V. All rights reserved.
30 B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39

(PSf) or polyethersulfone (PES) cast over a polyester fabric, typi- the presence of nanoparticles limit further development of the
cally polyethylene terephthalate (PET), using a phase inversion TFN membranes.
technique [15,16]. The composite structure of the TFC membranes In the present work, we present an easily-implemented ap-
provides beneficial flexibility in their design, as both the top active proach for fabrication of high throughput TFC-FO membranes by
and the bottom support layers can be tailored separately to opti- synthesizing an ultra-thin PA film over a PES microporous support.
mize the final performance [11]. Although the TFC membranes are The PA film is prepared via an IP reaction between m-phenyle-
very popular in pressure-driven separation processes like reverse nediamine (MPD) and trimesoyl chloride (TMC) at the surface of a
osmosis (RO) and nanofiltration (NF), their application in FO is at polyethersulfone (PES) substrate. The key innovative adjustment
the early stage [17]. TFC-RO and TFC-NF membranes typically have in this approach was reducing the temperature of the TMC-hep-
a dense active layer to provide high selectivity and a thick support tane solution to 20 °C. Three TFC membranes were prepared
layer to offer mechanical stability when external hydraulic pres- with TMC-heptane solutions at 25 °C, 1 °C and  20 °C, and their
sure is applied. But these membranes exhibit low permeation flux surface physico-chemical and morphological characteristics were
when tested for the FO process, as in the absence of hydraulic analyzed using field emission scanning electron microscopy
pressure, the dense active layer hinders the permeation flux (FESEM), transmission electron microscopy (TEM), and attenuated
through the membrane [3,8,10]. Additionally, the thick and dense total reflection-Fourier transform infrared (ATR-FTIR) spectro-
support layer provides a large resistance against the diffusion of scopy, X-ray photoelectron spectroscopy (XPS), atomic force mi-
the draw solute to the back side of the active layer, contributing to croscopy (AFM), and contact angle measurements. The permeation
internal concentration polarization (ICP) phenomenon, thereby performance and structural properties of the lab-made TFC
adversely affecting the water permeation of the membranes [18]. membranes were evaluated by both RO and FO tests and com-
The ICP generally occurs inside the pores of the porous support pared with a commercially available TFC-FO membrane (called
layer and depends mainly upon the thickness, porosity and tor- TFC-HTI here). Additionally, in order to evaluate the separation
tuosity of the support layer rather than the hydrodynamics of the performance of the synthesized TFC membranes for treatment of
flow [18]. Ideally, the support layer should be thin, highly porous industrial water with different types of contaminants, FO tests
with low tortuosity [18,19]. To date, numerous efforts have been were carried out using boiler feed water (BFW) of steam assisted
made on modification of TFC membranes in terms of the physico- gravity drainage (SAGD) process.
chemical characteristics of both the active and support layers to SAGD is a thermally enhanced heavy oil recovery method
make them efficient for the FO process. Much of these efforts were which is widely practiced for bitumen extraction from oil sands in
dedicated to improve the support layer characteristics by the fol- Alberta, Canada. In this process, steam is injected through a hor-
lowing strategies: izontal well into the bitumen-containing formation to decrease
the viscosity of the bitumen and effect its extraction. An emulsion
(1) Modifying the support layer morphology [10,17,20–23] of steam condensate and heated bitumen flows down along the
(2) Increasing the hydrophilicity of the common support materi- periphery of the steam chamber to the production well which is
als (i.e. PES or PSf), either by blending them with more hy- located below the injection well. This emulsion is then pumped to
drophilic materials such as sulfonated polysulfone (SPSf) [5], the surface where the bitumen and water are separated and the
polyphenylsulfone (PPSU) [21], carboxylated polysulfone water is treated for reuse as BFW. In a conventional SAGD plant,
(CPSf) [24], Montmorillonite (MMT)-sulfonated polyethersul- the re-used produced water is de-oiled and treated by lime soft-
fone (SPES) [25] or by coating them with hydrophilic polymers ening and ion exchange to reduce inorganic scalants, and then
like polydopamine [8] or sodium dodecyl sulfate (SDS) [26] used as BFW for a robust style of steam generator known as a once
(3) Using alternative hydrophilic support materials like cellulose through steam generator (OTSG), which can tolerate relatively
acetate (CA) [27], cellulose acetate propionate (CAP) [28], high amounts of TDS, but is limited to low-quality steam (70–85%).
polyacrylonitrile (PAN) [29], polyketone (PK) [30], sulfonated In this study, the feasibility of FO process to be used as a polishing
poly(ether ketone) (SPEK) [31], and sulfonated polyphenyle- step after the conventional lime softening/ion exchange treatment
nesulfone (sPPSU) [32] processes was evaluated to yield high-quality desalinated BFW
(4) Incorporating nanofillers like TiO2 and layered double hydro- suitable for higher efficiency drum boilers, while consuming less
xide (LDH) nanoparticles into the support polymer matrix energy than if desalination evaporators were used.
[33,34]
(5) Using a highly porous electrospun nanofiber matrix of poly-
vinylidene fluoride (PVDF), PES and PAN as support layer [35– 2. Materials and methods
39]
2.1. Chemicals and reagents
According to our search of the literature, there are only a few
studies focused on modifying the characteristics of the active thin A microporous polyethersulfone (PES, 0.2 μm) membrane was
layer of the TFC membranes for FO application [40], and these are provided by Sterlitech Co. (WA, USA) and used as the support for
mainly focused on incorporating nanoparticles like Zeolite, SiO2, the TFC membranes fabricated in this research. MPD (Z99%), TMC
graphene oxide (GO), and multi-walled carbon nanotube (98%) and camphorsulfonic acid (CSA) were obtained from Sigma-
(MWCNT) into the active layer of the TFC membrane [41–45]. Aldrich. Heptane (Z99%), triethylamine (TEA) and sodium dode-
However, the addition of nanoparticles in a thin film with the aim cyl sulfate (SDS) were purchased from Fisher Scientific. All mate-
of fabricating a thin film nanocomposite (TFN) membrane is more rials were used as they were received from the suppliers. Com-
complicated than it appears, due to (i) severe aggregation of the mercial PA TFC-FO membranes with embedded polyester fabric
nanoparticles in the organic solvent during the IP reaction, and (ii) support were purchased from Hydration Technology Innovation
weak compatibility of the nanoparticles with the synthesized (HTI, Albany, USA). The industrial process water used for mea-
polymer matrix. It is well known that non-uniform dispersion of suring the membranes permeation performance and fouling
nanoparticles forms non-selective voids at the interface of the characteristics was conventionally-treated SAGD BFW obtained
polymer and the nanoparticles, which significantly reduces the from an in-situ water treatment plant located in the Athabasca oil
rejection percentage [46,47]. The uncertainties related to the for- sands region of Alberta, Canada. The pH of this water, as received
mation of an integrally-skinned and defect-free thin film due to was 10, and was not adjusted during the filtration experiments.
B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39 31

Table 1 membranes was evaluated by the sessile drop method using a


Properties of conventionally-treated SAGD boiler feed water (BFW). Krüss DSA 100 (Krüss GmbH, Germany) contact angle analyzer.
The contact angle was measured at 5 different places on the
Parameter Unit Specification
membrane surface and the results are reported as an average.
Conductivity mS 1.8–2.2 Analysis of the samples for the presence of organic matter was
pH – 9.8–10.5 done using a combustion type total organic carbon (TOC) analyzer
TDS mg/L 1800–2000
(Shimadzu, model TOC-V; detection range 3–25,000 mg/L). The
TOC mg/L 450–550
Silica (as SiO2) mg/L 30–40 concentration of inorganic elements such as silica, Ca þ 2, Mg þ 2
Calcium mg/L 0.4–0.5 was measured using inductively coupled plasma-optical emission
Iron mg/L 0.2–0.3 spectroscopy (ICP-OES) (Agilent 735 ICP-OES).

2.4. Evaluation of membrane properties in reverse osmosis


Table 1 summarizes the properties of the boiler feed water used in
the present work. The pure water permeability coefficient, A, and solute (NaCl)
permeability coefficient, B, of the TFC membranes were evaluated
2.2. Synthesis of thin film composite (TFC) forward osmosis (FO) using a cross-flow reverse osmosis (RO) filtration setup following
membranes the standard protocol used by Cath et al. [49]. The detailed de-
scription of the RO test unit is presented elsewhere [50]. The fil-
The TFC membranes were prepared via an IP reaction between tration test was performed at room temperature (23 72 °C), con-
MPD and TMC at the surface of the PES support. The PES substrate trolled by water bath (Isotemp 3013, Fisher Scientific), at a con-
was first immersed in 2 wt% MPD-water solution with 2 wt% CSA, stant cross-flow velocity of 0.25 m/s. The pure water permeability
1 wt% TEA and 0.2 wt% SDS for 10 min. TEA was used as an acy- was measured using DI water as the feed solution at 5.1 atm
lation catalyst to promote the polymerization reaction by remov- (75 psi), 6.8 atm (100 psi) and 8.5 atm (125 psi) trans-membrane
ing HCl, which is formed as a reaction by-product. CSA and SDS pressures. The flux was allowed to stabilize at each pressure for at
were used to enhance the wettability and absorption of the MPD least 30 min and the corresponding water flux was measured by
solution into the PES support [48]. The substrate was then re- dividing the volumetric permeate rate by the membrane active
moved, and excess amine solution was squeezed off the surface area, 42 cm2. The flux values were then plotted against pressure
using a rubber roller. Next, the impregnated PES sheet was where the slope of the linear regression gave the pure water
brought into contact with a 0.2 wt% TMC-heptane solution for 30 s, permeability of the membrane.
to allow a polymerization reaction. The temperature of the TMC- The apparent salt rejection (R) was calculated using the fol-
heptane solution was set to  20 °C, 1 °C and 25 °C using a general lowing equation:
purpose water bath (Isotemp 3013, Fisher Scientific) and a freezer
(Fisher Scientific, Isotemp™ freezer). The resulting composite ⎛ Cp ⎞
R = ⎜1 − ⎟ × 100
membranes were then thermally cured in a digital oven for 5 min ⎝ Cf ⎠ (1)
at 70 °C. Finally, the residual solution from the surface was washed
away using 250 ml deionized (DI) water. The synthesized TFC where Cp and Cf are the concentration of salt in the permeate and
membranes were then kept in the DI water bath until character- feed, respectively. The values of Cp and Cf are obtained based on
ization tests were performed. the calibration curve of the solution conductivity. Sampling was
carried out after 3 h filtration of 2000 mg/L NaCl under an applied
2.3. Membrane characterization pressure of 8.5 atm. The salt permeability coefficient (B) of the
composite membranes was calculated using the following equa-
The surface morphology of the TFC membranes was in- tion [17,30]:
vestigated using FESEM (JEOL 6301F). The TFC membranes were ⎛1 ⎞
sputter coated with a thin film of chromium and imaged at a B = J W ⎜ − 1⎟
⎝R ⎠ (2)
magnification of 30,000  . The cross-sectional images of the TFC
membranes were examined using TEM (Philips/FEI Morgagni 268,
Netherlands) at an acceleration voltage of 80 kV. The samples were 2.5. Measurement of membrane performance in forward osmosis
prepared by first staining in uranyl acetate and lead citrate, then
embedding in spurr's resin, and finally sectioning using ultra- The permeation performance of TFC membranes was evaluated
microtome (Reichert-Jung Ultracut E, USA). ATR-FTIR (Thermo
Nicolet Nexus 670, USA) spectroscopy was used to study the
functional groups present at the surface of the TFC membranes.
The FTIR spectra of the samples were averaged from 512 scans and
were taken over the range of 600–4000 cm  1 at 4 cm  1 resolu-
tion. The elemental composition (C, O, N) of the top 5–10 nm of the
PA skin layer was analyzed using an XPS (Kratos AXIS ULTRA, UK)
equipped with a monochromatic Al Kα X-ray source. Survey
spectra were collected at constant pass energy of 160 eV, with a
scan step size of 0.4 eV, and a sweep time of 100 s in the range of
0–1100 eV. High-resolution spectra for carbon (C 1s) was collected
with pass energy of 20 eV, step size of 0.1 eV, and sweep time of
200 s. The surface topography of the TFC membranes was studied
using AFM (Bruker Dimension Icon, USA). An area of 5 μm  5 μm
of the TFC membranes was scanned three times using the tapping
mode at a scan rate of 1.0 Hz. The AFM data was analyzed using Fig. 1. Schematic view of FO setup. All TFC membranes were tested in active layer
Nanoscope analysis software V.1.40. The surface hydrophilicity of toward feed side (AL-FS) orientation.
32 B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39

using a cross-flow FO filtration setup which is illustrated sche- flux to pure water flux, Js/Jw, was considered as specific flux ratio.
matically in Fig. 1. The membrane cell was designed with channels The FO filtration of BFW was carried out with 0.5 M NaCl salt
on both sides of the membrane. The length, width and height of water as the draw solution and BFW as the feed solution. The FO
both channels were 145 mm, 96 mm, and 2 mm, respectively, tests were conducted over a time period of 6 h under similar op-
providing an effective filtration area of 140 cm2. Thin plastic mesh erating conditions.
spacers were used on both sides of the membrane to provide
mechanical support to the membrane and to induce better mixing 2.6. Determination of the membrane structural parameter
in the channel to reduce external concentration polarization (ECP).
Water permeation of the membrane in FO unit was measured The structural parameter (S) is one of the most important
using salt water with different NaCl concentrations (ranging from properties of a FO membrane, and is used as a measure to evaluate
0.25 M to 3 M) as draw solution and DI water as feed solution. The the vulnerability of the membrane to the ICP. The structural
test was first conducted with DI water on both sides of the parameter is basically a property of the support layer and depends
membranes for 15 min. The volumes of initial draw and feed so- on the support porosity (ε), tortuosity (τ) and thickness (t)
lutions were 2 L and 2.5 L, respectively. Afterwards, in a series of ( S = tτ /ε ). The structural parameter can be obtained using the
stepwise experimental runs, the concentration of the draw solu- following equation [49,52]:
tion was increased to desired values by adding proper amount of
⎛ D⎞ B + Aπ D,b
concentrated (5 M) NaCl solution. The temperature of the feed and S = ⎜ ⎟ ln
draw solutions were maintained at 217 2 °C using Isotemp water ⎝ JW ⎠ B + JW +Aπ F,m (4)
bath (Polyscience, model: MX-CA11B and model: 6560M11A120C,
USA). Variable speed gear pumps were used to maintain the flow where D is the salt (NaCl) diffusion coefficient, π D,b is the osmotic
pressures of the bulk draw solutions, and π F,m is the osmotic
of both solutions at 2.5 L/min (0.22 m/s cross flow velocity). The
pressures of the feed solution on the membrane interface. For high
draw solution tank was placed over a digital weighing balance
salt rejecting membranes, B is typically assumed to be zero [53].
(Mettler Toledo) and the change in the weight of the solution was
The structural parameters of the lab-made and commercial
recorded continuously. For all experiments, the initial volumes of
membranes were evaluated using 1 M NaCl solution and DI water
draw and feed solutions were 2 L and 2.5 L, respectively. During
as draw and feed solution, respectively. The cross-flow velocity of
the test, the conductivity of the feed and draw solutions was
both draw and feed solutions was maintained at 0.22 m/s. Using DI
monitored using in-line conductivity sensors and recorded auto-
water as feed Eq. (4) is simplified to:
matically every minute.
The water flux through the membrane was calculated by ⎛ D⎞ B + Aπ D,b
measuring the change in the weight of the draw solution (ΔmD ) S = ⎜ ⎟ ln
⎝ JW ⎠ B + JW (5)
passed through the effective surface area (Am) of the membrane
over a specific time period of the experiment (Δt) which was
about 30 min for each draw solution concentration:
ΔmD 3. Results and discussion
Jw =
ρ D AmΔt (3)
3.1. Membrane morphology
where ρD is the density of the draw solution [51]. The reverse salt
flux, Js, in gm  2 h  1 (gMH) was calculated by dividing the mass Figs. 2 and 3 show the surface and cross-sectional images of the
flow rate of NaCl (obtained via measuring the feed conductivity) PES support as well as the synthesized and commercial TFC
by the membrane effective surface area. The ratio of reverse solute membranes. The PES substrate has a smooth surface with average

Fig. 2. FESEM top surface images of (a) PES microporous support; (b) active-side of TFC-HTI; (c) support-side of TFC-HTI; (d) TFC1 (prepared at 25 °C); (e) TFC2 (prepared at
1 °C); (f) TFC3 (prepared at  20 °C).
B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39 33

Fig. 3. FESEM cross-sectional images of (a) PES microporous support; (b) TFC-HTI; (c) TFC1 (prepared at 25 °C); (d) TFC2 (prepared at 1 °C); (e) TFC3 (prepared at  20 °C).

micro-pores of 200 nm (Fig. 2a). The surface morphologies of the of the MPD molecules toward the reaction zone, thereby reducing
front and back sides of the commercial TFC-HTI membrane are the rate of the polymerization reaction. Hence, the size of features
presented in Fig. 2b and c, respectively. The active surface of the is reduced at lower temperatures of the organic solution [54].
TFC-HTI membrane has the typical “ridge-and-valley” structure of The cross-sectional image of the PES microporous membrane
the PA-based TFC membranes (Fig. 2b). According to Fig. 2c, the PA that utilized as a support in the present study displays a sym-
layer is formed on a polysulfone (PSf) support embedded in a metric sponge-type structure with a thickness of about 140 μm
woven polyester mesh. (Fig. 3a), whereas the support layer of the commercial TFC-HTI
The top surface morphologies of the lab-made TFC membranes membrane shows an asymmetric finger-like structure with a
which were prepared at different temperatures of the TMC-hep- denser skin layer (Fig. 3b). The different morphologies of these
tane solution are demonstrated in Fig. 2d‐f. The noticeable differ- support is predicted to provide different structural parameters as
ence in these figures is the structure of the PA skin layer which will be discussed later. The FESEM cross-sectional image of the
varied with temperature of the organic solution during the inter- TFC1 membrane (Fig. 3c) also confirms the homogeneous forma-
facial polymerization reaction. TFC1 membrane, prepared at 25 °C, tion of the PA film at the PES surface as was observed in its surface
has flexuous ridges and valleys, similar to the active surface of the image (Fig. 2d). A closer examination of Fig. (3d and e) reveals that
TFC-HTI membrane. The surface of the TFC1 membrane, prepared formation of PA ridges and valleys structure is hindered by de-
at 25 °C, has flexuous ridges and valleys which covered all the creasing the temperature of the organic solution.
surface of the porous sublayer, similar to the active surface of the The TEM images of the synthesized TFC membranes also con-
TFC-HTI membrane (Fig. 2b and d). However, by decreasing the firm formation of thicker PA layer for TFC1 membrane as com-
temperature of the organic solution, the size of the surface fea- pared to TFC2 and TFC3 (Fig. 3). This observation again demon-
tures markedly scaled down to small protuberances which is evi- strates the critical role of organic solvent temperature on the
dent in the FESEM images of the TFC2 and TFC3 membranes morphology and thickness of the synthesized TFC membranes. It is
prepared at þ1 °C and 20 °C, respectively (Fig. 2e and f). It is worth mentioning that, the visibility of the support pores in spite
believed that the formation of the PA selective layer during the of formation of a PA film over them in the active surface of TFC2
interfacial polymerization reaction occurs in two successive stages and TFC3 membranes (Fig. 2e and f) is due to the very low thick-
[14]. The IP reaction at the early stage is very fast, resulting in an ness of the PA layer in these membranes which could not cover the
incipient PA film at the PES substrate. Afterwards, the slower porous surface of the PES supports. However, these visible pores
growth stage starts when the initial PA film hinders the diffusion were not open-ended. Taking a closer look at the TEM images of
of the MPD molecules to the reaction zone. Decreasing the tem- these membranes (Fig. 4b and c), it can be noticed that the support
perature of the organic solution is believed to hinder the diffusion pores were internally closed by the PA film. If this were not the

Fig. 4. TEM cross-sectional images of (a) TFC1 (prepared at 25 °C); (b) TFC2 (prepared at 1 °C); (c) TFC3 (prepared at  20 °C).
34 B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39

over the PES sublayer.


Deconvolution of C (1s) high-resolution XPS spectra provided
valuable information about the difference in chemical bonding of
the PA layer between the lab-made TFC3 and commercial TFC-HTI
membrane. The C (1s) high-resolution spectra for TFC3 membrane
(Fig. 6b) shows the presence of three peaks, (i) a major peak at
285 eV assigned to a carbon atom without adjacent electron
withdrawing atoms (C–C and C–H), (ii) an intermediate peak at
286.5 eV associated with carbon with weak electron withdrawing
atoms (C–N), and (iii) a minor peak at 288.5 eV related to carbons
attached to strong electron withdrawing atoms (carboxylic O¼ C–O
and amides O ¼C–N) [59]. There are two distinct changes in the
high-resolution C (1s) spectrum of the TFC-HTI membrane as
compared to TFC3 membranes (Fig. 6a). First, the main peak
skewed to the left, which demonstrates the presence of more
hydroxyl groups (OH) on the surface. The commercial membranes
used in the present work were soaked in glycerol (having three OH
groups, C3H8O3) which indicates surface modification of these
membranes with polyols (sugar alcohol) to increase hydrophilicity,
and thereby increasing the water flux through the membranes.
Second, the peak related to carboxylic acid (O ¼C–O) is detached
from amide (O ¼C–N) which shows lower cross-link density of the
TFC-HTI membrane due to the higher amount of carboxylic func-
tional groups in the linear portion of the PA network at the
membrane surface.

3.3. Roughness and hydrophilicity of TFC membranes

The values of average roughness (Ra) and root mean squared


roughness (Rrms) of the TFC membranes are summarized in Ta-
ble 2. In overall, the average roughness of all TFC membranes was
comparable and varied in the range of 45–55 nm. The average
roughness slightly decreased from TFC1 to TFC3, implying that
synthesis of PA film at lower temperature of organic solution re-
Fig. 5. FTIR Spectra of the PES support, lab-made and commercial TFC membranes. sulted in a skin layer with little lower surface roughness. This can
also be observed in the two and three-dimensional AFM images of
the TFC membranes, illustrated in Fig. 7. The ridges and valleys are
case, the selectivity of these membranes would decrease drasti-
a bit sharper at the surface of the TFC1 (and also TFC-HTI) mem-
cally. The marked reduction in the thickness of the PA active layer
brane, while they are slightly flattened out at the surface of TFC2
of the TFC2 and TFC3 membranes is the main reason for the high
and TFC3 membranes. Table 2 presents the contact angle value of
permeation performance of these membranes, which will be dis-
TFC membranes. Measuring the contact angle is a common tech-
cussed in the following sections.
nique for evaluating the surface wettability of the membranes
[60,61]. According to this table, the contact angle value of lab-
3.2. Chemical bond information and elemental analysis of
made membranes decreased in the order of TFC1 to TFC3, sug-
membranes
gesting that the surface wettability of the TFC membrane in-
creased when the membrane was made at lower temperatures.
Fig. 5 presents the FTIR spectra of the PES support and TFC PA
Additionally, the very low contact angle of the TFC-HTI membrane
membranes. The FTIR spectrum detects the peaks attributed to
(24.0 71.3°) can be attributed to the surface modification of this
both the PES substrate and the PA skin layer of the TFC membranes
due to the high-depth of penetration of the IR beam at the wa- membrane by soaking in glycerol to induce more hydrophilicity as
venumbers of 1300–1800 cm  1. The peaks at 1486 and 1580 cm  1 discussed in previous section.
belong to the benzene rings of the PES sublayer in lab-made TFC
membranes and the PSf in TFC-HTI membranes [55]. The active 3.4. FO separation performance
layer of all TFC membranes exhibited the characteristic PA peaks at
1545 cm  1 (amide II, C-N stretching), 1610 cm  1 (aromatic ring) The performance of the lab-made and commercial TFC mem-
and 1655 cm  1 (amide I, C ¼O stretching) [56–58]. branes is evaluated by FO tests over a range of osmotic pressure
XPS analysis provides information about the elements and the difference and the results are presented in Fig. 8. The osmotic
chemical bonds present in the top 1–5 nm of the PA layer. The pressure was calculated by the following polynomial equation as a
survey and high-resolution C (1s) XPS spectra for the TFC3 and function of NaCl concentration (mol Lit  1) [62]:
TFC-HTI membranes along with the chemical structure of the PA
π = 6.2971 C2 + 40.714 C (6)
skin layer are presented in Fig. 6. The XPS survey spectra of both
TFC-HTI and TFC3 membranes show the presence of only three The calculated osmotic pressure by this equation increases
elements, including oxygen (O 1s), nitrogen (N 1s) and carbon (C non-linearly with increase in NaCl concentration, especially at
1s) at the membrane surface. The absence of a sulfur peak, which higher concentrations. According to Fig. 8, the water permeability
is the main peak of the sulfone polymers typically used as support of all TFC membranes increased with increases in the osmotic
(e.g. PES and PSf), suggests successful formation of the PA layer pressure difference between the draw and the feed solutions.
B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39 35

Fig. 6. XPS survey and high-resolution C 1s for (a) TFC-HTI and (b) TFC3 membranes along with (c) chemical structure of a typical PA synthesized by IP reaction; m and n are
the cross-linked and linear part, respectively (m þ n ¼ 1).

Table 2 the TFC3 membrane is primarily attributed to the formation of a


Surface roughness and surface wettability of the composite membranes. thinner PA selective layer at the surface of the PES support, as
indicated in FESEM and TEM images, which provided lower mass
Membrane Surface roughness (nm) Contact angle (°)
transfer resistance toward water passage through the membrane.
Ra Rrms The reverse solute flux and the specific salt flux through the
lab-made and commercial TFC membranes are shown in Fig. 9. The
TFC1 53.0 7 1.6 66.37 2.4 81.2 7 1.6 reverse solute flux takes place due to the imperfection of the ac-
TFC2 51.2 7 2.3 68.07 3.8 56.9 7 1.1
TFC3 49.57 2.1 65.17 2.2 53.3 7 1.2
tive layer of the TFC membrane. An ideal selective membrane
TFC-Com 46.3 7 1.8 56.7 7 2.3 24.07 1.3 would allow no draw solute to pass through and reach the feed
side. The specific solute flux is a caliber indicating how much salt
passes through the membrane per unit volume of permeated
However, the rate of flux enhancement was initially high at lower water. In general, a lower value of both reverse salt flux and spe-
values of osmotic driving forces and then declined at higher os- cific solute flux is desirable for development of high-performance
motic pressure difference, mainly due to the more significant di- membranes [63].
lutive ICP phenomenon within the porous support layer. Com- According to Fig. 9, the reverse salt flux of the synthesized
paring the performance of the membranes reveals that all the lab- membranes increased from TFC1 to TFC3. This can be attributed to
made (TFC1-3) membranes showed higher water permeation than the formation of a thinner active layer with lower selectivity by
the commercial TFC-HTI membrane. The TFC3 membrane which using organic solution at lower temperatures. Comparing with
was prepared at the lowest organic solution temperature TFC-HTI membrane, the TFC2 membrane showed superior per-
(  20 °C), outperformed the other membranes with a high per- formance with two times higher water permeation and 50% less
meation rate of 38.5 LMH at osmotic pressure of 179 bar. The TFC2 reverse salt flux. Although TFC3 had slightly higher reverse salt
(prepared at þ 1 °C) and TFC1 (prepared at þ25 °C) showed water flux, its specific salt flux was significantly lower than the TFC-HTI
permeation of 17.6 LMH and 24.1 LMH, respectively, implying that membranes. The specific solute flux of the other two lab-made
the water permeability of the membrane significantly improved by membranes was also remarkably lower than TFC-HTI. The specific
decreasing the temperature of the organic solution during the solute flux results clearly indicate that all lab-synthesized mem-
interfacial polymerization reaction. The high permeation rate of branes perform more efficiently than the commercial membrane
36 B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39

Fig. 7. Surface topography of (a) TFC-HTI; (b) TFC1; (c) TFC2; (d) TFC3 membranes.

in a FO process.
Table 3 provides a comparison between the intrinsic permea-
tion properties of lab-made TFC membranes and the literature PA
TFC flat sheet membranes under both RO and FO conditions. Ac-
cording to this table, the TFC3 membrane exhibited the highest
pure water permeability coefficient (A) among all the other
membranes.
The water permeability of this membrane is two folds higher
than that of the commercial TFC-HTI (5.78 LMH/bar compared to
2.48 LMH/bar) with comparable salt rejection percentage (93.4%
for TFC1 and 95% for TFC-HTI). This result again confirms the
significant effect of organic solution temperature for controlling
the thickness of the PA active layer and thereby enhancing the
water permeability. Although TFC1 and TFC2 membranes showed
lower water permeability coefficient than the TFC-HTI membrane
under RO test conditions, they provided more water flux in FO
experiment (Fig. 8) which can be presumably attributed to their
significantly lower structural parameter than the TFC-HTI mem-
brane (4517 13 mm for lab-made TFC membranes compared to
1770 for TFC-HTI membrane). From the FESEM images of the TFC-
HTI membrane (Fig. 2c), it is noticed that a polyester woven fabric
Fig. 8. FO performance of lab-made and commercial TFC membranes at different
osmotic pressure difference between draw and feed solutions. Test conditions:
is used to provide extra mechanical support to the membrane. The
draw solution: 0.25, 0.5, 1, 1.5, 2 and 3 M NaCl solutions; Feed solution: DI water; use of this woven fabric might be the main reason for the high-
Cross flow velocity: 0.22 m/s for feed and draw solutions. structural parameter of the TFC-HTI membrane by adding extra
resistance against solute transport inside the porous support.
Furthermore, the asymmetric structure of the PSf support with a
dense skin layer in the TFC-HTI membrane could enhance the ICP
phenomenon and thus decrease the flux. In general, an ideal
support layer for FO separation process needs to be very thin and
hydrophilic, with high-porosity and low-tortuosity in order to fa-
cilitate draw solution passage toward the membrane active layer
[10]. A lower value of the structural parameter is highly desirable
for a membrane to counter the negative impact of ICP during the
FO process.
As discussed above, the flux through a TFC membrane in a
pressure-driven process is mainly governed by a compromise be-
tween the hydrophilicity and the thickness of the PA skin layer,
whereas in FO process, another influential factor, i.e. the structural
parameter, is also brought to action. With a low structural para-
meter along with an ultra-thin selective layer, the TFC3 membrane
provided the highest water permeability, whereas the water flux
through TFC1 and TFC2 was compromised by the increased
thickness of the active layer. It must be noted that the reduction of
Fig. 9. Reverse solute flux and specific solute flux of lab-made and commercial TFC
the active layer thickness in the TFC3 is accompanied by a slight
membranes. Test conditions: draw solution: 1 M NaCl solutions; feed solution: DI reduction in the rejection percentage of the membrane (93.4% for
water; cross flow velocity: 0.22 m/s for feed and draw solutions. TFC3 compared to 97.5% in TFC2). Furthermore, the larger B value
B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39 37

Table 3
Performance of PA TFC flat-sheet membranes in AL-FS orientation.

Membrane, material Modified layer FO performance RO performance S (lm) Reference

Feed solution Draw solution Water flux (LMH) A (LMH/bar) B (LMH) R (%)

TFC1, PA/PES Active DI water 1M NaCl 11.0 0.66 0.35 97.8 460 Present work
TFC2, PA/PES Active DI water 1M NaCl 17.0 1.80 1.00 97.5 458 Present work
TFC3, PA/PES Active DI water 1M NaCl 26.5 5.78 4.96 93.4 436 Present work
Commercial TFC-HTI, PA/PSf – DI water 1M NaCl 7.5 2.48 0.82 95.0 1770 Present work
PA/PSf Support DI water 1M NaCl 15.8 1.16 0.47 97.4 492 [10]
PA/PSf Support DI water 1M NaCl 25.0 1.90 0.33 98.6 312 [20]
PA/PES-co-sPPSU Support DI water 1M NaCl 20.0 0.73 0.25 91.0 324 [21]
PA/PES Support DI water 1M NaCl 47.0 1.70 NA 97.0 80 [35]
PA/PES-SPSF Support DI water 1M NaCl 32.0 0.77 0.11 93.5 238 [5]
PA/CAP Support DI water 1M NaCl 10.0 1.82 0.19 89.2 789 [28]
PA/PSf-SPEK Support DI water 1M NaCl 23.0 0.75 0.07 89.5 107 [31]
PA/PVDF Support DI water 1M NaCl 28.0 3.15 2.33 84.4 325 [38]
PA/PK Support DI water 1M NaCl 27.0 2.50 0.18 NA 280 [30]
Zeolite NaY-PA/PSf Active DI water 1M NaCl 11.0 2.57 1.57 77.6 782 [42]
PA/PVDF Support DI water 1M NaCl 22 1.28 0.28 NA 193 [39]
PA/PSf-LDHs Support DI water 1M NaCl 18.1 0.61 0.27 NA 148 [34]

of the TFC3 compared to the other lab-made membranes can be 3.5. FO separation performance with BFW
presumably attributed to the formation of ultra-thin PA layer. In
spite of the fact that, employing support layer with large pores can The performance of the lab-made and commercial TFC mem-
be beneficial to yield a lower structural parameter, it can make the branes is also evaluated for filtration of a real wastewater (SAGD
top layer more prone to fail when working under high applied BFW) and the reduction of water flux and the draw solution
pressure of RO process due to the lack of sufficient supporting area concentration is plotted over time, as shown in Fig. 10. Similar to
[37]. It has been reported that the pore size distribution of the the trend observed in Fig. 8, the lab-made TFC membranes dis-
support layer has a significant influence on the physico-chemical played higher water flux than the commercial membrane under
properties (e.g. degree of cross-linking, surface morphology, me- identical test conditions. The initial flux for TFC3 (18.1 LMH) and
chanical stability and permselectivity) of the PA active layer TFC2 (12.6 LMH) was about three and two times higher than the
[23,64,65]. Considering a series of hydrophilic microporous sup- TFC-HTI (6.4 LMH) membrane, respectively.
ports with average pore sizes of 0.025, 0.1, 0.2 and 0.45 mm, Huang The water permeation of the TFC3 membrane declined rapidly
and McCutcheon found that water permeability increases gradu- from 18.1 LMH to 12.8 LMH after 6 h operation. There are two
likely causes for such a sharp water flux decline. First, fouling of
ally with increase in the support pore size up to 0.2 mm while the
membrane by depositing the BFW contaminants, such as organic
salt rejection slightly decreases following the permselectivity
matters, dissolved and suspended solids, on the surface induces
trade-off relationship [23]. Regarding that, at higher pore sizes
extra resistance against water transport through the membranes,
(4 100 nm), proper adjustment of the synthesis parameters in
and thus decreases the water flux. Second, the high permeation
interfacial polymerization reaction, particularly the monomers
rate of water through the TFC3 membrane quickly dilutes the draw
concentration ratio and drying process, becomes important for
solution and at the same time concentrates the feed solution
making a defect-free PA selective layer [14,29]. However, unlike
which results in a significant reduction in osmotic pressure dif-
the RO operation, the FO process does not require a large hydraulic
ference. According to Fig. 10, the draw solution concertation of
pressure that allows more flexibility in the synthesis and mod-
TFC3 membrane decreased about 30% and diluted from 0.46 M to
ification of both active and support layers in terms of mechanical 0.3 M over the six hours filtration test. In contrast, for TFC1 and
stability. TFC-HTI membranes, where the permeate flux was comparatively
low, there was not a significant change in the osmotic driving force
(around 14% reduction in draw solution concentration), leading to
a relatively slow decline in the water flux with time. In order to
find the contribution of the aforementioned parameters (fouling
and reduction of osmotic pressure difference) in flux decline
through TFC3, the BFW filtration results were compared with the
test results obtained using DI and saline water as feed and draw
solutions, respectively (Fig. 8). As can be seen, at the same osmotic
pressure difference, the TFC3 membrane showed 3 LMH less initial
water flux during filtration of the BFW which can be attributed to
the fouling phenomenon. Following the same approach for TFC1
and TFC2, however, shows that the fouling contribution for these
membranes is negligible. These findings suggest that the “critical
flux” concept, above which an irreversible fouling appears, is also
valid for the FO process. Hence, a very likely explanation for the
less fouling of TFC1 and TFC2 membrane is operation of these
membranes below critical flux, where the effect of permeation
drag on the deposition of foulants is moderated [66,67]. The se-
Fig. 10. Water permeation of lab-made and commercial TFC membranes. Test
conditions: feed solution: conventionally-treated SAGD BFW; draw solution: 0.5 M paration performance of the TFC membranes was evaluated by
NaCl solution; cross-flow velocity: 0.22 m/s. establishing a mass balance between the concentration of TOC,
38 B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39

silica and divalent ions (calcium and magnesium) in the final and performance thin-film composite forward osmosis membrane, Environ. Sci.
initial feed solutions. All the TFC membranes showed very good Technol. 44 (2010) 3812–3818, http://dx.doi.org/10.1021/es1002555.
[11] C. Klaysom, T.Y. Cath, T. Depuydt, I.F.J. Vankelecom, Forward and pressure re-
rejection percentages (∼99%) toward the removal of organic mat- tarded osmosis: potential solutions for global challenges in energy and water
ter, silica and divalent ions. supply, Chem. Soc. Rev. 42 (2013) 6959–6989, http://dx.doi.org/10.1039/
c3cs60051c.
[12] J. Ren, J.R. McCutcheon, A new commercial thin film composite membrane for
forward osmosis, Desalination 343 (2014) 187–193, http://dx.doi.org/10.1016/j.
4. Conclusion desal.2013.11.026.
[13] W.J. Lau, A.F. Ismail, N. Misdan, M.A. Kassim, A recent progress in thin film
composite membrane: a review, Desalination 287 (2012) 190–199, http://dx.
In the present work, high performance TFC membranes were doi.org/10.1016/j.desal.2011.04.004.
developed for FO operation by controlling the thickness of PA se- [14] B. Khorshidi, T. Thundat, B. Fleck, M. Sadrzadeh, Thin film composite poly-
lective layer via temperature adjustment of organic solution dur- amide membranes: parametric study on the influence of synthesis conditions,
RSC Adv. 5 (2015) 54985–54997, http://dx.doi.org/10.1039/C5RA08317F.
ing the IP reaction. It was found that by reducing the temperature
[15] R.J. Petersen, Composite reverse osmosis and nanofiltration membranes, J.
of the organic solution down to  20 °C, an ultra-thin layer of PA Membr. Sci. 83 (1993) 81–150, http://dx.doi.org/10.1016/0376-7388(93)
film was formed at the support surface. The synthesized TFC 80014-O.
membranes, prepared via thermally-tuned IP reaction, demon- [16] M. Sadrzadeh, S. Bhattacharjee, Rational design of phase inversion membranes
by tailoring thermodynamics and kinetics of casting solution using polymer
strated higher permeation performance compared to the un- additives, J. Membr. Sci. 441 (2013) 31–44, http://dx.doi.org/10.1016/j.
modified as well as commercial TFC membranes. The results ob- memsci.2013.04.009.
tained in this study demonstrate the significant role of the active [17] J. Wei, C. Qiu, C.Y. Tang, R. Wang, A.G. Fane, Synthesis and characterization of
flat-sheet thin film composite forward osmosis membranes, J. Membr. Sci. 372
layer properties (thickness and morphology) on performance of a (2011) 292–302, http://dx.doi.org/10.1016/j.memsci.2011.02.013.
FO membrane. The microporous support used in this work had a [18] J.R. McCutcheon, M. Elimelech, Influence of concentrative and dilutive internal
thickness of about 140 μm with a structural parameter of concentration polarization on flux behavior in forward osmosis, J. Membr. Sci.
284 (2006) 237–247, http://dx.doi.org/10.1016/j.memsci.2006.07.049.
451 713 μm. For further improvement, the support layer can be [19] J.R. Mccutcheon, M. Elimelech, Modeling water flux in forward osmosis: im-
modified either in terms of structure or material to provide a plications for improved membrane design, AIChE J. 53 (2007) 1736–1744, http:
lower structural parameter and minimize the negative impact of //dx.doi.org/10.1002/aic.
[20] A. Tiraferri, N.Y. Yip, W. a Phillip, J.D. Schiffman, M. Elimelech, Relating per-
the ICP, thus maximizing the permeation flux. FO trials using formance of thin-film composite forward osmosis membranes to support
conventionally-treated SAGD boiler feed water indicate that high- layer formation and structure, J. Membr. Sci. 367 (2011) 340–352, http://dx.
performance TFC membranes show promise for the production of doi.org/10.1016/j.memsci.2010.11.014.
[21] N. Widjojo, T.S. Chung, M. Weber, C. Maletzko, V. Warzelhan, The role of
desalinated boiler feed water from SAGD produced water.
sulphonated polymer and macrovoid-free structure in the support layer for
thin-film composite (TFC) forward osmosis (FO) membranes, J. Membr. Sci.
383 (2011) 214–223, http://dx.doi.org/10.1016/j.memsci.2011.08.041.
Acknowledgments [22] Y. Yu, S. Seo, I.-C. Kim, S. Lee, Nanoporous polyethersulfone (PES) membrane
with enhanced flux applied in forward osmosis process, J. Membr. Sci. 375
(2011) 63–68, http://dx.doi.org/10.1016/j.memsci.2011.02.019.
The authors gratefully acknowledge financial support provided [23] L. Huang, J.R. Mccutcheon, Impact of support layer pore size on performance
of thin film composite membranes for forward osmosis, J. Membr. Sci. 483
by Alberta Innovates Technology Futures (AITF) and oil sands
(2015) 25–33, http://dx.doi.org/10.1016/j.memsci.2015.01.025.
companies (Suncor Energy Inc., Devon Canada Corporation, and [24] Y.H. Cho, J. Han, S. Han, M.D. Guiver, H.B. Park, Polyamide thin-film composite
ConocoPhillips Canada Resources Corporation) via the Grant membranes based on carboxylated polysulfone microporous support mem-
number RES 27789. branes for forward osmosis, J. Membr. Sci. 445 (2013) 220–227, http://dx.doi.
org/10.1016/j.memsci.2013.06.003.
[25] Y. Wang, T. Xu, Anchoring hydrophilic polymer in substrate: an easy approach
for improving the performance of TFC FO membrane, J. Membr. Sci. 476 (2015)
References 330–339, http://dx.doi.org/10.1016/j.memsci.2014.11.025.
[26] J.R. McCutcheon, M. Elimelech, Influence of membrane support layer hydro-
phobicity on water flux in osmotically driven membrane processes, J. Membr.
[1] S. Zhao, L. Zou, C.Y. Tang, D. Mulcahy, Recent developments in forward os- Sci. 318 (2008) 458–466, http://dx.doi.org/10.1016/j.memsci.2008.03.021.
mosis: opportunities and challenges, J. Membr. Sci. 396 (2012) 1–21, http://dx. [27] I.L. Alsvik, K.R. Zodrow, M. Elimelech, M.B. Hägg, Polyamide formation on a
doi.org/10.1016/j.memsci.2011.12.023. cellulose triacetate support for osmotic membranes: effect of linking mole-
[2] K. Lutchmiah, A.R.D. Verliefde, K. Roest, L.C. Rietveld, E.R. Cornelissen, Forward cules on membrane performance, Desalination 312 (2013) 2–9, http://dx.doi.
osmosis for application in wastewater treatment: a review, Water Res. 58 org/10.1016/j.desal.2012.09.019.
(2014) 179–197, http://dx.doi.org/10.1016/j.watres.2014.03.045. [28] X. Li, K.Y. Wang, B. Helmer, T.S. Chung, Thin-film composite membranes and
[3] T.Y. Cath, A.E. Childress, M. Elimelech, Forward osmosis: principles, applica- formation mechanism of thin-film layers on hydrophilic cellulose acetate
tions, and recent developments, J. Membr. Sci. 281 (2006) 70–87, http://dx.doi. propionate substrates for forward osmosis processes, Ind. Eng. Chem. Res. 51
org/10.1016/j.memsci.2006.05.048. (2012) 10039–10050, http://dx.doi.org/10.1021/ie2027052.
[4] T.S. Chung, X. Li, R.C. Ong, Q. Ge, H. Wang, G. Han, Emerging forward osmosis [29] C. Klaysom, S. Hermans, A. Gahlaut, S. Van Craenenbroeck, I.F.J. Vankelecom,
(FO) technologies and challenges ahead for clean water and clean energy Polyamide/Polyacrylonitrile (PA/PAN) thin film composite osmosis mem-
applications, Curr. Opin. Chem. Eng. 1 (2012) 246–257, http://dx.doi.org/ branes: film optimization, characterization and performance evaluation, J.
10.1016/j.coche.2012.07.004. Membr. Sci. 445 (2013) 25–33, http://dx.doi.org/10.1016/j.
[5] K.Y. Wang, T.S. Chung, G. Amy, Developing thin-film-composite forward os- memsci.2013.05.037.
mosis membranes on the PES/SPSf substrate through interfacial polymeriza- [30] M. Yasukawa, S. Mishima, M. Shibuya, D. Saeki, T. Takahashi, T. Miyoshi,
tion, AIChE J. 58 (2012) 770–781, http://dx.doi.org/10.1002/aic. T. Miyoshi, H. Matsuyama, Preparation of a forward osmosis membrane using
[6] I. Pinnau, B.D. Freeman, Formation and modification of polymeric membranes: a highly porous polyketone microfiltration membrane as a novel support, J.
overview, in: I. Pinnau, B.D. Freeman (Eds.), Membrane Formation and Mod- Membr. Sci. 487 (2015) 51–59, http://dx.doi.org/10.1016/j.
ification, American Chemical Society, Washington, DC, 1999, memsci.2015.03.043.
pp. 1–22, http://dx.doi.org/10.1021/bk-2000-0744. [31] G. Han, T.S. Chung, M. Toriida, S. Tamai, Thin-film composite forward osmosis
[7] M. Qasim, N.A. Darwish, S. Sarp, N. Hilal, Water desalination by forward (di- membranes with novel hydrophilic supports for desalination, J. Membr. Sci.
rect) osmosis phenomenon: a comprehensive review, Desalination 374 (2015) 423–424 (2012) 543–555, http://dx.doi.org/10.1016/j.memsci.2012.09.005.
47–69, http://dx.doi.org/10.1016/j.desal.2015.07.016. [32] N. Widjojo, T.-S. Chung, M. Weber, C. Maletzko, V. Warzelhan, A sulfonated
[8] J.T. Arena, B. McCloskey, B.D. Freeman, J.R. McCutcheon, Surface modification polyphenylenesulfone (sPPSU) as the supporting substrate in thin film com-
of thin film composite membrane support layers with polydopamine: en- posite (TFC) membranes with enhanced performance for forward osmosis
abling use of reverse osmosis membranes in pressure retarded osmosis, J. (FO), Chem. Eng. J. 220 (2013) 15–23, http://dx.doi.org/10.1016/j.
Membr. Sci. 375 (2011) 55–62, http://dx.doi.org/10.1016/j.memsci.2011.01.060. cej.2013.01.007.
[9] C.H. Tan, H.Y. Ng, A novel hybrid forward osmosis – nanofiltration (FO–NF) [33] D. Emadzadeh, W.J. Lau, T. Matsuura, M. Rahbari-Sisakht, A.F. Ismail, A novel
process for seawater desalination: draw solution selection and system con- thin film composite forward osmosis membrane prepared from PSf–TiO2
figuration, Desalin. Water Treat. 13 (2010) 356–361, http://dx.doi.org/10.5004/ nanocomposite substrate for water desalination, Chem. Eng. J. 237 (2014)
dwt.2010.1733. 70–80, http://dx.doi.org/10.1016/j.cej.2013.09.081.
[10] N.Y. Yip, A. Tiraferri, W.A. Phillip, J.D. Schiffman, M. Elimelech, High [34] P. Lu, S. Liang, L. Qiu, Y. Gao, Q. Wang, Thin film nanocomposite forward
B. Khorshidi et al. / Journal of Membrane Science 511 (2016) 29–39 39

osmosis membranes based on layered double hydroxide nanoparticles blen- memsci.2015.11.015.


ded substrates, J. Membr. Sci. 504 (2015) 196–205, http://dx.doi.org/10.1016/j. [51] Handbook of Chemistry and Physics, in: W.M. Haynes (Ed.), 95th ed.,CRC Press,
memsci.2015.12.066. 2014.
[35] X. Song, Z. Liu, D.D. Sun, Nano gives the answer: breaking the bottleneck of [52] S. Loeb, Effect of porous support fabric on osmosis through a Loeb-Sourirajan
internal concentration polarization with a nanofiber composite forward os- type asymmetric membrane, J. Membr. Sci. 129 (1997) 243–249, http://dx.doi.
mosis membrane for a high water production rate, Adv. Mater. 23 (2011) org/10.1016/S0376-7388(96)00354-7.
3256–3260, http://dx.doi.org/10.1002/adma.201100510. [53] K.L. Lee, R.W. Baker, H.K. Lonsdale, Membranes for power generation by
[36] N.N. Bui, M.L. Lind, E.M.V. Hoek, J.R. McCutcheon, Electrospun nanofiber pressure-retarded osmosis, J. Membr. Sci. 8 (1981) 141–171.
supported thin film composite membranes for engineered osmosis, J. Membr. [54] B. Khorshidi, T. Thundat, B.A. Fleck, M. Sadrzadeh, A novel approach toward
Sci. 385–386 (2011) 10–19, http://dx.doi.org/10.1016/j.memsci.2011.08.002. fabrication of high performance thin film composite polyamide membranes,
[37] N.N. Bui, J.R. McCutcheon, Hydrophilic nanofibers as new supports for thin Sci. Rep. (2016), http://dx.doi.org/10.1038/srep22069.
film composite membranes for engineered osmosis, Environ. Sci. Technol. 47 [55] M. Oldani, G. Schock, Characterization of ultrafiltration membranes by infra-
(2013) 1761–1769, http://dx.doi.org/10.1021/es304215g. red spectroscopy, esca, and contact angle measurements, J. Membr. Sci. 43
[38] M. Tian, C. Qiu, Y. Liao, S. Chou, R. Wang, Preparation of polyamide thin film (1989) 243–258, http://dx.doi.org/10.1016/S0376-7388(00)85101-7.
composite forward osmosis membranes using electrospun polyvinylidene [56] B.C. Smith, Infrared Spectral Interpretation: A Systematic Approach, CRC Press,
fluoride (PVDF) nanofibers as substrates, Sep. Purif. Technol. 118 (2013) Boca Raton, Florida, USA, 1998.
727–736, http://dx.doi.org/10.1016/j.seppur.2013.08.021. [57] C. Tang, Y. Kwon, J. Leckie, Probing the nano- and micro-scales of reverse
[39] L. Huang, J.T. Arena, J.R. McCutcheon, Surface modified PVDF nanofiber sup- osmosis membranes  a comprehensive characterization of physiochemical
ported thin film composite membranes for forward osmosis, J. Membr. Sci. properties of uncoated and coated membranes by XPS, TEM, ATR-FTIR, and
499 (2016) 352–360, http://dx.doi.org/10.1016/j.memsci.2015.10.030. streaming potential measurements, J. Membr. Sci. 287 (2007) 146–156, http:
[40] Y. Wang, X. Li, C. Cheng, Y. He, J. Pan, T. Xu, Second interfacial polymerization //dx.doi.org/10.1016/j.memsci.2006.10.038.
on polyamide surface using aliphatic diamine with improved performance of [58] Y. Song, F. Liu, B. Sun, Preparation, characterization, and application of thin
TFC FO membranes, J. Membr. Sci. 498 (2016) 30–38, http://dx.doi.org/10.1016/ film composite nanofiltration membranes, J. Appl. Polym. Sci. 95 (2005)
j.memsci.2015.09.067. 1251–1261, http://dx.doi.org/10.1002/app.21338.
[41] A. Tiraferri, Y. Kang, E.P. Giannelis, M. Elimelech, Highly hydrophilic thin-film [59] C.Y. Tang, Y.N. Kwon, J.O. Leckie, Effect of membrane chemistry and coating
composite forward osmosis membranes functionalized with surface-tailored
layer on physiochemical properties of thin film composite polyamide RO and
nanoparticles, ACS Appl. Mater. Interfaces 4 (2012) 5044–5053, http://dx.doi.
NF membranes II. Membrane physiochemical properties and their depen-
org/10.1021/am301532g.
dence on polyamide and coating layers, Desalination 242 (2009) 168–182.
[42] N. Ma, J. Wei, R. Liao, C.Y. Tang, Zeolite-polyamide thin film nanocomposite
[60] D. Quéré, Wetting and roughness, Annu. Rev. Mater. Res. 38 (2008) 71–99,
membranes: towards enhanced performance for forward osmosis, J. Membr.
http://dx.doi.org/10.1146/annurev.matsci.38.060407.132434.
Sci. 405–406 (2012) 149–157, http://dx.doi.org/10.1016/j.memsci.2012.03.002.
[61] G. Hurwitz, G.R. Guillen, E.M.V. Hoek, Probing polyamide membrane surface
[43] M. Amini, M. Jahanshahi, A. Rahimpour, Synthesis of novel thin film nano-
charge, zeta potential, wettability, and hydrophilicity with contact angle
composite (TFN) forward osmosis membranes using functionalized multi-
measurements, J. Membr. Sci. 349 (2010) 349–357, http://dx.doi.org/10.1016/j.
walled carbon nanotubes, J. Membr. Sci. 435 (2013) 233–241, http://dx.doi.org/
memsci.2009.11.063.
10.1016/j.memsci.2013.01.041.
[62] C.H. Tan, H.Y. Ng, Modified models to predict flux behavior in forward osmosis
[44] N. Niksefat, M. Jahanshahi, A. Rahimpour, The effect of SiO2 nanoparticles on
in consideration of external and internal concentration polarizations, J.
morphology and performance of thin film composite membranes for forward
osmosis application, Desalination 343 (2014) 140–146, http://dx.doi.org/ Membr. Sci. 324 (2008) 209–219, http://dx.doi.org/10.1016/j.
10.1016/j.desal.2014.03.031. memsci.2008.07.020.
[45] L. Shen, S. Xiong, Y. Wang, Graphene oxide incorporated thin-film composite [63] N.T. Hancock, T.Y. Cath, Solute coupled diffusion in osmotically driven mem-
membranes for forward osmosis applications, Chem. Eng. Sci. 143 (2016) brane processes, Environ. Sci. Technol. 43 (2009) 6769–6775, http://dx.doi.
194–205, http://dx.doi.org/10.1016/j.ces.2015.12.029. org/10.1021/es901132x.
[46] J.W. Krumpfer, T. Schuster, M. Klapper, K. Müllen, Make it nano – keep it nano, [64] A.K. Ghosh, E.M.V. Hoek, Impacts of support membrane structure and chem-
Nano Today 8 (2013) 417–438, http://dx.doi.org/10.1016/j.nantod.2013.07.006. istry on polyamide–polysulfone interfacial composite membranes, J. Membr.
[47] M.M. Pendergast, E.M.V. Hoek, A review of water treatment membrane na- Sci. 336 (2009) 140–148, http://dx.doi.org/10.1016/j.memsci.2009.03.024.
notechnologies, Energy Environ. Sci. 4 (2011) 1946–1971, http://dx.doi.org/ [65] P.S. Singh, S.V. Joshi, J.J. Trivedi, C.V. Devmurari, A.P. Rao, P.K. Ghosh, Probing
10.1039/c0ee00541j. the structural variations of thin film composite RO membranes obtained by
[48] A.K. Ghosh, B.-H. Jeong, X. Huang, E.M.V. Hoek, Impacts of reaction and curing coating polyamide over polysulfone membranes of different pore dimensions,
conditions on polyamide composite reverse osmosis membrane properties, J. J. Membr. Sci. 278 (2006) 19–25, http://dx.doi.org/10.1016/j.
Membr. Sci. 311 (2008) 34–45, http://dx.doi.org/10.1016/j.memsci.2007.11.038. memsci.2005.10.039.
[49] T.Y. Cath, M. Elimelech, J.R. McCutcheon, R.L. McGinnis, A. Achilli, D. Anastasio, [66] M.A. Al Mamun, M. Sadrzadeh, R. Chatterjee, S. Bhattacharjee, S. De, Colloidal
A.R. Brady, A.E. Childress, I.V. Farr, N.T. Hancock, J. Lampi, L.D. Nghiem, M. Xie, fouling of nanofiltration membranes: a novel transient electrokinetic model
Yip, Ngai Yin, Standard methodology for evaluating membrane performance and experimental study, Chem. Eng. Sci. 138 (2015) 153–163, http://dx.doi.
in osmotically driven membrane processes, Desalination 312 (2013) 31–38, org/10.1016/j.ces.2015.08.022.
http://dx.doi.org/10.1016/j.desal.2012.07.005. [67] M. Hayatbakhsh, M. Sadrzadeh, D. Pernitsky, S. Bhattacharjee, J. Hajinasiri,
[50] B. Khorshidi, J. Hajinasiri, G. Ma, S. Bhattacharjee, M. Sadrzadeh, Thermally Treatment of an in situ oil sands produced water by polymeric membranes,
resistant and electrically conductive PES/ITO nanocomposite membrane, J. Desalin. Water Treat. (2015) 1–19, http://dx.doi.org/10.1080/
Membr. Sci. 500 (2016) 151–160, http://dx.doi.org/10.1016/j. 19443994.2015.1069216.

You might also like