Sie sind auf Seite 1von 12

Journal of Membrane Science 521 (2017) 53–64

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Understanding functionalized silica nanoparticles incorporation in thin


film composite membranes: Interactions and desalination performance
Masoumeh Zargar, Yusak Hartanto, Bo Jin n, Sheng Dai n
School of Chemical Engineering, The University of Adelaide, SA 5005, Australia

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

Article history: Incorporation of functional nanoparticles in polymer membranes is a promising approach for the de-
Received 29 March 2016 velopment of advanced thin film nanocomposite (TFN) membranes with improved desalination perfor-
Received in revised form mance. This study was aimed to fabricate TFN membranes incorporated with functionalized silica na-
26 August 2016
noparticles (SNs), which were synthesized with different sizes (  50 and  100 nm) and surface func-
Accepted 31 August 2016
Available online 31 August 2016
tionalities (hydroxyl, epoxy and amine). Research focused on getting a better understanding of the in-
terfacial interactions between SNs and polymer matrix and structure-property correlation of the TFN
Keywords: membranes. The physicochemical properties of the functionalized SNs and the TFN membranes were
Thin film nanocomposite membrane characterized using ATR-FTIR, TGA, DLS, SEM, TEM, XPS, tensiometer and streaming potential analysis.
Surface functionalization
The results confirm successful incorporation of the surface functionalized SNs into the polymer mem-
Water flux
branes. The concentration, size and surface functionality of SNs were found to have strong impact on the
Salt rejection
Structure-property correlation surface hydrophilicity, morphology and chemistry of the TFN membranes. The SNs hybridized TFN
members showed increased permeate flux and insignificant change in salt rejection. The surface func-
tionalization of SNs with amine and epoxy moieties could facilitate the chemical interaction between SNs
and polymer monomers, resulting in stabilizing the TFN membranes. Our research outcomes have
provided new insight into the structure-performance correlation of TFN membranes and can be bene-
ficial for fabrication of a wide range of nanoparticle incorporated membranes.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction groups: zero-dimensional metal and metal oxides, one-dimen-


sional nanotubes, and two-dimensional porous NPs [7]. The most
Polymeric thin film composite (TFC) membranes have received commonly used NPs in TFN membranes include zeolite [5,8,9],
increasing research interests and applications in separation and silica [10–14], titanium dioxide [15,16], silver [17,18], graphene
purification processes such as desalination and wastewater treat- oxide [19–21] and carbon nanotubes [6,22,23].
ment. TFC membranes demonstrate higher separation perfor- The TFN membranes can be fabricated by either direct de-
mance and stability over a wider range of temperature and pH position of NPs on the TFC membranes or encapsulation of NPs
compared with the conventional cellulose acetate membranes [1]. into the TFC membrane during the interfacial polymerization
Although many studies have been offered to the development of stage. Surface deposition is a post modification approach and is
TFC membranes during last decades, there is still an on-going need mainly used to improve the antifouling capability of the mem-
to develop technically and economically more feasible TFC mem- branes. However, the deposited NPs might leak during the high-
pressure filtration process. Hence, it is necessary to improve the
branes with a high separation performance and low energy costs
stability of NPs in the polymer matrix by chemical functionaliza-
for their industrial applications [2,3].
tion strategies such as NP surface modification. Further, the gen-
Recent studies showed that the integration of nanoparticle (NP)
eration of stress weak points due to the possible NP aggregation is
with the TFC membranes to fabricate thin film nanocomposite
another issue associated with the fabrication of TFN membranes.
(TFN) membranes can improve the membranes’ physicochemical
These may initiate the formation of defects during the high-
properties such as hydrophilicity, mechanical stability, and ther-
pressure filtration process. Incorporation of NPs into the TFC
mal resistance, as well as their permselectivity [4–6]. The NPs used
membrane during the interfacial polymerization process can be a
as nanofillers in the TFN membranes can be classified into three promising approach. Jeong et al. fabricated the first TFN mem-
brane by in-situ incorporation of zeolite-A NPs (50–200 nm) into
n
Corresponding authors. the polyamide (PA) layer and their zeolite-incorporated TFN
E-mail addresses: bo.jin@adelaide.edu (B. Jin), s.dai@adelaide.edu.au (S. Dai). membranes showed an enhanced separation performance and

http://dx.doi.org/10.1016/j.memsci.2016.08.069
0376-7388/& 2016 Elsevier B.V. All rights reserved.
54 M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64

fouling resistance [8]. The sustainable improvement in the struc- temperature. 2.9 ml of TEOS was rapidly injected into the mixture
ture and performance of TFN membranes is highly dependent on and the solution was stirred at 25 °C for 24 h. The resulting SNs
the NP stability and distribution in the membrane matrix [4]. The were washed 3–4 times with ethanol and then redispersed and
overall stability of these NPs incorporated into the TFC membrane stored in ethanol.
through interfacial polymerization is expected to be higher than SNs with the average size of  50 nm were synthesized ac-
those being incorporated via deposition approach. That is due to a cording to a modified approach reported by Fouilloux et al. [30].
large quantity of NPs which are partially or fully encapsulated 30 ml of 6 mM L-arginine solution was prepared. 4.65 ml cyclo-
inside the PA thin film layers during the interfacial polymerization hexane was then added as the solvent for the organic phase. The
process. mixture was transferred to a pre-heated 60 °C oil bath to warm up,
Silica nanoparticles (SNs) have superior biocompatibility and followed by the dropwise injection of 5.7 ml TEOS to the organic
mechanical stability for the fabrication of TFN membranes, and phase. The stirring rate of the solution was fixed at a low value
their surface can be easily modified with functional chemical such that the aqueous phase was well mixed while the organic
groups [11,24]. Further, SNs show low toxicity as evidenced by phase was left intact. The reaction was further continued for 72 h
their wide biomedical applications [25–27]. Importantly, SN sur- under constant stirring and heating. Finally, bluish silica solution
face hydrophilic groups can be beneficial for enhancing perme- was washed several times with Milli-Q water and stored in etha-
ability and fouling resistance of the TFN membranes [4,10]. The nol. The obtained NPs possess hydroxyl surface functional groups
prospect of enhancing surface hydrophilicity and consequently and are noted as SN-OH. The two SN-OH NPs with sizes of 50
fouling resistance of the TFC membranes has been well docu- and  100 nm are named as SN50 and SN100 in this study. Further
mented in previous studies [12,28]. For instance, Tiraferri et al. chemical modifications were then performed to introduce amine
studied the incorporation of SNs (  14 nm) on the TFC membranes or epoxy groups to the SNs.
via the deposition approach, resulting in increasing the fouling
resistance of their TFN membranes [12]. 2.2.2. Preparation of amine and epoxy-functionalized silica
Until now, there are very few reported studies on the in-situ nanoparticles
incorporation of NPs and their interactions within the polymer A modified approach from Chang et al. was applied to func-
matrix of the TFN membranes. Further, the extensive evaluation of tionalize the SN surface with amine groups [31]. Typically, 100 mg
the physicochemical properties of TFN membranes following the of the synthesized SNs were dispersed in 50 ml of ethanol and
hybridization of functionalized NPs and their performance corre- degassed with nitrogen for 1 h. This was followed by the increase
lations are missing in the literature. This study aims to address of the temperature to 80 °C and the quick injection of 135 mL
these fundamental gaps by developing novel TFN membranes in- APTMS to the reaction vessel. The solution was then refluxed for
corporated with functionalised SNs, which are noted as “SN in- 6 h. After centrifugation and washing with ethanol and water,
corporated TFC” (SN-TFC) membranes in this study. Our research amine-functionalized SNs (SN-NH2) were obtained and finally re-
will focus on understanding the structure-property correlation of dispersed in ethanol.
the TFN membranes. In particular, we will synthesize SNs with The epoxy group was introduced to the SN surface according to
different sizes (  50 and  100 nm) and functionalize their sur- an established approach [32]. In brief, 100 mg of synthesized SNs
faces with the amine or epoxy groups. These functionalities make was added into 2.5 ml of 5 wt% GPMS in anhydrous toluene and
the SNs ready to interact with the PA monomers during the in- then stirred for 12 h at room temperature. The epoxy functiona-
terfacial polymerization and can improve the stability of the TFN lized SNs (SN-EPX) were centrifuged and washed with anhydrous
membranes. The surface functionalization of the SNs and the toluene and absolute ethanol, respectively, followed by drying
physicochemical properties of the TFN membranes are extensively under vacuum. The general schematics for the preparation of SN-
characterized. The desalination performance of the SN-TFC mem- NH2 and SN-EPX are illustrated in Supporting information (Fig.
branes is evaluated using a cross-flow filtration unit. The research S1).
outcomes will provide new insights into the influence of different
surface functionality of nanofillers on the overall performance and 2.3. Fabrication of TFC and SN-TFC membranes
physicochemical properties of the resulting TFN membranes.
To obtain the TFC membranes, an ultra-thin PA layer was casted
on ultrafiltration PES support membranes using the interfacial
2. Experimental polymerization. The PES support was first clamped between an
acrylic plastic plate and an acrylic plastic frame. Then, 2% (w/v)
2.1. Materials MPD aqueous solution was poured into the frame and allowed to
penetrate to the PES support for 10 min. Excess MPD solution was
Polyether sulfone support (PES-MQ-50 kDa) was obtained from drained and the PES support was rolled with a soft rubber roller
Synder filtration (USA). 1, 3-phenylenediamine (MPD, 499%), 1, 3, until no visible droplets remained on the surface of the membrane.
5-benzenetricarbonyl trichloride (TMC, 98%), n-hexane, tetraethyl This was followed by reassembling the holder and pouring
orthosilicate (TEOS, 499%), L-arginine, (3-aminopropyl) tri- 0.1 (w/v)% TMC in hexane solution into the frame. After 1 min
methoxysilane (APTMS, 97%) and (3-glycidyloxypropyl) tri- reaction, excess TMC solution was disposed and the membrane
methoxysilane (GPMS, 498%) were purchased from Sigma-Al- was air-dried for another 1 min, followed by 10 min curing in a
drich. Sodium chloride (NaCl), aqueous ammonia solution (28%) recirculating oven at 60 °C. The fabricated TFC membranes
and cyclohexane were obtained from VWR International. were washed with 22 °C distilled water and stored wet at 5 °C for
further characterization and evaluation. The TFC membrane
2.2. Synthesis and surface modification of silica nanoparticles without any nanoparticle was used as a control membrane in this
study.
2.2.1. Synthesis of silica nanoparticles SN-TFC membranes were fabricated exactly as described, ex-
SNs with an average size of  100 nm were fabricated accord- cept that different amounts of functionalized SNs (SN-OH, SN-NH2,
ing to a modified Stöber approach [29]. Briefly, 50 ml absolute and SN-EPX) were dispersed in MPD aqueous solution. To obtain
ethanol, 0.9 ml Milli-Q water and 2.04 ml of 28% aqueous ammo- good SN dispersion, the mixtures of SNs and MPD solution were
nia were mixed together and stirred for 10 min at room sonicated for 1 h at 22 °C exactly before the interfacial
M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64 55

polymerization. The resulting membranes incorporated with SN- energy of 1486.6 eV. The area of analysis (Iris aperture) was
OH, SN-NH2 and SN-EPX NPs are noted as SN-OH-TFC, SN-NH2-TFC 0.3 mm  0.7 mm and the electron take off angle was normal to
and SN-EPX-TFC membranes, respectively. the sample surface with the analysis depth of approximately
15 nm from the sample surface. The spectra were scanned in the
range of  10 to 1110 eV and the pass energy and step of 160 eV
2.4. Characterization of nanoparticles, TFC and SN-TFC membranes
and 500 meV were used. The charge neutraliser system was used
to reduce the surface charge of the samples. CasaXSP software
Attenuated total reflectance Fourier transform infrared spec-
package was eventually used to interpret the obtained survey
troscopy (ATR-FTIR) was performed using a NICOLET 6700 spec-
spectra.
trometer equipped with a diamond ATR to evaluate the chemical
Contact angle measurements were performed on an Attention
structures and functionalities of the fabricated nanoparticles and
Theta Optical tensiometer in sessile drop mode at room tem-
membranes. The obtained FTIR spectra are shifted vertically for
perature. The membranes were attached to clean glass slides using
clarity.
a double-sided tape and a milli-Q water droplet (1 mL) was auto-
The surface charges of the synthesized nanoparticles were
matically dispensed on the surface of the membranes. A FireWire
measured using a Malvern Zetasizer Nano ZS (Malvern Inst. Ltd., U.
camera (55 mm focus length) recorded the droplet surface contact
K.) based on the dynamic light scattering (DLS). The samples were
angle equilibrated for 10 s from the time it was dispensed. Curve
dispersed in milli-Q water and the measurements were performed
fitting and data analysis were then performed using the OneAt-
at room temperature.
tension software. The reported contact angles were the average of
Thermogravimetric analysis (TGA) was performed on a TGA/
at least 8 different positions on each membrane.
DSC 2 STARe system to estimate the surface functionalities, where
5 mg dried nanoparticles were loaded and measurement tem-
perature was from 30 to 800 °C with a heating rate of 10 °C/min 2.5. Evaluation of membrane desalination performance
under air. An isotherm step at 120 °C was set for 1 h to remove the
trace amount of adsorbed water. The residual weight of the sam- A reverse osmosis cross-flow filtration setup was used to
ples at the end of this isotherm was considered as the initial evaluate the desalination performance of the TFC and SN-TFC
weight and used to calculate the normalized residual weights of membranes. The effective membrane surface area of the filtration
the following stages. cell (CF042, Sterlitech) was 42 cm2. Prior to evaluation, the
The morphology of fabricated nanoparticles and membranes membrane coupons were compacted with distilled water in the
were characterized using a scanning electron microscope (FEI filtration unit for at least 16 h under the pressure of 15 bar to reach
Quanta 450 FEG Environmental SEM (ESEM)) equipped with EDX their equilibrium flux. The system was then operated with 15 L of
(energy dispersive x-ray spectroscopy). The instrument was op- 2000 mg/L NaCl at a flowrate of 4 L/min under 15 bar and
erated at the voltage of 5–20 kV and the working distance of 257 0.1 °C. The concentrate was recycled back to the feed tank
10 mm. For the cross-section SEM imaging of the membrane while operating the system, and the permeate was collected in a
samples, small pieces of fabric free and wet membranes were container positioned on a digital balance for flux measurement.
freeze-fractured in liquid nitrogen and mounted vertically on SEM The schematic illustration of this lab-scale filtration setup is pre-
stands. The prepared samples were dried overnight and sputter sented in Supporting Information (Fig. S2).
coated with  5 nm layer of platinum using a CRESSINGTON 208 The collected permeate weight was recorded using the LabX
high-resolution sputter coater before imaging. Direct Software (Mettler Toledo) and the volumetric permeate flux
TEM images were taken on an FEI Tecnai G2 Spirit Transmission (Jw) was calculated using Eq. (1) accordingly:
Electron Microscope at an acceleration voltage of 100 kV.
Nanoparticle samples were prepared by placing one drop of VP
Jw =
ethanol dispersed nanoparticles on 200 mesh carbon-coated A⋅t (1)
copper grids. Membrane samples for cross-section TEM imaging where Jw is the permeate water flux (LMH), VP is the collected
were prepared by gently removing the polyester backing layer to permeate volume (l), A is the active surface area of the membrane
ensure PA and PSF layers remained together. Small pieces (m2) and t is the operation time (h). Intrinsic water permeability
of fabric free membranes were cut and embedded in Epon resin coefficients of the fabricated membranes can be then calculated
and cured in the oven at 50 °C overnight. From the cured samples, from the obtained Jw values using Eq. (2):
60–70 nm cross sections were cut on a Leica EM UC6 Ultra-
microtome and mounted on formvar and carbon coated copper Jw
Pw =
grids for imaging. (∆P−∆π ) (2)
An Anton Paar SurPASS Electrokinetic Analyser (Anton Paar 2 1 1
where Pw (L m h bar ) is the specific membrane perme-
GmbH) was used to evaluate the streaming potentials of the fab-
ability coefficient, ΔP (bar) is the applied transmembrane pressure
ricated membranes. The wet membranes samples were cut in
and Δπ (bar) is the transmembrane osmotic pressure [9].
rectangular profiles and mounted in the adjustable gap cell ap-
In order to evaluate the salt rejection, the conductivity of the
paratus by applying double- sided tape. The measurements were
permeate and feed water were measured using a conductometer
carried out using 10 mM KCl solution as the electrolyte and the pH
(k ¼1.0, AQUA, Cond./pH, TPS, Australia) and were subsequently
was varied by the addition of 0.1 M HCl or 0.1 M NaOH solutions to
converted to the equivalent salt concentrations using a pre-cali-
400 ml KCl solution. Multiple rinsing and flow check steps were
bration curve which is shown in Fig. S3. Salt rejections (R) were
carried out prior to the tests to ensure reproducibility. For each
then calculated using Eq. (3):
sample, four repeats were recorded and the averages in each pH
were reported. Zeta potentials were obtained from the slope of ⎛ C ⎞
streaming potential versus pressure according to the Helmholtz- R%=⎜ 1− P ⎟× 100
⎝ CF ⎠ (3)
Smoluchowski approach.
X-ray photoelectron spectroscopy (XPS) was performed on where R is the salt rejection and CP and CF (g/L) are the equivalent
the samples using a Kratos Axis Ultra with a monochromatic NaCl concentrations in the collected permeate and feed,
aluminium x-ray running at 225 W and the characteristic respectively.
56 M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64

The solute diffusive permeation coefficients, B (L m  2 h  1), 3. Results and discussion


were then calculated using Eq. (4):
3.1. Characterization of fabricated silica nanoparticles

B = Jw
( 1−R)
(4) SEM micrographs and TEM images (inset) of the synthesized
R
SNs in Fig. 1 show that the nanoparticles have uniform spherical
structures with the average diameters of  50 and  100 nm, re-
The desalination tests for each TFC and SN-TFC membrane were
spectively. SEM and TEM images of the surface functionalized SNs
performed in three replicates and the average values are reported
are presented in Fig. 1. The results confirm that there is no sig-
for further evaluation and comparison. A homoscedastic (assum-
nificant SN aggregation following their surface functionalization.
ing equal variances) two-tailed distribution using student t-test The thermogravimetric analysis was performed to study the
analysis was finally used to study the statistically significant dif- surface functionalization of SNs and to estimate the quantity of
ferences (α ¼0.05) between the performance results. their surface functional groups. Physically adsorbed water was

Fig. 1. SEM micrographs and TEM images (inset) of the different sized SN, A) SN50, B) SN100, C) SN50-NH2, D) SN100-NH2, E) SN50-EPX and F) SN100-EPX. Particle size and
morphology are similar as the SN before modification.
M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64 57

removed by heating the dried SNs at 120 °C for 1 h. The tem- confirming PA fabrication [36]. Since the IR penetration depth
perature was then raised step-wise to 800 °C. The TGA thermo- ( 0.5  1 mm) exceeds the thickness of the thin PA layer [36,37],
grams in Fig. 2A show that SN50 was reduced about 2.5% residual PA, PES and SN peaks are all emerged in these FTIR spectra. The
weight which can be associated with the loss of the chemically vibration bands of 955 cm  1 (stretching vibration of Si–OH) and
bonded water and organics from the sol-gel fabrication process or 1090 cm  1 (asymmetric vibration of Si–O–Si) represent the SNs
the dihydroxylation of the SNs’ surface hydroxyl groups [33,34]. and the vibration bands of the SN surface functional groups
The residual weight loss is higher for the surface functionalized overlap with the PES and PA bands. To further identify the over-
SNs due to the decomposition of grafted silane moieties and in- lapped peaks, the SN, PA and PA-SN thin films extracted from their
dicates the successful surface functionalization with amine or PES supports were characterized. The results are reported in Fig.
epoxy [34]. The estimated surface functional groups on the SN50- S6.
NH2 and SN50-EPX are determined as  6% and  9%, respectively. Fig. 3 presents the high and low magnification surface and
The successful surface modification by amine or epoxy groups cross-section SEM micrographs of the synthesized TFC and SN-TFC
can be further confirmed by zeta potential measurement. Fig. S4 membranes incorporated with the 0.05 wt% SNs. All membranes
shows the surface zeta potential of different SNs. The amine in- show a leaf-like structure, which is the expected surface mor-
troduction on the SN-NH2 surface results in changing the zeta phology of thin PA layers made from MPD and TMC. The SN50-TFC
potentials from negative to positive values. On the other hand, the membrane represents leafier surface morphology compared to the
SN-EPX shows a lower magnitude of surface potential due to the SN100-TFC membrane which has some lobe-like surface features
loss of some surface hydroxyl groups. It should be noted that the (Fig. 3B and E). These surface morphology and features are re-
zeta potential values (4 15 mV) indicate that the fabricated SNs sulted from the different sizes of SNs which may influence the
are highly stable in the aqueous medium. formation of the surface strands. As observed in the high magni-
Table 1 summarizes the XPS analysis results for the SN50 with fication surface images (Fig. 3C, D, F, and G), the membrane surface
different surface functionalities and their survey spectra are pre- features are denser and the polymer strands are more packed for
sented in Fig. S5. SN50 shows the typical silica peaks at 532.72 eV, the SN-NH2-TFC and SN-EPX-TFC membranes compared to the
284.72 eV, 153.3 eV and 103.22 eV corresponding to O 1s, C 1s, and control TFC and SN-OH-TFC membranes. These compact surface
Si 2s, and Si 2p, respectively [35]. For SN50-NH2, N 1s peak at features are due to the chemical interaction of silane groups with
399.62 eV is emerged with 1.92% composition which confirms the the PA monomers which may interfere with the interfacial poly-
successful amine functionalization. On the other hand, SN-EPX merization. However, high magnification SEM images of the SN-
shows 30.60% increase in the composition of C 1s which indicates TFC membranes indicates that all membranes still represent leaf-
the successful epoxy functionalization. like morphology with some scattered nodule-like features. The
presence of some SNs with the sizes of  50 and  100 nm on the
3.2. Characterization of TFC and SN-TFC membranes thin film layer of SN-TFC membranes confirms the successful in-
corporation of the functionalized SNs into the PA layers. Cross-
Fig. 2B shows ATR-FTIR spectra of PES support, TFC and SN-TFC section SEM images of the TFC and SN-TFC membranes (Fig. 3)
membranes. Two new peaks at 1541 and 1612 cm  1 are found in show the structure of the different membranes with maximum
the ATR-FTIR spectra of the TFC and SN-TFC membranes compared  450 nm PA layers. Furthermore, no significant aggregation of
to that of the PES support which is associated with the vibrations SNs can be observed, indicating the fair distribution of different
of amide groups: C–N (amide II band) and N–H (aromatic amide) SNs into the SN-TFC membranes.
Fig. S7 presents SEM images of the SN50-TFC membranes in-
Table 1 corporated with SN50 at different concentration. To further elu-
XPS results for SN50 with different surface functionalities. cidate the distribution of nanoparticles, SEM images were cap-
Sample name Atomic percentage of elements
tured from some fractured areas on the surface of SN50-NH2-TFC
and SN100-EPX-TFC membranes as illustrated in Fig. S8. The re-
N (%) O (%) C (%) Si (%) sults confirm the successful SN incorporation and fair distribution
in these membranes.
SN-OH – 59.53 21.17 19.30
To extensively analyse the SN distribution in the PA thin film
SN-NH2 1.92 56.14 21.68 20.26
SN-EPX – 54.18 27.65 18.17 layer, cross-section TEM imaging was performed on TFC/SN-TFC
membranes incorporated with functionalized SN50. Three cross-

PES support
Normalized residual weight loss

1
TFC membrane
Typical SN-TFC membrane
0.98 SN
Transmittance

0.96
SN-NH2

0.94 1612 1541

0.92 SN-EPX

0.9
0 100 200 300 400 500 600 700 800 900 1500 1000 500
Temperature (ºC) Wavenumbers (cm - ¹)
Fig. 2. A) TGA thermograms for SN50 before and after surface functionalization, B) FTIR spectra for PES support, TFC membrane and a typical SN-TFC membrane.
58 M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64

3 µm 500 nm 50 µm 5 µm
)

3 µm 500 nm 50 µm 5 µm

3 µm 500 nm 50 µm 5 µm

3 µm 500 nm 50 µm 5 µm

3 µm 500 nm 50 µm 5 µm

3 µm 500 nm 50 µm 5 µm

3 µm 500 nm 50 µm 5 µm

Fig. 3. SEM micrographs of the surfaces and cross-section of different TFC/SN-TFC membranes: A) TFC, B) SN50-TFC, C) SN50-NH2-TFC, D) SN50-EPX-TFC, E) SN100-TFC, F)
SN100-NH2-TFC, G) SN100-EPX-TFC. For the surface images, the low magnification images were taken at 55000  and the high magnification images were taken at
200000  and for the cross-section images, the low magnification images were at 2000  and the high magnification ones were at 25000  .
M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64 59

section samples were prepared from each membrane and their the MPD/TMC molar ratio is beneficial for forming a denser core
typical images are shown in Fig. 4. TEM images further confirm the PA layer, while reducing the thickness of the PA layer during the
successful SN distribution into the PA thin film layers. The quantity diffusion stage and vice versa [40–42]. Therefore, the SN-NH2-TFC
of SNs seems to be higher for the SN50-NH2-TFC and SN50-EPX- would have a less dense core layer but thicker secondary PA layer
TFC membranes compared to the SN50-TFC membranes which can formed by the diffusion, while the SN-EPX-TFC would have a
be attributed to the chemical interaction of SN-NH2 and SN-EPX denser core layer but thinner secondary PA layer.
with the PA monomers as well as the zeta potentials. These In order to understand the surface electrical double layers of
functionalized SNs could be incoportated to the TFC matrix due to the fabricated membranes, streaming potential measurements
the faverable interactions, including the electrostatic attraction were performed on the TFC and SN-TFC membranes incorporated
between SN-NH2 and negatively charged PES support [38], and the with functionalized SN50. Surface zeta potentials of the mem-
weaker electrostatic repulsion between SN-EPX and PES support branes can be highly influenced by the unreacted surface func-
than those between the SN-OH and PES support. Consistent with tional groups or impurities [43]. Therefore, the surface function-
previous studies [8,9,39], thickness of the PA thin film layers varies ality of the introduced SNs can alter the surface potential of the
in a range of 50  450 nm. The average thicknesses based on TEM SN-TFC membranes. Fig. 5 shows the average of four pH depen-
are 255 7178, 154 752, 271 7125, and 102 747 nm for the TFC, dent zeta potentials for various membranes measured using tan-
SN50-TFC, SN50-NH2-TFC, and SN50-EPX-TFC membranes, re- gential streaming potential (TSP) method with the Fairbrother and
spectively. The higher thickness of the SN50-NH2-TFC membrane Mastin approximation. In general, the surfaces of TFC and SN-TFC
than that of the SN50-EPX-TFC membrane is attributed to their membranes are more positively charged at low pH due to the
different amine/acyl chloride molar ratios due to the interaction of proton ions’ adsorption and are more negatively charged at high
SN-NH2 or SN-EPX with PA monomers. The amine/acyl chloride pH due to the deprotonation of the PA monomers [12,20].
ratio is identical to the feed's MPD/TMC ratio for the SN-OH-TFC. Fig. 5 shows that the isoelectric point (IEP) of the TFC mem-
However, this ratio slightly increases for the SN-NH2-TFC due to brane occurs at pH  3.6 which is consistent with the previous
the introduction of more –NH2 groups from the NP surface. Con- studies [43,44]. However, the zeta potential and IEP of the mod-
versely, SN-EPX can consume some –NH2 groups of MPD mono- ified membranes are slightly changed which are influenced by the
mer, resulting in slightly decreasing the ratio. The interfacial SNs’ surface functional groups. In the case of the SN50-NH2-TFC
polymerization of MPD and TMC consists of an extremely fast membrane, the extra amine groups raise the amine/acyl chloride
stage which forms an ultrathin dense core layer of PA in the oil- molar ratio and make the membrane surface more positively
water interface and a slow growth stage, making a thicker layer of charged compared to the other membranes increasing the IEP to
PA by the monomers’ diffusion (mainly MPD) [40]. The decrease of pH  4.5. However, the surface potentials and IEPs of the SN50-TFC

Fig. 4. Cross-section TEM images of the A) TFC, B) SN50-TFC, C) SN50-NH2-TFC and SN50-EPX-TFC membranes; the scale bars are 100 nm.
60 M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64

20
Table 2
15 XPS results for TFC membrane and TFN membranes incorporated with different
Plain TFC
surface functionalized SN50.
Zeta potential (mV)

10 SN50-TFC
SN50-NH2-TFC Sample name Atomic percentage of elements
5 SN50-EPX-TFC
*
0 N (%) O (%) C (%) Si (%) N/O

-5 TFC 9.89 16.12 73.71 – 0.61


-10 SN-TFC 8.97 16.51 73.92 0.58 0.60
SN-NH2-TFC 9.58 16.85 72.59 0.96 0.67
-15 SN-EPX-TFC 9.19 17.09 72.44 1.26 0.69

-20
2 3 4 5 6 7 8 9 10
pH 80

Contact Angle (Degree)


Fig. 5. Comparison of the streaming potential measurements for the TFC mem- 70
branes with and without SN50 having different surface functionalities. 60
50
and SN50-EPX-TFC membranes become more negative due to the
hydroxyl and/or epoxy functional groups on SNs. It should be 40

noted that all zeta potential curves show amphoteric membrane 30


surfaces or surfaces with both acidic and basic functional groups 20
[43,45] and their variation can be a strong indicator of the suc-
10
cessful SN incorporation into the TFC membranes. Further, all
0
curves appear to be stable at pH over 7, indicating the complete
deprotonation of the carboxylic groups [20].
XPS analysis was performed to estimate the cross-linking
density of the PA layers by comparing their elemental composi-
Fig. 6. Contact angle measurement results for the TFC membranes incorporated
tions. Normally, a higher N/O ratio indicates a higher cross-linking with the different concentrations of SN50 and 0.05 wt% SN-NH2 and SN-EPX.
density of the PA layers [46–48]. The SNs incorporation introduces
more C, O or N (for SN-NH2 only) in SN-TFC matrix. Thus, the N/O to a high water flux during the desalination process. Therefore, the
ratio in SN-TFC membrane needs to be determined for the cross-
lower contact angle of all SN-TFC membranes compared with the
linking density evaluation [9,49]. XPS analysis results of SNs in
TFC membrane appears to be beneficial for enhancing their de-
Table 1 can be used to correct the elemental composition of each
salination performance.
element and calculate the real N/O ratio in the SN-TFC membranes,
High hydrophilicity of membranes is deemed advantageous for
which is noted as *N/O. This can be performed based on Si com-
fouling mitigation, where anti-fouling characteristic comes from
position of each particular TFN membrane. Table 2 shows the in-
the formation of hydration layers on hydrophilic surfaces. This
itial values of each constituent element and their corresponding
* hydration layer acts as a repulsive barrier against various fouling
N/O ratios.
matters, such as microorganisms, organic and inorganic materials,
Results in Table 2 explain that SN50-TFC membrane has a lower
* proteins, etc. [12,28,55].
N/O ratio compared to the control TFC membrane which suggests
a slight decrease of the PA layer's cross-linking density following
the SN50 incorporation, indicating that the unmodified SNs can 3.3. Desalination performance of TFC and SN-TFC membranes
alter the bulk PA structure. This finding agrees with the results
reported in previous studies [8,9,48]. The increase in *N/O ratio The desalination performance of TFC and SN-TFC membranes in
after the introduction of SN50-NH2 and SN50-EPX into the PA terms of water flux and salt rejection were systematically in-
layers leads to high cross-linking density. It is due to the chemical vestigated and calculated using Eqs. (1) and (3). To evaluate the
interactions of the functional SNs with the PA monomers and can desalination behavior in a more accurate and independent man-
be beneficial for enhancing the SNs’ stability in the PA matrix. That ner, the intrinsic water and salt permeability coefficients, which
can also explain why more SNs could be detected in the TEM are independent of the operating conditions, were then calculated
samples of the SN50-NH2-TFC and SN50-EPX-TFC membranes from Eqs. (2) and (4).
compared to the SN50-TFC membrane. These observations are
supported by the FTIR analysis results reported in Fig. S6. 3.3.1. Effect of nanoparticle concentration
Contact angle is a macroscopic manifestation of the molecular Fig. 7A shows the desalination results of the SN-TFC mem-
level surface-water interactions and represents the magnitude of branes incorporated with the different concentrations of SN50.
the affinity towards water or the level of hydrophilicity of the Previous studies on desalination performance of the TFC mem-
membranes [8,50–52]. The contact angle was calculated from the branes reported that the permeability and the salt rejection of the
average of the tangent lines to both sides of the dispensed water membranes are significantly influenced by the structure and
droplets on the membrane surface. Fig. 6 shows the average con- properties of their supporting layers and fabrication procedure
tact angle values of at least 8 random points on the surface of each (e.g. manual/automated coating) [10,37,56,57]. The TFC mem-
membrane. The contact angles decrease with increasing the con- branes are usually made by polysulfone ultrafiltration support. In
centration of SNs in the PA thin film layers which is consistent our study, we used PES support with different structure as shown
with the results of previous studies. [35,51–54]. Further, the in Fig. S9, as well as the manual coating procedure. Thus, our de-
equilibrium contact angle decreases for the amine and epoxy salination data cannot be comparable with those membranes that
functionalized SN-TFC membranes. Similar contact angle profiles are manufactured by polysulfone supports or are automatically
are observed from all SN-TFC membranes. It is interesting to note coated. Our TFC membranes perform very much similar salt re-
that high hydrophilicity of TFC and TFN membranes corresponds jections as those reported studies [10,58–61]. This study is to
M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64 61

Table 3
24 The average solution diffusion permeability coefficients of the water and salt across
100
Water Flux (LMH) Salt Rejection (%) the fabricated TFC/SN-TFC membranes. All membranes were incorporated with
98 0.05 wt% SNs unless specified otherwise.
22
Water Flux (L/m²h)

96

Salt Rejection (%)


20 Sample name Pw (L m  2 h  1 bar) B (L m  2 h  1)
94

18 92 TFC 1.116 70.045 2.0157 0.551


90 SN50-TFC (0.05 wt%) 1.5047 0.039 2.890 7 0.817
16 SN50-TFC (0.1 wt%) 1.5167 0.037 2.898 7 0.779
88
SN50-TFC (0.2 wt%) 1.3517 0.049 2.375 70.754
14 86 SN50-NH2-TFC 1.1537 0.037 0.984 7 0.276
84 SN50-EPX-TFC 1.239 70.091 2.155 70.378
12 SN100-TFC 1.3687 0.039 2.537 70.378
82
SN100-NH2-TFC 1.1177 0.040 1.2737 0.222
10 80 SN100-EPX-TFC 1.189 70.049 1.639 70.367
TFC membrane 0.05% SN50 0.1% SN50 0.2% SN50

TFC/TFN Membrane
incorporation of hydrophilic nanofillers into the TFC membranes
[14,62,63,65,66].
TFC membrane 0.05% SN50-TFC
26 SN-TFC membranes perform a slightly lower salt rejection than
0.1% SN50-TFC 0.20% SN50-TFC
TFC membrane, while the rejection values are found to be im-
24
proved following the concentration increase. However, this change
Water Flux (L/m²h)

22
appeases insignificant. This observation is in line with those data
20 reported in the literature, and indicates that no major defects are
18 formed in the PA structure of the SN-TFC membranes following the
SN concentration up to 0.2 wt% [5,10,65]. Considering the statis-
16
tically insignificant correlation (p¼ 0.1824–0.5345) of the B values
14 for TFC membranes incorporated with the 0.05–0.2 wt% SN50, it
12 can be concluded that the incorporation of SNs at low con-
centrations could significantly improve the water permeability of
10
0 100 200 300 400 500 600 700 800 900 1000 the membranes without sacrificing their salt rejections.
EDX mapping of the SN-TFC membranes can provide further
Compaction Time (min)
information showing the NP distribution. Fig. S10 presents Si
Fig. 7. A) the performance evaluation of the various TFC membranes incorporated mappings of the SN-TFC membranes incorporated with different
with the different concentrations of SN50, and B) water flux vs. compaction time of SN50 concentrations. All SN-TFC membrances show homogeneous
the TFC and SN-TFC membranes incorporated with the different concentrations of
SN distributions across the membranes, while slight aggregation
SN50.
was found on the SN-TFC membrane with 0.2 wt% SNs. These
observations justifie the desalination results as shown in Fig. 7A,
evaluate the structure-performance correlation of the TFC mem-
and agree with those reported data [13,14,62].
branes incorporated with the surface functionalized SNs. The re-
0.05 wt% SN50-TFC membrane performs a slightly low (only
lative desalination data obtained at the same fabrication and
0.7% lower), but insignificant (p ¼0.7472) reduction in water flux
evaluation procedure are valid rather than the absolute values.
and an insignificant (p ¼0.9932) change of salt rejection compared
The specific water and salt permeability coefficients of the TFC
to the 0.1 wt% SN50-TFC membrane. Considering the statistical
membranes incorporated with SN50 at different concentrations
analysis results, 0.05 wt% was determined as the optimum SN
are presented in Table 3. Fig. 7A shows the equivalent water flux
concentration, and thus 0.05 wt% SN50-TFC membrane was cho-
and salt rejection of these membranes. All experiments were
sen for further investigations.
performed in three replicates, and the average values are statisti-
Compaction resistance is an important parameter for evalua-
cally assessed. The incorporation of 0.1 wt% SN50 increases the
tion of the TFC membranes for industrial desalination application.
water permeability by  40% over the TFC membrane. Student t-
Incorporation of inorganic nanofillers such as SNs into the TFC
test analysis shows statistically significant correlations of water membranes is expected to improve the mechanical strength and
permeability coefficients of various SN50-TFC membranes and the eventually compaction resistance of the resulting TFN membranes.
TFC membrane with the p-values of 0.0004, 0.0003, and 0.0037 for Both ultrafiltration support and PA layer of TFC membranes ex-
the 0.05 wt%, 0.1 wt%, and 0.2 wt% SN50 incorporated TFC mem- perience irreversible compaction under high pressure which re-
branes, respectively. As the incorporated SN concentration in- sults in the permanent decrease of the water flux [67–69]. Fig. 7B
creases from 0.05 to 0.1 wt%, similar high water permeability va- shows the compaction rates of the SN-TFC membranes in-
lues can be observed. The introduction of SNs into the PA layer corporated with SN50 at different concentrations. The results were
may change bulk PA structure and result in the increase of free obtained by measuring the water flux as a function of time during
volume which contributes to higher permeation rates [9,13,51]. On the compaction stage. The compaction rates are 22.8%, 17.6%, 14.2%
the other hand, the hydrophilicity of the TFC membranes increases and 13.3% for TFC, 0.05 wt%, 0.1 wt% and 0.2 wt% SN50-TFC
after SN incorporation, resulting in further accelerating water so- membranes, respectively. Student t-test analysis results show
lubilization and diffusion [51,62–64]. statistically significant improvement of the compaction rates fol-
Although the hydrophilic SNs can enhance the water perme- lowing SN50 concentration increment up to 0.1 wt%. The correla-
ability, there is an optimum concentration to provoke positive tion between the compaction rates and SN50 concentrations of
changes in membrane performance. Beyond this optimum level, 0.1–0.2 wt% is insignificant (p 40.10). Compaction data for the
the permeability may drop due to the pore blocking effect and/or functionalized SN hybridized TFC do not show significant corre-
the aggregation of excess nanoparticles [14,62,63,65,66]. The low lations with these compaction results obtained from other TFC
water flux of the 0.2 wt% SN50-TFC membrane can be attributed to membranes. Prendergast et al. observed similar results with im-
this. Similar trend has been reported in the previous studies on the proved compaction resistance for their zeolite NPs incorporated
62 M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64

TFC membrane and reported the importance of the NP-enhanced Statistic results show significant correlation is associated with
mechanical properties to improve the overall TFC membranes’ desalination performance tests using SN50-OH-TFC (p¼ 0.0004)
performance [69]. Compaction results suggest the NP incorpora- and SN100-OH-TFC (p¼ 0.0019) membranes. The average salt re-
tion could be beneficial for strengthening TFN membranes which jections are slightly higher for the SN-NH2-TFC and SN-EPX-TFC
is another advantage of their application. membranes while the correlations are not significant for all of the
samples. In fact, when comparing the salt permeability coefficients
3.3.2. Effect of nanoparticle surface functionality (B) of the developed membranes, only SN50-NH2-TFC membrane
MPD monomer has –NH2 groups while TMC monomer has – (p ¼0.0430) and SN100-NH2-TFC membrane (p ¼0.0932) show
COCl groups. Interfacial interaction of these functional groups in significantly lower B values compared to the control TFC mem-
the interface of MPD aqueous and TMC organic solutions (on the brane. When the SN-TFC membranes are compared against SN-
PES support) results in the formation of TFC membranes with a OH-TFC membranes, the same results are obtained except that the
thin PA selective layer. To fabricate SN-TFC membranes, SNs can be B values are significantly lower for both of the SN50-NH2-TFC
either dispersed in the aqueous or organic monomer solutions. If (p ¼0.0181) and SN100-NH2-TFC (p ¼0.0450) membranes. These
the SNs can be conjugated with the PA by the chemical interaction data show that all membranes perform almost similar salt rejec-
with either of these monomers, the resulting SN-TFC membranes
tions with slight improvement of the rejection for the amine
are expected to have more stable structure and excellent durability
functionalized SN-TFC membranes. Salt rejection data and corre-
compared to the membranes made through only physical inter-
lations are found to be independent of size variables of the in-
actions with nanoparticles. Hence, we designed SNs with two
corporated SNs.
different surface functional groups: epoxy and amine. It is ex-
The observed performance results can be explained by the
pected that epoxy groups of SN-EPX can interact with the –NH2
higher crosslinking density of the PA layers in the SN-NH2-TFC and
groups of MPD monomer while amine groups of SN-NH2 can in-
SN-EPX-TFC membranes compared to those of the SN-OH-TFC and
teract with the -COCl groups of TMC monomer. After the in-
TFC membranes as evidenced by the XPS analysis results. For the
corporation of the functionalized SNs to the TFC membranes, their
SN-NH2 and SN-EPX, the SNs can be entrapped into the PA layer by
desalination performance was evaluated subject to the SN sizes
the covalent bond formation which ensures the higher stability of
and surface functionalities.
Fig. 8A and Table 3 show that all SN-TFC membranes perform a the resulting SN-TFC membranes. However, SN-OH has no inter-
higher water flux/permeability compared to the control TFC action with the PA monomers, making a looser PA structure with
membrane. The highest water flux is given by the SN-OH-TFC the possibility of the nanoparticles’ leakage after long-time high-
membranes followed by the SN-EPX-TFC and SN-NH2-TFC mem- pressure operation.
branes, respectively. Systematic data for desalination were ana- Another factor that contributes to this performance variation is
lysed by student t-test comparing the intrinsic water permeability the alteration of the amine/acyl chloride molar ratios by the in-
coefficients (Pw) of the different SN-TFC membranes with the troduction of different SNs. That can influence the formation of the
control TFC membrane. A higher average permeability coefficient dense core and the loose secondary PA layers which both play an
is given by all SN-TFC membranes than the control TFC membrane, important role in the membranes’ filtration behavior [40–42]. As
observed from TEM images, the average thickness of the PA layer
in the SN50-NH2-TFC membranes is higher than the SN50-EPX-TFC
Water Flux (LMH) Salt Rejection (%) membranes. Apart from the hydrophilic variation, the thickness
25 100
variation can also contribute to the slightly lower water flux and
Water Flux (L/m²h)

higher salt rejection of the SN-NH2-TFC than the SN-EPX-TFC


20
Salt Rejection (%)

95 membranes. However, due to the large –NH2/–COCl molar ratio in


15 the PA fabrication stage and the small ratio of SN/MPD in the MPD
90 monomer solution, the chemical interaction of the functionalized
10 SNs with the PA monomers could have a larger influence on the
85
overall permeation behavior than the molar ratio variation. That
5 could explain why both SN-NH2-TFC and SN-EPX-TFC membranes
have almost a similar variation trend in terms of water flux and
0 80
TFC SN50 SN50-NH2 SN50-EPX SN100 SN100-NH2 SN100-EPX salt rejection as compared to the SN-OH-TFC and control TFC
TFC/ TFN Membrane Samples membranes.
Fig. 8B shows the normalized salt rejection curves of the
membranes during a 6 h operation against 2000 ppm NaCl solu-
1
tion, where the data are the average of three replicates for each
Normalized Salt Rejection

0.995 membrane and each point is the ratio of salt rejection at a certain
0.99 time normalized by the initial salt rejection values. Overall, all salt
0.985
SN50-TFC rejections experience a slight decrease by time. However, the salt
SN50-NH2-TFC rejection values of the SN-NH2-TFC and SN-EPX-TFC membranes
0.98 SN50-EPX-TFC
show an insignificant change when compared with the SN-OH-TFC
SN100-TFC
0.975 membranes. This is due to the formation of strong chemical bonds
SN100-NH2-TFC
0.97 SN100-EPX-TFC between the SN-NH2 or SN-EPX with the PA layers and the com-
0.965 pact structures of the SN-TFC membranes. However, the SN-OH
can only physically bond with the PA layer; and hence, might leak
0.96
0 1 2 3 4 5 6 under high-pressure process and generate some defects on the SN-
Filtration Time (h) OH-TFC membranes. The small salt rejection variability of the SN-
NH2-TFC and SN-EPX-TFC membranes confirms the high perfor-
Fig. 8. A) The desalination performance evaluation of the TFC membranes in-
corporated with various SNs having different surface functionalities, and B) The mance stability of these membranes which is beneficial for their
normalized salt rejection of the fabricated SN-TFC membranes vs. filtration time. optimum performance over a longer period of time.
M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64 63

4. Conclusions functionalizing
SN-EPX Epoxy functionalized silica nanoparticles
In this study, new TFN membranes incorporated with various SN-NH2 Amine functionalized silica nanoparticles
SNs having different sizes (  50 and  100 nm) and surface func- SN-TFC SN incorporated TFC membranes
tionalities (amine, epoxy or hydroxyl groups) are developed. The PA-SN Polyamide-silica nanoparticle hybrids
research focus is to understand the interactions between the dif- APTMS (3-aminopropyl) trimethoxysilane
ferent SNs and the PA thin film layers, and study the SNs’ influ- GPMS (3-glycidyloxypropyl) trimethoxysilane
ences on the physicochemical properties and the desalination TEOS Tetraethyl orthosilicate
performance of the fabricated membranes. Different SNs could be TEM Transmission electron microscopy
successfully incorporated into the thin film PA layers of the SN-TFC SEM Scanning electron microscopy
membranes as evidenced by detailed characterizations (streaming ATR-FTIR Attenuated total reflectance Fourier transform in-
potential, SEM, TEM, XPS and FTIR). The water flux was sig- frared spectroscopy
nificantly increased by approximate 40% using the 0.1 wt% SN50 DLS Dynamic light scattering
incorporated TFC membrane while the salt rejection changes in- TGA Thermogravimetric analysis
significantly. The cross-section TEM images and FTIR and XPS EDX Energy dispersive x-ray spectroscopy
analysis results show that the functionalized SN-NH2 and SN-EPX XPS X-ray photoelectron spectroscopy
could be stably encapsulated into the PA layers. XPS analysis re- Jw Permeate water flux (LMH)
sults revealed that the SN-NH2-TFC and SN-EPX-TFC membranes Pw Intrinsic water permeability coefficient
show more stable salt rejection performance. Desalination per- Vp Collected permeate volume (L)
formance evaluation of the fabricated membranes shows that the A Active membrane surface area (m2)
SN-TFC membranes have a higher average water flux and almost t Operation time (h)
comparable salt rejection values compared to the control TFC R Salt rejection
membrane. The most significant water permeability correlation CP Equivalent NaCl concentrations in collected
coefficient is given by the SN-OH-TFC membranes, while sig- permeate
nificant salt permeability coefficient corresponding to SN-NH2-TFC CF Equivalent NaCl concentrations in feed water
membranes. Overall, the covalent bond formation between the B Intrinsic salt permeability coefficient
SN-NH2 or SN-EPX and the PA structure was found to improve the
incorporation stability of the SNs inside the PA as well as im-
proving the performance stability of the resulting SN-TFC
membranes.
References

Acknowledgements [1] L.Y. Ng, A.W. Mohammad, C.P. Leo, N. Hilal, Polymeric membranes in-
corporated with metal/metal oxide nanoparticles: a review, Desalination 308
(2013) 15–33.
The authors would like to thank our workshop team, Mr. Jason [2] M. Zargar, B. Jin, S. Dai, Development and application of reverse osmosis for
Peak, Mr. Jeffrey Hiorns and Mr. Michael Jung at the University of separation, in: K. Hu, J. Dickson (Eds.), Membrane Processing for Dairy In-
Adelaide for their help in the establishment of the cross-flow fil- gredient Separation, Wiley-Blackwell, 2015, pp. 139–175.
[3] K.P. Lee, T.C. Arnot, D. Mattia, A review of reverse osmosis membrane materials
tration unit. We also sincerely thank Dr. Sara Azari and Ms. Hanaa for desalination- Development to date and future potential, J. Membr. Sci. 370
Hegab for their help to perform the streaming potential mea- (2011) 1–22.
surements. This project was funded by a Category One Initiative [4] Gd Kang, Ym Cao, Development of antifouling reverse osmosis membranes for
water treatment: a review, Water Res. 46 (2012) 584–600.
Research Grant (IRG 13103302) and IPRS Scholarship for Masou-
[5] S.G. Kim, D.H. Hyeon, J.H. Chun, B.H. Chun, S.H. Kim, Nanocomposite poly
meh Zargar at the University of Adelaide, Australia. (arylene ether sulfone) reverse osmosis membrane containing functional
zeolite nanoparticles for seawater desalination, J. Membr. Sci. 443 (2013)
10–18.
[6] H. Zhao, S. Qiu, L. Wu, L. Zhang, H. Chen, C. Gao, Improving the performance of
Appendix A. Supporting information polyamide reverse osmosis membrane by incorporation of modified multi-
walled carbon nanotubes, J. Membr. Sci. 450 (2014) 249–256.
Supplementary data associated with this article can be found in [7] H. Dong, L. Wu, L. Zhang, H. Chen, C. Gao, Clay nanosheets as charged filler
materials for high-performance and fouling-resistant thin film nanocomposite
the online version at http://dx.doi.org/10.1016/j.memsci.2016.08.069. membranes, J. Membr. Sci. 494 (2015) 92–103.
[8] B.H. Jeong, E.M.V. Hoek, Y. Yan, A. Subramani, X. Huang, G. Hurwitz, A.
K. Ghosh, A. Jawor, Interfacial polymerization of thin film nanocomposites: a
new concept for reverse osmosis membranes, J. Membr. Sci. 294 (2007) 1–7.
[9] M.L. Lind, A.K. Ghosh, A. Jawor, X. Huang, W. Hou, Y. Yang, E.M.V. Hoek, In-
Nomenclature fluence of zeolite crystal size on zeolite-polyamide thin film nanocomposite
membranes, Langmuir 25 (2009) 10139–10145.
[10] G.L. Jadav, P.S. Singh, Synthesis of novel silica-polyamide nanocomposite
TFC Thin film composite
membrane with enhanced properties, J. Membr. Sci. 328 (2009) 257–267.
TFN Thin film nanocomposite [11] L. Jin, W. Shi, S. Yu, X. Yi, N. Sun, C. Ma, Y. Liu, Preparation and characterization
PA Polyamide of a novel PA-SiO2 nanofiltration membrane for raw water treatment, Desa-
lination 298 (2012) 34–41.
PES Polyether sulfone
[12] A. Tiraferri, Y. Kang, E.P. Giannelis, M. Elimelech, Highly hydrophilic thin-film
MPD 1, 3-phenylenediamine composite forward osmosis membranes functionalized with surface-tailored
TMC 1,3,5-benzenetricarbonyl trichloride nanoparticles, ACS Appl. Mater. Interfaces 4 (2012) 5044–5053.
SN Silica nanoparticles [13] H. Wu, B. Tang, P. Wu, Optimizing polyamide thin film composite membrane
covalently bonded with modified mesoporous silica nanoparticles, J. Membr.
SN50 Silica nanoparticles with average particle size of Sci. 428 (2013) 341–348.
50 nm [14] J. Yin, E.S. Kim, J. Yang, B. Deng, Fabrication of a novel thin-film nanocomposite
SN100 Silica nanoparticles with average particle size of (TFN) membrane containing MCM-41 silica nanoparticles (NPs) for water
purification, J. Membr. Sci. 423–424 (2012) 238–246.
100 nm [15] R.X. Zhang, L. Braeken, P. Luis, X.L. Wang, B. Van der Bruggen, Novel binding
SN-OH Original silica nanoparticles without further surface procedure of TiO2 nanoparticles to thin film composite membranes via self-
polymerized polydopamine, J. Membr. Sci. 437 (2013) 179–188.
64 M. Zargar et al. / Journal of Membrane Science 521 (2017) 53–64

[16] D. Emadzadeh, W.J. Lau, T. Matsuura, A.F. Ismail, M. Rahbari-Sisakht, Synthesis [44] S. Yu, X. Liu, J. Liu, D. Wu, M. Liu, C. Gao, Surface modification of thin-film
and characterization of thin film nanocomposite forward osmosis membrane composite polyamide reverse osmosis membranes with thermo-responsive
with hydrophilic nanocomposite support to reduce internal concentration polymer (TRP) for improved fouling resistance and cleaning efficiency, Sep.
polarization, J. Membr. Sci. 449 (2014) 74–85. Purif. Technol. 76 (2011) 283–291.
[17] J. Yin, Y. Yang, Z. Hu, B. Deng, Attachment of silver nanoparticles (AgNPs) onto [45] S.S. Deshmukh, A.E. Childress, Zeta potential of commercial RO membranes:
thin-film composite (TFC) membranes through covalent bonding to reduce influence of source water type and chemistry, Desalination 140 (2001) 87–95.
membrane biofouling, J. Membr. Sci. 441 (2013) 73–82. [46] S.H. Kim, S.Y. Kwak, T. Suzuki, Positron annihilation spectroscopic evidence to
[18] Y. Liu, E. Rosenfield, M. Hu, B. Mi, Direct observation of bacterial deposition on demonstrate the flux-enhancement mechanism in morphology-controlled
and detachment from nanocomposite membranes embedded with silver na- thin-film-composite membrane, Environ. Sci. Technol. 39 (2005) 1764–1770.
noparticles, Water Res. 47 (2013) 2949–2958. [47] J. Duan, Y. Pan, F. Pacheco, E. Litwiller, Z. Lai, I. Pinnau, High-performance
[19] F. Perreault, M.E. Tousley, M. Elimelech, Thin-film composite polyamide polyamide thin-film-nanocomposite reverse osmosis membranes containing
membranes functionalized with biocidal graphene oxide nanosheets, Environ. hydrophobic zeolitic imidazolate framework-8, J. Membr. Sci. 476 (2015)
Sci. Technol. Lett. 1 (2014) 71–76. 303–310.
[20] L. He, L.F. Dumée, C. Feng, L. Velleman, R. Reis, F. She, W. Gao, L. Kong, Pro- [48] S.Y. Kwak, S.H. Kim, S.S. Kim, Hybrid organic/inorganic reverse osmosis (RO)
moted water transport across graphene oxide–poly(amide) thin film compo- Membrane for bactericidal anti-fouling. 1. Preparation and characterization of
site membranes and their antibacterial activity, Desalination 365 (2015) TiO2 nanoparticle self-assembled aromatic polyamide thin-film-composite
126–135. (TFC) membrane, Environ. Sci. Technol. 35 (2001) 2388–2394.
[21] H.R. Chae, J. Lee, C.H. Lee, I.C. Kim, P.K. Park, Graphene oxide-embedded thin- [49] D. Emadzadeh, W.J. Lau, M. Rahbari-Sisakht, H. Ilbeygi, D. Rana, T. Matsuura, A.
film composite reverse osmosis membrane with high flux, anti-biofouling, and F. Ismail, Synthesis, modification and optimization of titanate nanotubes-
chlorine resistance, J. Membr. Sci. 483 (2015) 128–135. polyamide thin film nanocomposite (TFN) membrane for forward osmosis
[22] C.F. de Lannoy, E. Soyer, M.R. Wiesner, Optimizing carbon nanotube-re- (FO) application, Chem. Eng. J. 281 (2015) 243–251.
inforced polysulfone ultrafiltration membranes through carboxylic acid [50] S. Romero-Vargas Castrillón, X. Lu, D.L. Shaffer, M. Elimelech, Amine enrich-
functionalization, J. Membr. Sci. 447 (2013) 395–402. ment and poly(ethylene glycol) (PEG) surface modification of thin-film com-
[23] L. Dumée, J. Lee, K. Sears, B. Tardy, M. Duke, S. Gray, Fabrication of thin film posite forward osmosis membranes for organic fouling control, J. Membr. Sci.
composite poly(amide)-carbon-nanotube supported membranes for enhanced 450 (2014) 331–339.
performance in osmotically driven desalination systems, J. Membr. Sci. 427 [51] N. Niksefat, M. Jahanshahi, A. Rahimpour, The effect of SiO2 nanoparticles on
(2013) 422–430. morphology and performance of thin film composite membranes for forward
[24] H. Yamada, C. Urata, Y. Aoyama, S. Osada, Y. Yamauchi, K. Kuroda, Preparation osmosis application, Desalination 343 (2014) 140–146.
of colloidal mesoporous silica nanoparticles with different diameters and their [52] M. Bao, G. Zhu, L. Wang, M. Wang, C. Gao, Preparation of monodispersed
unique degradation behavior in static aqueous systems, Chem. Mater. 24 spherical mesoporous nanosilica–polyamide thin film composite reverse os-
(2012) 1462–1471. mosis membranes via interfacial polymerization, Desalination 309 (2013)
[25] K. Fujiwara, H. Suematsu, E. Kiyomiya, M. Aoki, M. Sato, N. Moritoki, Size- 261–266.
dependent toxicity of silica nano-particles to Chlorella kessleri, J. Environ. Sci. [53] B. Rajaeian, A. Rahimpour, M.O. Tade, S. Liu, Fabrication and characterization
Health Part A 43 (2008) 1167–1173. of polyamide thin film nanocomposite (TFN) nanofiltration membrane im-
[26] I.Y. Kim, E. Joachim, H. Choi, K. Kim, Toxicity of silica nanoparticles depends on pregnated with TiO2 nanoparticles, Desalination 313 (2013) 176–188.
size, dose, and cell type, Nanomed.: Nanotechnol. Biol. Med. 11 (2015) [54] M. Zargar, Y. Hartanto, B. Jin, S. Dai, Hollow mesoporous silica nanoparticles: A
1407–1416. peculiar structure for thin film nanocomposite membranes, J. Membr. Sci. 519
[27] J.S. Chang, K.L.B. Chang, D.F. Hwang, Z.L. Kong, In vitro cytotoxicitiy of silica (2016) 1–10.
nanoparticles at high concentrations strongly depends on the metabolic ac- [55] L.W. Zhikang Xu, Xiaojun Huang, Surface Engineering of Polymer Membranes,
tivity type of the cell line, Environ. Sci. Technol. 41 (2007) 2064–2068. Springer, Berlin Heidelberg, 2009.
[28] A. Tiraferri, Y. Kang, E.P. Giannelis, M. Elimelech, Superhydrophilic thin-film [56] A. Prakash Rao, N.V. Desai, R. Rangarajan, Interfacially synthesized thin film
composite forward osmosis membranes for organic fouling control: fouling composite RO membranes for seawater desalination, J. Membr. Sci. 124 (1997)
behavior and antifouling mechanisms, Environ. Sci. Technol. 46 (2012) 263–272.
11135–11144. [57] G.D. Vilakati, M.C.Y. Wong, E.M.V. Hoek, B.B. Mamba, Relating thin film com-
[29] W. Stöber, A. Fink, J. Colloid Interface Sci. 26 (1968) 62. posite membrane performance to support membrane morphology fabricated
[30] S. Fouilloux, A. Désert, O. Taché, O. Spalla, J. Daillant, A. Thill, SAXS exploration using lignin additive, J. Membr. Sci. 469 (2014) 216–224.
of the synthesis of ultra monodisperse silica nanoparticles and quantitative [58] A.K. Ghosh, E.M.V. Hoek, Impacts of support membrane structure and chem-
nucleation growth modeling, J. Colloid Interface Sci. 346 (2010) 79–86. istry on polyamide–polysulfone interfacial composite membranes, J. Membr.
[31] B. Chang, D. Chen, Y. Wang, Y. Chen, Y. Jiao, X. Sha, W. Yang, Bioresponsive Sci. 336 (2009) 140–148.
controlled drug release based on mesoporous silica nanoparticles coated with [59] S. Hermans, R. Bernstein, A. Volodin, I.F.J. Vankelecom, Study of synthesis
reductively sheddable polymer shell, Chem. Mater. 25 (2013) 574–585. parameters and active layer morphology of interfacially polymerized poly-
[32] J. Lu, W. Wei, L. Yin, Y. Pu, S. Liu, Flow injection chemiluminescence im- amide–polysulfone membranes, React. Funct. Polym. 86 (2015) 199–208.
munoassay of microcystin-LR by using PEI-modified magnetic beads as cap- [60] S.H. Maruf, A.R. Greenberg, Y. Ding, Influence of substrate processing and in-
turer and HRP-functionalized silica nanoparticles as signal amplifier, Analyst terfacial polymerization conditions on the surface topography and permse-
138 (2013) 1483–1489. lective properties of surface-patterned thin-film composite membranes, J.
[33] K. Möller, J. Kobler, T. Bein, Colloidal suspensions of nanometer-sized meso- Membr. Sci. 512 (2016) 50–60.
porous silica, Adv. Funct. Mater. 17 (2007) 605–612. [61] M. Son, Hg Choi, L. Liu, E. Celik, H. Park, H. Choi, Efficacy of carbon nanotube
[34] Y. Sun, Z. Zhang, C.P. Wong, Study on mono-dispersed nano-size silica by positioning in the polyethersulfone support layer on the performance of thin-
surface modification for underfill applications, J. Colloid Interface Sci. 292 film composite membrane for desalination, Chem. Eng. J. 266 (2015) 376–384.
(2005) 436–444. [62] E.S. Kim, B. Deng, Fabrication of polyamide thin-film nano-composite (PA-TFN)
[35] B. Deng, J. Yin, Thin-film nano-composite membrane with mesoporous silica membrane with hydrophilized ordered mesoporous carbon (H-OMC) for
nanoparticles, in, Google Pat. (2014). water purifications, J. Membr. Sci. 375 (2011) 46–54.
[36] V. Freger, J. Gilron, S. Belfer, TFC polyamide membranes modified by grafting [63] A. Peyki, A. Rahimpour, M. Jahanshahi, Preparation and characterization of
of hydrophilic polymers: an FT-IR/AFM/TEM study, J. Membr. Sci. 209 (2002) thin film composite reverse osmosis membranes incorporated with hydro-
283–292. philic SiO2 nanoparticles, Desalination 368 (2015) 152–158.
[37] A. Prakash Rao, S.V. Joshi, J.J. Trivedi, C.V. Devmurari, V.J. Shah, Structure– [64] S. Yu, Z. Lü, Z. Chen, X. Liu, M. Liu, C. Gao, Surface modification of thin-film
performance correlation of polyamide thin film composite membranes: effect composite polyamide reverse osmosis membranes by coating N-iso-
of coating conditions on film formation, J. Membr. Sci. 211 (2003) 13–24. propylacrylamide-co-acrylic acid copolymers for improved membrane prop-
[38] S. Salgın, Effects of ionic environments on bovine serum albumin fouling in a erties, J. Membr. Sci. 371 (2011) 293–306.
cross-flow ultrafiltration system, Chem. Eng. Technol. 30 (2007) 255–260. [65] M. Safarpour, A. Khataee, V. Vatanpour, Thin film nanocomposite reverse os-
[39] A.K. Ghosh, B.H. Jeong, X. Huang, E.M.V. Hoek, Impacts of reaction and curing mosis membrane modified by reduced graphene oxide/TiO2 with improved
conditions on polyamide composite reverse osmosis membrane properties, J. desalination performance, J. Membr. Sci. 489 (2015) 43–54.
Membr. Sci. 311 (2008) 34–45. [66] S. Bano, A. Mahmood, S.J. Kim, K.H. Lee, Graphene oxide modified polyamide
[40] V. Freger, Nanoscale Heterogeneity of polyamide membranes formed by in- nanofiltration membrane with improved flux and antifouling properties, J.
terfacial polymerization, Langmuir 19 (2003) 4791–4797. Mater. Chem. A 3 (2015) 2065–2071.
[41] A.V. Berezkin, A.R. Khokhlov, Mathematical modeling of interfacial poly- [67] E.M.V. Hoek, A.K. Ghosh, X. Huang, M. Liong, J.I. Zink, Physical–chemical
condensation, J. Polym. Sci. Part B: Polym. Phys. 44 (2006) 2698–2724. properties, separation performance, and fouling resistance of mixed-matrix
[42] W. Xie, G.M. Geise, H.S. Lee, G. Byun, J.E. McGrath, Polyamide interfacial ultrafiltration membranes, Desalination 283 (2011) 89–99.
composite membranes prepared from m-phenylene diamine, trimesoyl [68] M.T.M. Pendergast, J.M. Nygaard, A.K. Ghosh, E.M.V. Hoek, Using nano-
chloride and a new disulfonated diamine, J. Membr. Sci. 403–404 (2012) composite materials technology to understand and control reverse osmosis
152–161. membrane compaction, Desalination 261 (2010) 255–263.
[43] M. Elimelech, W.H. Chen, J.J. Waypa, Measuring the zeta (electrokinetic) po- [69] M.M. Pendergast, A.K. Ghosh, E.M.V. Hoek, Separation performance and in-
tential of reverse osmosis membranes by a streaming potential analyzer, terfacial properties of nanocomposite reverse osmosis membranes, Desalina-
Desalination 95 (1994) 269–286. tion 308 (2013) 180–185.

Das könnte Ihnen auch gefallen