You are on page 1of 11

Showcasing a study on the graphene oxide quantum As featured in:

dot incorporated reverse osmosis membrane by


Zhining Wang et al. from Ocean University of China.

Pressure-assisted preparation of graphene oxide quantum


dot-incorporated reverse osmosis membranes: antifouling
and chlorine resistance potentials

A graphene oxide quantum dot incorporated thin film


nanocomposite reverse osmosis membrane was developed
by filtration-assisted interfacial polymerization. The prepared
membrane exhibited excellent performance, high antifouling
capacity and improved chlorine resistance.

See Zhining Wang et al.,


J. Mater. Chem. A, 2016, 4, 16896.

www.rsc.org/MaterialsA
Registered charity number: 207890
Journal of
Materials Chemistry A
View Article Online
PAPER View Journal | View Issue

Pressure-assisted preparation of graphene oxide


quantum dot-incorporated reverse osmosis
Cite this: J. Mater. Chem. A, 2016, 4,
16896 membranes: antifouling and chlorine resistance
potentials†
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

Xiangju Song,a Quanzhu Zhou,b Tian Zhang,a Haibo Xub and Zhining Wang*a

Graphene-based nanomaterials have opened a new era in the fabrication of novel membranes with
outstanding performance. In this study, we propose a new method of fabricating a graphene oxide
quantum dot (GOQD)-incorporated thin-film nanocomposite (TFN) membrane for reverse osmosis (RO)
applications. Inspired by free-standing graphene membranes prepared by vacuum filtration, an aqueous
suspension of GOQD/m-phenylenediamine (MPD) was first filtered onto a polysulfone (PSf) substrate to
obtain a cushion layer. A GOQD-incorporated polyamide (PA) selective layer was then constructed by
interfacial polymerization between MPD and trimesoyl chloride (TMC). This method can easily be scaled
up owing to its simple preparation procedures. The newly developed TFN membrane exhibited stable
high performance with a flux of 37.5 L m2 h1 and a NaCl rejection of 98.8% at 16 bar. Remarkably, this
represents a 51.8% increase in permeate flux relative to a GOQD-free TFC membrane without
Received 3rd August 2016
Accepted 12th September 2016
compromising the solute rejection. In addition, the GOQD-incorporated TFN membrane displayed
long-term durability over 120 h of RO testing. More importantly, the introduction of GOQD into the TFN
DOI: 10.1039/c6ta06636d
membrane resulted in improved antifouling and chlorine resistance properties, which are greatly desired
www.rsc.org/MaterialsA in the membrane desalination and water reclamation processes.

characteristics in comparison with conventional thin-lm


1 Introduction composite (TFC) membranes. The structure and physical and
Water crises represented the top global risk of 2015 and it is chemical properties of the membranes can be tuned by the
believed that these will be a continuing environmental and social introduction of selected nanomaterials to meet the needs of
problem in the coming decade.1 Membrane separation technology, specic water treatment applications. Certain TFN membranes
in particular reverse osmosis (RO), has become the most effective have helped to overcome the trade-off between water permeability
means of producing large quantities of freshwater from abun- and solute selectivity and have also demonstrated excellent anti-
dantly available brackish water or seawater owing to the advan- fouling capacity and improved bactericidal effects.10,11
tages of a high permeate ux, ease of operation, and low process Recently, the advent of graphene oxide (GO) has opened a new
costs.2,3 Among novel RO membranes, thin-lm nanocomposite era in the fabrication of TFN membranes with high separation
(TFN) membranes fabricated by incorporating nanomaterials into performance owing to its distinctive two-dimensional structure
a dense polyamide (PA) layer have emerged as a promising solu- and single-atom layer motif, high mechanical strength, good
tion to the bottlenecks of current membrane desalination tech- exibility, cost-effectiveness and scalable production.4,12–18 GO-
nologies.4–9 TFN membranes can provide enhanced separation containing membranes promise to improve water permeability
by permitting the rapid diffusion of water through the nanopores
a
Key Laboratory of Marine Chemistry Theory and Technology, Ministry of Education,
in GO,19–21 the two-dimensional nanochannels between GO
Ocean University of China, Qingdao 266100, PR China. E-mail: wangzhn@ouc.edu.cn akes,22,23 and the nanocorridors between GO and the polymer
b
College of Chemistry and Chemical Engineering, Ocean University of China, Qingdao matrix.24 The unique structure and surface properties of GO also
266100, PR China impede the adsorption of hydrophobic foulants and provide
† Electronic supplementary information (ESI) available: Characterization of GO; strong antimicrobial characteristics to prevent colonization of
characterization of GO-incorporated membranes; photographs of the prepared
the membrane by a wide variety of microorganisms.25–28 In
membranes; SEM and AFM images of GO-incorporated membranes; RO
performance of GO-incorporated membranes; photographs of the
addition, GO akes incorporated into the membranes hinder
GO-incorporated membranes before and aer RO testing; SEM images of the dissolved active chlorine ions from contacting the PA selective
polyamide layers of TFC and TFN-GOQD5.0 membranes; comparison of the RO layer, which results in efficient chlorine resistance without the
performance of membranes prepared with and without the assistance of loss of intrinsic water permeability and solute selectivity.26,29,30
pressure. See DOI: 10.1039/c6ta06636d

16896 | J. Mater. Chem. A, 2016, 4, 16896–16905 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper Journal of Materials Chemistry A

The size of GO has been found to have an important inuence selectivity, excellent stability, and improved antifouling capacity
on the performance of various GO-based membranes. Chen et al. and chlorine tolerance. These ndings have important impli-
reported that a free-standing GO membrane composed of small cations for the design of high-performance TFN membranes.
GO akes provided additional shorter paths for water molecules
to penetrate the membrane in comparison with large GO akes.31
2 Experimental
The inuence of the size of GO on the permeation properties of
solutes in water through free-standing GO membranes has been 2.1 Materials
investigated by Sun et al.32 As the size of GO akes decreased GOQD was kindly provided by Qingdao Haida Haixi New
from several micrometers to several hundred nanometers, the Materials Co., Ltd (China). GO was obtained from Shanxi Fen-
ion permeability increased signicantly, which can be attributed ghuiyuan Science and Technology Co., Ltd (China). The poly-
to an increase in nanocapillaries formed between neighboring sulfone membrane was purchased from Hangzhou Water
GO akes. Lee et al. prepared membranes with a GO-embedded Treatment Center (China). m-Phenylenediamine (MPD, >99%)
polysulfone (PSf) substrate and found that small GO akes were and bovine serum albumin (BSA, $98%, MW ¼ 67 000 g mol1)
more effective in improving the membrane's mechanical prop- were purchased from Sigma-Aldrich Chemical Co., Ltd (China).
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

erties.33 On the one hand, small GO akes possess a large surface Trimesoyl chloride (TMC, 98%) was obtained from Aladdin Co.,
area, which induces strong interfacial interactions between the Ltd (China). Other chemicals were obtained from Sinopharm
nanollers and the PSf matrix. On the other hand, the excellent Chemical Reagent Co., Ltd (China) and were used without
dispersion of the small GO akes contributes to an improvement further purication.
in mechanical strength. In addition, the large surface area in the
nanochannels also ensures that a more porous nanocomposite 2.2 Membrane preparation
substrate can be obtained. Perreault et al. reported another
A 0.1 mg mL1 suspension of GOQD was prepared by ultra-
important nding, namely, that a GO-modied membrane surface
sonication for 1 h. Then, 2 wt% MPD was added to obtain
exhibited prominent size-dependent antimicrobial activity based
a GOQD/MPD suspension. A GOQD/PSf membrane with certain
on GO akes.28 A membrane surface coated with small GO akes
amounts of GOQD (1.0, 2.0, 5.0, and 10.0 mg) was fabricated
exhibited higher antimicrobial activity than that of membranes
using various volumes of the GOQD/MPD suspension with the
coated with large GO akes, which can be explained by an increase
assistance of N2 pressure. As shown in Fig. 1, a certain volume
in oxidative behavior that resulted from the higher defect density
of the GOQD/MPD suspension was subjected to pressure-
as the size of GO akes decreased.
assisted ltration with N2 at a pressure of 2 bar using a dead-
Given this, we proposed to utilize graphene oxide quantum
end ltration device to deposit a cushion of GOQD/MPD onto
dot (GOQD) as nanollers to fabricate novel TFN membranes.
a PSf substrate. The composite membrane was dried in air.
GOQD adopt the same one-atom-thick structure as GO but have
A 0.1 w/v% solution of TMC/n-hexane was poured onto this
lateral dimensions of less than 100 nm.34,35 GOQD possess
membrane for 1 min to form a GOQD-incorporated polyamide
hydrophilic groups that are similar to those of GO, including
layer via interfacial polymerization. The resulting TFN membrane
epoxy, hydroxyl and carboxyl groups. These groups help to
was dried in an oven at 80  C for 5 min and was then washed
produce excellent dispersity of GOQD. Furthermore, GOQD
and stored in DI water for at least 24 h before testing. According
have a greater edge-to-volume ratio than micron-sized GO.36
to the GOQD loading (1.0, 2.0, 5.0, and 10.0 mg), the membranes
Therefore, GOQD have great potential for use as nanollers for
were denoted as TFN-GOQD1.0, TFN-GOQD2.0, TFN-GOQD5.0,
fabricating high-performance membranes.37 In this study, extra
and TFN-GOQD10.0. As controls, a TFC membrane and a GO-
pressure was applied to lter an aqueous suspension of GOQD/
incorporated TFN (TFN-GO) membrane were also prepared via
m-phenylenediamine (MPD) to accumulate and deposit a GOQD
the same procedure, employing MPD and GO/MPD solution as
cushion layer on a PSf substrate (Fig. 1). Subsequently, tri-
the ltrate.
mesoyl chloride (TMC) was reacted with MPD to form a GOQD-
incorporated PA layer by interfacial polymerization. In compar-
ison with a reference TFC membrane, the TFN membrane 2.3 Characterization
exhibited enhanced water permeability with retention of solute Atomic force microscopy (AFM) images of GOQD were obtained
using a Digital Instrument Nanoscope V (Veeco MultiMode)
atomic force microscope. A sample dispersion was drop-cast
onto a clean mica wafer and dried at room temperature. A FEI
TECNAI G2 F20 transmission electron microscope (TEM) oper-
ating at 200 kV was used to investigate the morphology of
GOQD. Before measurement, GOQD was sonically dispersed in
water and then dropped onto a carbon lm-coated copper grid
and dried in air for 2 h. The particle size and zeta potential of
GOQD were determined by dynamic light scattering (DLS,
Malvern Zetasizer Nano series, UK). The pH value of the GOQD
Fig. 1 Schematic of the fabrication procedure of GOQD-incorporated suspension was tuned by adding 0.1 M HCl or 0.1 M NaOH
RO membranes with the assistance of pressure. solution. The chemical properties of GOQD and the membranes

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. A, 2016, 4, 16896–16905 | 16897
View Article Online

Journal of Materials Chemistry A Paper

were investigated using a Fourier transform infrared (FTIR) where DP is the transmembrane pressure difference and Dp is
spectrometer (Tensor 27, Bruker, Germany) and micro-Raman the difference in osmotic pressure across the membrane.
spectroscopy (Renishaw, UK). Raman spectra were acquired at The stability and durability of the TFC and TFN-GOQD5.0
l ¼ 532 nm and 10 mW. For the TFN-GOQD membranes, membranes were measured via continuous long-term ltration
a micro-Raman laser probe was focused on the island-shaped tests with 2 g L1 NaCl solution as the feed solution at 16 bar
domains on the membrane surface. All the samples were and 25  C for 120 h. In the beginning, the water ux and salt
vacuum-dried at 40  C overnight before FTIR and Raman rejection were recorded every 1 h and then the intervals were
measurements. The thermal stability of the selective layers of extended to 3 h and 5 h and eventually to 10 h.
the TFC and TFN membranes was investigated by thermogra- A fouling test was performed using bovine serum albumin
vimetric analysis (TGA) in N2. Prior to measurement, the thin (BSA) and humic acid (HA) as model foulants. Prior to the
polyamide layer on the PSf substrate was separated from the PSf fouling experiment, a RO experiment with a 2 g L1 NaCl feed
support by dissolving PSf using dichloromethane. In each run, solution was rst conducted for 3 h to determine the initial
samples of about 5 mg were loaded into a crucible and heated value of the permeate ux. Then, BSA or HA was added to the
to 800  C at a rate of 10  C min1. The gas ow rate was 2 g L1 NaCl solution to obtain a concentration of 0.5 g L1. The
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

maintained at 10 mL min1. fouling test was continued for 21 h. Subsequently, the fouled
The surface charge of the membrane was measured by an membrane was cleaned using DI water for 2 h at 6 bar with
electrokinetic analyzer (Anton Paar SurPASS, Austria) based on a cross-ow velocity of 72 cm s1 and then the ux recovery was
the measurement of streaming potential with 1 mM KCl as the measured using a 2 g L1 NaCl feed solution for 3 h. The
electrolyte solution at a pH of 7.5. Contact angle measurements recovery rate (FRR) and total fouling rate (Rt) were calculated as
were obtained using a drop shape analysis system (DSA 100, follows:40
Kruss, Germany) to measure the hydrophilicity of the membrane.
A 5.0 mL droplet of DI water was placed on the membrane surface FRR ¼ JR/J0  100% (5)
using a microsyringe with a stainless-steel needle. To reduce
error, average values of the surface charge and contact angle were Rt ¼ (J0  JS)/J0  100% (6)
obtained by measuring six different spots on at least two different
where J0 is the initial ux, JR is the recovered ux aer washing
membranes. To determine the surface and cross-sectional
and JS is the steady ux for BSA or HA solution.
morphologies of the membranes, an S-4800 (Hitachi, Japan)
The chlorine resistance of the membrane was investigated by
scanning electron microscope (SEM) was employed. AFM was
immersing the membrane in NaClO solutions of different
also used to analyze the surface morphology and roughness. The
concentrations. Firstly, the membrane was tested using a 2 g L1
root-mean-square (RMS) roughness was obtained from three
NaCl feed solution to measure the initial water ux and salt
different areas of each membrane.
rejection. Next, the membrane was exposed to a NaClO solution
(100, 500, 1000, 1500, 2000, 3000 and 5000 ppm) at a pH of 4 for
2.4 RO performance 1 h. Aer chlorination, the membrane was rinsed thoroughly
The RO performance was measured using a lab-scale cross-ow with DI water and stored in DI water for 24 h before further
RO unit, which consisted of three parallel ltration cells with an testing. Finally, the water ux and salt rejection of each chlorine-
effective membrane area of 19.6 cm2. The average cross-ow treated membrane were measured.
velocity was 17 cm s1. A 2 g L1 solution of NaCl was used as
a feed solution. The membrane was compacted at 18 bar for
0.5 h and then measured at 16 bar. The concentrations of NaCl
3 Results and discussion
solutions were measured using a conductivity meter. During the 3.1 GOQD characterization
RO measurements, the water ux (J) and salt rejection (R) were The chemical and physical properties of GOQD are shown in
calculated by the following equations:38 Fig. 2. Fig. 2A and B show an AFM image of GOQD and its height
distribution. The height of GOQD varied from 0.4 nm to 1.8 nm,
J ¼ DVfeed/(AmDt) (1) whereas 83% of the materials were less than 1 nm in height,
which meant that a majority of GOQD comprised 1–3 graphene
R ¼ (Cf  Cp)/Cf  100% (2)
layers.41,42 Dynamic light scattering (DLS) measurements were
where DVfeed is the change in the volume of the feed solution, conducted and it was found that GOQD had an approximate
Am is the effective area of the membrane and Dt is the perme- diameter distribution of 7.5–21 nm (average diameter of 11 nm),
ation time. The parameters Cf and Cp are the salt concentrations as shown in Fig. 2C. These results correlated well with previ-
in the feed and permeate. ously reported ndings.35,43
The water permeability (A) and solute permeability (B) were A TEM image of GOQD (Fig. 2D) revealed a dot-like
calculated as follows:39 morphology. The crystal structure of GOQD was investigated
using HR-TEM, as shown in Fig. 2E, which reveals a group of
A ¼ J/DP (3) parallel crystal planes with lattice spacings of 0.203 nm and
0.339 nm, respectively, which match the (100) and (002) crystal
B ¼ A(1  R)(DP  Dp)/R (4) planes of graphite.44 In addition, the corresponding fast Fourier

16898 | J. Mater. Chem. A, 2016, 4, 16896–16905 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper Journal of Materials Chemistry A


Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

Fig. 2 Characterization of GOQD. (A) AFM image of GOQD, (B) height distribution of GOQD from (A), (C) size distribution of GOQD obtained
from dynamic light scattering (DLS) measurements, (D) TEM image of GOQD, (E) high-resolution transmission electron microscopy (HR-TEM)
image of GOQD; the insets show fast Fourier transform (FFT) patterns, (F) size distribution of GOQD from (D), (G) zeta potential of GOQD at
different pH values, (H) FTIR spectrum, and (I) Raman spectrum of GOQD.

transform (FFT) patterns (inset in Fig. 2E) display these repre- in Fig. S1 in the ESI†), which indicated that there were more
sentative characteristics. By determining the size of well-dispersed structural defects or disorders on GOQD nanosheets as a result of
GOQD from the TEM image (Fig. 2F), a narrow distribution of the higher edge-to-volume ratio of GOQD than that of GO.50
5–12 nm was observed. As shown in Fig. 2G, GOQD were negatively
charged over a wide pH range and the zeta potential attained
a minimum value of 36.5 mV at a pH of 7, owing to the ionization 3.2 Membrane characterization
of COOH groups to COO groups.45 Fig. 3 illustrates the chemical and physical properties of the TFC
FTIR analysis was employed to reveal the chemical nature of and GOQD-incorporated TFN membranes. To conrm the
the functional groups on the surface of GOQD. Fig. 2H shows successful interfacial polymerization of polyamide and the
peaks at 3436 cm1, 1720 cm1, 1635 cm1, and 1072 cm1, incorporation of GOQD in the PA layer, ATR-FTIR and micro-
which were attributed to the stretching vibrations of hydroxyl Raman analysis were performed. Fig. 3A shows the FTIR spectra
moieties, the stretching vibrations of the carbonyl unit in of the TFC and TFN-GOQD5.0 membranes. The TFC membrane
carboxyl groups, the asymmetrical stretching vibrations of exhibited characteristic peaks at 1660 cm1, 1610 cm1 and
carboxylate groups and the stretching vibrations of epoxy 1542 cm1, which corresponded to amide I C]O stretching
groups, respectively.46 To further reveal the structural properties vibrations, hydrogen-bonded C]O stretching vibrations, and
of GOQD, Raman spectroscopic analysis was employed. The two amide II N–H stretching vibrations, respectively.30 In the case of
characteristic peaks of the D band at 1347 cm1 and the G band the TFN-GOQD5.0 membrane, the broad band at approximately
at 1590 cm1 can be observed in Fig. 2I. In general, the D band 3440 cm1 was intensied, which conrmed the incorporation
provides evidence of the presence of structural defects and of GOQD nanosheets with abundant carboxylic acid and
disorders in the graphite lattice generated by vacancies, edges, hydroxyl groups.51 In the Raman results (Fig. 3B), the TFC
and epoxide and hydroxyl groups47,48 and the G band is associ- membrane exhibited three characteristic peaks at 1457 cm1,
ated with the sp2 hybridization of graphitic carbon.49 Therefore, 1525 cm1, and 1635 cm1, which were produced by aromatic
the ratio of the peak intensity of the D band to that of the G band rings in the PSf substrate and C]O groups in the PA layer.52,53
(ID/IG) demonstrates the extent of surface defects. GOQD exhibi- The TFN-GOQD5.0 membrane exhibited a typical D band at
ted a higher value of ID/IG (1.12) than that of GO (0.94, as shown 1346 cm1 and a G band at 1584 cm1, which demonstrated the

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. A, 2016, 4, 16896–16905 | 16899
View Article Online

Journal of Materials Chemistry A Paper


Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

Fig. 3 Characterization of the prepared membranes. (A) ATR-FTIR


spectra, (B) micro-Raman spectra, (C) zeta potential and (D) contact
angle of the TFC and GOQD-incorporated membranes. The micro-
Raman spectrum of TFN-GOQD5.0 was obtained by focusing the laser
spot on the island-shaped domains on the membrane surface.
Fig. 4 SEM images of surface morphology and cross-sectional
morphology and AFM images of the prepared membranes. SEM and
AFM (scan size 1 mm  1 mm, z-scale 1 mm) images of (A, D and G) TFC,
incorporation of GOQD into the membrane. It is supposed that (B, E and H) TFN-GOQD5.0, and (C, F and I) TFN-GOQD10.0
membranes.
carboxyl groups on GOQD reacted with the amine groups of
MPD in the mixed solution. During the interfacial polymeriza-
tion process, carboxyl groups on GOQD formed covalent bonds However, based on the RMS roughness values calculated from the
via condensation reactions with terminal carboxyl groups of AFM images, it was concluded that the TFN membrane was
TMC in the linear portion of polyamide. Both factors made the rougher than the TFC membrane, which was attributed to the
selective layer more compact, which improved the stability of appearance of island-shaped protrusions. A mixture of GOQD and
the membrane. MPD was pressed and ltered onto a PSf substrate to form
The surface charge and contact angle of the prepared a cushion before adding TMC. Fig. S3† shows homogeneous
membranes are shown in Fig. 3C and D. Both parameters GOQD/MPD cushion lms when the GOQD content was less than
decreased with an increase in the GOQD content. The presence 5 mg. The color of the membrane became darker with an increase
of numerous negatively charged oxygen-containing groups on in the GOQD content (Fig. S3 in the ESI†). During the interfacial
GOQD was believed to increase the surface charge density. These polymerization process, MPD needed to penetrate the GOQD layer
oxygen-containing groups also provided excellent hydrophilicity to react with TMC. The steric hindrance of the GOQD layer
to GOQD. Accordingly, larger amounts of GOQD in the PA layer decreased the diffusion of MPD and thereby retarded the forma-
endowed the TFN membrane with higher hydrophilicity.26 A tion of the PA layer. Moreover, GOQD were squeezed out from the
similar phenomenon was also observed for the GO-incorporated layer to form agglomerates, which resulted in a signicant
TFN membranes (Fig. S2 in the ESI†). increase in RMS roughness. The introduction of GOQD led to
Fig. 4 shows the morphologies and structures of the prepared additional nanocorridors, reduced the thickness of polyamide
membranes as characterized by SEM and AFM. The TFC and increased its roughness, which promoted the permeation of
membrane exhibited a rough ridge-and-valley structure (Fig. 4A), water through the membrane.
which indicated the successful formation of a PA active layer.54 As When GO was used in place of GOQD, at surface
shown in Fig. 4B and C, the GOQD-incorporated TFN membranes morphologies were observed in SEM and AFM images (Fig. S4 in
exhibited a different surface morphology. As the GOQD content the ESI†). GO was of a larger size than GOQD and therefore
increased in the TFN membrane, the upper surface appeared to deposition of a GO layer on the PSf substrate resulted in more
have more protruding domains with a blocky shape, which were severe steric hindrance. The GO layer blocked the diffusion of
composed of grainy structures. As shown in the cross-sectional MPD such that a poor-quality PA layer was formed and
SEM images (Fig. 4D–F), GOQD impeded the diffusion of MPD a resulting at surface morphology was found.26
molecules, which resulted in a lower thickness of the polyamide
layer.
The surface morphology and RMS roughness of the prepared 3.3 Membrane performance
RO membranes were also investigated using AFM. Evidently, both To determine the effects of GOQD on RO performance, a series
TFC and TFN membranes possessed rough upper surfaces. of prepared TFN membranes were tested using 2 g L1 NaCl as

16900 | J. Mater. Chem. A, 2016, 4, 16896–16905 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper Journal of Materials Chemistry A

the feed solution at 16 bar. As shown in Fig. 5A, the TFC in water treatment and desalination applications remains
membrane exhibited a permeate ux of 24.7 L m2 h1 (LMH) a challenge. Fig. 6A illustrates the water ux and salt rejection of
with a NaCl rejection of 98.5%. The permeate ux of the TFN the TFC and TFN-GOQD5.0 membranes during a 120 h
membrane increased with an increase in GOQD content and continuous RO test. Both TFC and TFN-GOQD5.0 membranes
the NaCl rejection was maintained at a high level (>98.5%). exhibited very stable uxes and rejections. Fig. 6B shows
TFN-GOQD5.0 achieved a ux of 37.5 LMH, which represented photographs of the TFC and TFN-GOQD5.0 membranes before
a 51.8% increase in comparison with that of the TFC membrane and aer RO operation. As shown, there was hardly any differ-
without GOQD. The NaCl rejection of TFN-GOQD5.0 was 98.8%. ence in the morphologies of these membranes aer the RO test.
This increase in permeate ux was attributed to the addition of This demonstrated that a robust PA layer was formed on the
GOQD, which affected the water permeability of the membrane TFN-GOQD5.0 membrane. However, the active layer of the
in the following ways. Firstly, the incorporation of GOQD GO-incorporated membranes peeled away from the top of the
improved the hydrophilicity of the membrane.55 Secondly, membranes aer the RO test (Fig. S6 in the ESI†). The poor
nanochannels between GOQD nanosheets served as a perme- stability of the TFN-GO membranes further demonstrated that
ation route, which shortened the paths of water through the an unintegrated PA layer had been formed as a result of the
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

membrane.56 Moreover, some voids were unavoidably intro- severe steric hindrance of GO, which strongly hindered the
duced into the thin lm layer at the surface of GOQD. These diffusion of MPD.
nanocorridors also facilitated the permeation of water through
the membrane.57 In addition, the greater roughness also played 3.4 Antifouling capacity
an important role in increasing water ux as a result of the
larger effective contact area between water molecules and the Membrane fouling is one of the main issues that affect the
membrane. However, the incorporation of GOQD did not efficiency of membranes during their use in RO. The onset of
provide high selectivity and the dense PA layer was the only fouling causes a sharp decline in ux and increases the main-
determining factor of the membrane's salt permeability.30,58 In tenance costs. To investigate the antifouling capacity of the
other words, the incorporation of a selected amount of GOQD prepared membranes, the TFC and TFN-GOQD5.0 membranes
(less than 5.0 mg) maintained an integrated PA layer on top of were assessed for RO fouling and cleaning employing bovine
the TFN membrane. serum albumin (BSA) as a model protein foulant and humic
A further increase in GOQD in the membrane (TFN- acid (HA) as a natural organic foulant.59 To ensure that the
GOQD10.0) resulted in a decline in the NaCl rejection to 81.0%. transverse hydrodynamic force was identical for the membrane
This phenomenon agreed well with the solute permeability (B) fouling, the fouling tests were conducted using the same initial
shown in Fig. 5B, where TFN-GOQD10.0 exhibited an extremely permeate ux of 24.7 LMH. As shown in Fig. 7A and C, the uxes
high solute permeability. At higher GOQD contents, it was dramatically decreased when BSA or HA was added to the feed
found that GOQD aggregated and prevented the diffusion of solutions and reached a steady value aer a 21 h fouling test.
MPD, which resulted in the formation of defects in the PA layer. Subsequently, the recovered uxes were measured aer the
Both water and the solute could penetrate the membrane fouled membranes were rinsed. The results of the fouling tests
through these defects. indicated that the incorporation of GOQD signicantly
As a control, the RO performance of GO-incorporated TFN improved the antifouling capacity of the RO membranes. As
membranes is shown in Fig. S5 (ESI†). As can be seen, the shown in Fig. 7B, the TFC membrane exhibited a total fouling
permeate ux increased with an increase in GO content, but the rate (Rt) for BSA of 41.7% and a recovery rate (FRR) of 83.8%.
salt rejection dramatically decreased. This is consistent with the The TFN-GOQD5.0 membrane exhibited higher BSA resistance
SEM images of this membrane shown in Fig. S4 (ESI†), where with an Rt of 29.2% and an FRR of 93.5%. A similar trend was
only a low-quality PA layer was formed on the GO/MPD cushion. observed in the fouling test that used HA (Fig. 7D). The TFC
Membrane stability is of great importance in RO applica-
tions. The long-term stability of membranes that contain
carbon materials (such as CNT, GO, GOQD, and so forth) for use

Fig. 6 Long-term separation performance of the TFC and TFN-


GOQD5.0 membranes as a function of time using 2 g L1 NaCl solution
at 16 bar and 25  C. (A) Water flux and salt rejection, (B) photographs of
Fig. 5 Performance of the prepared membranes with different GOQD TFC (B1 and B3) and TFN-GOQD5.0 (B2 and B4) membranes before
contents. (A) RO flux (J) and salt rejection (R), and (B) water perme- and after the RO test.
ability (A) and solute permeability (B).

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. A, 2016, 4, 16896–16905 | 16901
View Article Online

Journal of Materials Chemistry A Paper

foulants from the surface during the rinsing process and led to
better ux recovery.61 Furthermore, both BSA and HA contain
negative charges at a pH of 7.62 This increased the amount of
negative surface charge in the TFN-GOQD5.0 membrane and
produced stronger electrostatic repulsion between the foulants
and the TFN membrane.63

3.5 Chlorine resistance


In a practical RO process, a chlorine detergent is commonly
used as a disinfectant to control biofouling and/or as
a membrane-cleaning agent for chemical washing. However,
the PA matrix is vulnerable to active chlorine, which will reduce
the salt rejection of the membrane and degrade the selective
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

layer.64 To investigate the chlorine resistance of the prepared


membranes, the TFC and TFN-GOQD5.0 membranes were
tested by comparing the changes in salt rejection aer exposing
the membranes to 100–5000 ppm h chlorine (Fig. 8A). The salt
Fig. 7 Antifouling behavior of the TFC and TFN-GOQD5.0 rejection of the TFC membrane signicantly decreased from
membranes in RO processes. (A) Time-dependent flux for BSA, (B)
fouling resistance for BSA, (C) time-dependent flux for HA, and (D)
98.5% to 59.4% aer exposure to 5000 ppm h active chlorine.
fouling resistance for HA. Chlorination destroyed the hydrogen bonds in the polyamide
chains, which caused severe conformational deformation of the
polyamide chains. This resulted in partial destruction of the
membrane exhibited an Rt of 48.7% and an FRR of 69.2% and polyamide layer and an increase in salt passage.65,66 The
the TFN-GOQD5.0 membrane exhibited an Rt of 32.5% and an TFN-GOQD5.0 membrane exhibited a lesser decline in salt
FRR of 90.0%. rejection from 98.8% to 83.6% aer chlorination at 5000 ppm h.
The improved antifouling properties of the experimental The salt rejection remained higher than 90% aer chlorination at
membranes resulted from the increase in surface hydrophilicity 3000 ppm h. These results indicated that the chlorine tolerance of
on incorporating GOQD.60 Higher surface hydrophilicity can the TFN-GOQD5.0 membrane was improved in comparison with
reduce the adsorption of foulants on the membrane surface, that of the TFC membrane.
thus mitigating surface fouling. In addition, the higher surface To further investigate the chemical changes in the TFC and
hydrophilicity also promoted the desorption of hydrophobic TFN-GOQD5.0 membranes before and aer chlorination,
ATR-FTIR was employed to analyze the characteristic peaks of

Fig. 8 (A) Separation performance of the TFC and TFN-GOQD5.0 membranes, (B) ATR-FTIR spectra of the TFC membrane, (C) ATR-FTIR spectra
of the TFN-GOQD5.0 membrane, and (D) SEM images of the TFC and TFN-GOQD5.0 membranes before and after chlorination.

16902 | J. Mater. Chem. A, 2016, 4, 16896–16905 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper Journal of Materials Chemistry A

band and amide II band for the TFC membrane signicantly


declined as the chlorination became more intensive. Aer chlo-
rination at 5000 ppm h, both peaks were very weak, which
demonstrated the degradation of the PA selective layer.66 For the
TFN-GOQD5.0 membrane, the decrease in the peak intensities of
the amide I band and amide II band was relatively smaller than
that for the virgin TFC membrane, which indicated that the
incorporated GOQD appeared to protect the PA matrix from
Fig. 9 TGA curves of the selective layers separated from the TFC and chlorine.
TFN-GOQD5.0 membranes. Fig. 8D shows surface SEM images of the TFC and TFN-
GOQD5.0 membranes before and aer chlorination. For the
TFC membrane, some of the ridge and valley structures dis-
the membranes at different active chlorine capacities. As shown appeared aer chlorination. When the TFC membrane was
in Fig. 8B and C, the peaks at 1660 cm1 and 1545 cm1 cor- treated with active chlorine at 5000 ppm h, some defects were
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

responded to amide I band C]O stretching vibrations and observed. These were attributed to the collapse of polyamide
amide II band N–H in-plane bending vibrations, respectively. chains or the separation of the polyamide layer from the PSf
These two peaks were sensitive to chlorine and changed substrate. For the TFN-GOQD5.0 membrane, less damage can
remarkably aer chlorination. The intermolecular hydrogen be observed in the SEM images.
bonds between polyamide chains became weaker and were As shown in Fig. 9, the TGA curves of the selective layers of
broken on exposure to chlorine. Therefore, the peak at 1660 cm1 the TFC and TFN-GOQD5.0 membranes provided some
shied to a higher frequency with an increase in active chlorine evidence of the increase in thermal stability on loading with
capacity.65,66 It is worth noting that the intensity of the amide I GOQD. A major weight loss at approximately 500  C was

Table 1 Comparison of performance of GO- and GOQD-containing RO membranes

Flow Feed
Pressure rate (NaCl, J R Chlorine resistance
Approach (bar) (cm s1) ppm) (LMH) (%) Durability Antifouling (NaClO, ppm h)

GO in PSf support, 15.5 22.2 2000 84 98.2 — — —


PA as selective layer33
rGO/TiO2 in PA 15 — 2000 51.3 99.45 — 500 mg L1 BSA, 4000 ppm h, pH ¼ 4,
selective layer 3 h, Rt: 25% R decreased 3%
(rGO/TiO2 dispersed in
aqueous solution of MPD)68
GO in PA selective layer 20.68 — 2000 59.4 93.8 J and R — —
(GO dispersed in were stable
TMC/hexane solution)51 aer 72 h
GO in PA selective layer 15.5 — 2000 16.6 99 — Diluted PAO1 in 48 000 ppm h,
(GO dispersed in 1/100 TSB, 24 h, pH ¼ 7,
aqueous solution of MPD)26 attached biovolume R decreased <1%
of cells decreased 98%
CNTa/GO in PA selective 15.5 8.4 2000 58.96 96.2 J decreased — 20 000 ppm h,
layer (CNTa/GO dispersed 15% aer R decreased 10%
in aqueous solution of MPD)29 72 h
GO/aGO LbL on PA30 15.5 — 2000 14.0 96.4 — 100 ppm BSA, 6000 ppm h,
12 h, Rt: 15% pH ¼ 10, R declined
by <4%;
60 000 ppm h,
pH ¼ 10, R decreased
>80%
aPES/aGO/GO LbL 55 — 32 000 28.4 98.4 — — 300 ppm NaClO,
assembly69 R decreased 14.4%
GO covalently 27.6 21.4 2925 41.4 97.8 — 107 units per mL E. coli —
bonded to PA27 in contact with membrane
for 1 h, number of viable
E. coli cells decreased 64.5%
GOQD in PA selective 16.0 17.0 2000 37.5 98.8 J and R were 500 ppm BSA, 21 h, Rt 29.2%, 5000 ppm h, pH ¼ 4,
layer (GOQD/MPD stable aer FRR 93.5%; 500 ppm HA, 21 h, R decreased 15%
ltered onto PSf) 120 h Rt: 32.5%, FRR 90.0%
(present work)

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. A, 2016, 4, 16896–16905 | 16903
View Article Online

Journal of Materials Chemistry A Paper

observed for both membranes, which was consistent with the Notes and references
thermochemical decomposition of polyamides.67 The mass of
the polyamide layer of the TFC membrane continued to 1 World Economic Forum, Global risks, 2015, Geneva, 2015.
decrease gradually with an increase in temperature, whereas the 2 M. Elimelech and W. A. Phillip, Science, 2011, 333, 712–717.
mass of the selective layer of TFN-GOQD5.0 was almost 3 A. G. Fane, R. Wang and M. X. Hu, Angew. Chem., Int. Ed.,
unchanged. Therefore, the incorporation of GOQD increased 2015, 54, 3368–3386.
both the chlorine tolerance and the thermal stability of the TFN 4 G. Liu, W. Jin and N. Xu, Chem. Soc. Rev., 2015, 44, 5016–5030.
membranes. 5 B.-H. Jeong, E. M. V. Hoek, Y. Yan, A. Subramani, X. Huang,
It was known that GOQD hindered the diffusion of organic G. Hurwitz, A. K. Ghosh and A. Jawor, J. Membr. Sci., 2007,
molecules; therefore, they could also block active chlorine.9 The 294, 1–7.
reduced diffusion rate of OCl ions favored the increase in 6 G. L. Jadav and P. S. Singh, J. Membr. Sci., 2009, 328, 257–267.
chlorine resistance. In addition, hydrogen bonding between 7 E.-S. Kim, G. Hwang, M. Gamal El-Din and Y. Liu, J. Membr.
GOQD and the polyamide layer made the polymer chains more Sci., 2012, 394–395, 37–48.
compact and inhibited the replacement of amidic hydrogens, 8 E.-S. Kim and B. Deng, J. Membr. Sci., 2011, 375, 46–54.
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

which contributed to both chlorine tolerance and thermal 9 H. Zhao, S. Qiu, L. Wu, L. Zhang, H. Chen and C. Gao, J.
stability.26 Another possibility lay in the protection of active sites Membr. Sci., 2014, 450, 249–256.
in residual MPD by electron-rich GOQD from being attacked by 10 W. J. Lau, S. Gray, T. Matsuura, D. Emadzadeh, J. P. Chen
active chlorine. Charge exclusion could impede the contact of and A. F. Ismail, Water Res., 2015, 80, 306–324.
OCl with the membrane surface and thus prevent active 11 J. Yin and B. Deng, J. Membr. Sci., 2015, 479, 256–275.
chlorine from disrupting the PA selective layer. 12 F. Perreault, A. F. de Faria and M. Elimelech, Chem. Soc. Rev.,
Table 1 summarizes the performance of GO- and GOQD- 2015, 44, 5861–5896.
containing RO membranes reported in the recent literature. 13 Z. P. Smith and B. D. Freeman, Angew. Chem., Int. Ed., 2014,
The TFN-GOQD5.0 membrane exhibited comparable ux and 53, 10286–10288.
excellent NaCl rejection. It was worth mentioning that the 14 H. Huang, Y. Ying and X. Peng, J. Mater. Chem. A, 2014, 2,
TFN-GOQD5.0 membrane exhibited a very stable ux and rejec- 13772–13782.
tion during the extended RO test, which suggested outstanding 15 L. Huang, M. Zhang, C. Li and G. Shi, J. Phys. Chem. Lett.,
durability during membrane processes. In addition, the TFN 2015, 6, 2806–2815.
membrane prepared in the present work possessed excellent 16 H. M. Hegab and L. Zou, J. Membr. Sci., 2015, 484, 95–106.
antifouling capacity and chlorine resistance, which are greatly 17 P. S. Goh and A. F. Ismail, Desalination, 2015, 356, 115–128.
desired in membrane desalination and water reclamation 18 P. Sun, K. Wang and H. Zhu, Adv. Mater., 2016, 28, 2287–
processes. 2310.
19 S. C. O'Hern, D. Jang, S. Bose, J.-C. Idrobo, Y. Song, T. Laoui,
J. Kong and R. Karnik, Nano Lett., 2015, 15, 3254–3260.
4 Conclusions 20 S. C. O'Hern, M. S. H. Boutilier, J.-C. Idrobo, Y. Song, J. Kong,
T. Laoui, M. Atieh and R. Karnik, Nano Lett., 2014, 14, 1234–
In summary, a novel GOQD-incorporated TFN membrane was
1241.
developed by pressure-ltering a GOQD/MPD cushion layer onto
21 S. P. Surwade, S. N. Smirnov, I. V. Vlassiouk, R. R. Unocic,
a PSf substrate prior to interfacial polymerization. In compar-
G. M. Veith, S. Dai and S. M. Mahurin, Nat. Nanotechnol.,
ison with a reference TFC membrane, the TFN membrane
2015, 10, 459–464.
exhibited a higher permeate ux and maintained an excellent
22 Y. Han, Z. Xu and C. Gao, Adv. Funct. Mater., 2013, 23, 3693–
NaCl rejection efficiency. Furthermore, the TFN membrane
3700.
possessed outstanding stability, excellent antifouling capacity,
23 C.-H. Tsou, Q.-F. An, S.-C. Lo, M. De Guzman, W.-S. Hung,
and improved chlorine resistance. The GOQD-incorporated TFN
C.-C. Hu, K.-R. Lee and J.-Y. Lai, J. Membr. Sci., 2015, 477,
membrane has the potential for applications in membrane
93–100.
ltration processes such as reverse osmosis, forward osmosis,
24 L. He, L. F. Dumee, C. Feng, L. Velleman, R. Reis, F. She,
and nanoltration on the basis of its high water permeability
W. Gao and L. Kong, Desalination, 2015, 365, 126–135.
and solute selectivity, as well as its antifouling capacity and
25 S. Bano, A. Mahmood, S.-J. Kim and K.-H. Lee, J. Mater.
chlorine tolerance.
Chem. A, 2015, 3, 2065–2071.
26 H.-R. Chae, J. Lee, C.-H. Lee, I.-C. Kim and P.-K. Park, J.
Acknowledgements Membr. Sci., 2015, 483, 128–135.
27 F. Perreault, M. E. Tousley and M. Elimelech, Environ. Sci.
The authors are grateful for the nancial support from the Technol. Lett., 2014, 1, 71–76.
National Natural Science Foundation of China (21476219) and 28 F. Perreault, A. F. de Faria, S. Nejati and M. Elimelech, ACS
the Science and Technology Development Plan of Shandong Nano, 2015, 9, 7226–7236.
province (2014GSF116006). We would like to thank Qingdao 29 H. J. Kim, M.-Y. Lim, K. H. Jung, D.-G. Kim and J.-C. Lee, J.
Haida Haixi New Materials Co., Ltd (China) for providing Mater. Chem. A, 2015, 3, 6798–6809.
GOQD. This is MCTL contribution no. 100.

16904 | J. Mater. Chem. A, 2016, 4, 16896–16905 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper Journal of Materials Chemistry A

30 W. Choi, J. Choi, J. Bang and J.-H. Lee, ACS Appl. Mater. 49 F.-Y. Zhao, Q.-F. An, Y.-L. Ji and C.-J. Gao, J. Membr. Sci.,
Interfaces, 2013, 5, 12510–12519. 2015, 492, 412–421.
31 J. Chen, Y. Li, L. Huang, N. Jia, C. Li and G. Shi, Adv. Mater., 50 C. Delamarche, D. Thomas, J. P. Rolland, A. Froger,
2015, 27, 3654–3660. J. Gouranton, M. Svelto, P. Agre and G. Calamita, J.
32 P. Sun, F. Zheng, M. Zhu, Z. Song, K. Wang, M. Zhong, Bacteriol., 1999, 181, 4193–4197.
D. Wu, R. B. Little, Z. Xu and H. Zhu, ACS Nano, 2014, 8, 51 J. Yin, G. Zhu and B. Deng, Desalination, 2016, 379, 93–101.
850–859. 52 L. Phelane, F. N. Muya, H. L. Richards, P. G. L. Baker and
33 J. Lee, J. H. Jang, H.-R. Chae, S. H. Lee, C.-H. Lee, P.-K. Park, E. I. Iwuoha, Electrochim. Acta, 2014, 128, 326–335.
Y.-J. Won and I.-C. Kim, J. Mater. Chem. A, 2015, 3, 22053– 53 M. Ionita, A. M. Pandele, L. Crica and L. Pilan, Composites,
22060. Part B, 2014, 59, 133–139.
34 P. He, J. Sun, S. Tian, S. Yang, S. Ding, G. Ding, X. Xie and 54 S. Karan, Z. Jiang and A. G. Livingston, Science, 2015, 348,
M. Jiang, Chem. Mater., 2015, 27, 218–226. 1347–1351.
35 C. Wu, C. Wang, T. Han, X. Zhou, S. Guo and J. Zhang, Adv. 55 G. N. B. Baroña, J. Lim, M. Choi and B. Jung, Desalination,
Healthcare Mater., 2013, 2, 1613–1619. 2013, 325, 138–147.
Published on 12 September 2016. Downloaded on 12/27/2020 8:14:31 PM.

36 H.-H. Cho, H. Yang, D. J. Kang and B. J. Kim, ACS Appl. 56 J. Yin, E.-S. Kim, J. Yang and B. Deng, J. Membr. Sci., 2012,
Mater. Interfaces, 2015, 7, 8615–8621. 423, 238–246.
37 Z. Zeng, D. Yu, Z. He, J. Liu, F.-X. Xiao, Y. Zhang, R. Wang, 57 H. Wu, B. Tang and P. Wu, J. Membr. Sci., 2013, 428, 341–348.
D. Bhattacharyya and T. T. Y. Tan, Sci. Rep., 2016, 6, 20142. 58 M. Hu and B. Mi, Environ. Sci. Technol., 2013, 47, 3715–3723.
38 M. Wang, Z. Wang, X. Wang, S. Wang, W. Ding and C. Gao, 59 B. Ma, W. Yu, W. A. Jefferson, H. Liu and J. Qu, Water Res.,
Environ. Sci. Technol., 2015, 49, 3761–3768. 2015, 71, 140–149.
39 W. Ding, J. Cai, Z. Yu, Q. Wang, Z. Xu, Z. Wang and C. Gao, J. 60 M. F. A. Goosen, S. S. Sablani, H. Al-Hinai, S. Al-Obeidani,
Mater. Chem. A, 2015, 3, 20118–20126. R. Al-Belushi and D. Jackson, Sep. Sci. Technol., 2005, 39,
40 S. Zinadini, A. A. Zinatizadeh, M. Rahimi, V. Vatanpour and 2261–2297.
H. Zangeneh, J. Membr. Sci., 2014, 453, 292–301. 61 L.-L. Hwang, J.-C. Chen and M.-Y. Wey, Desalination, 2013,
41 Y. Liu, B. Gao, Z. Qiao, Y. Hu, W. Zheng, L. Zhang, Y. Zhou, 313, 166–175.
G. Ji and G. Yang, Chem. Mater., 2015, 27, 4319–4327. 62 Q. She, R. Wang, A. G. Fane and C. Y. Tang, J. Membr. Sci.,
42 X. Li, X. Wang, L. Zhang, S. Lee and H. Dai, Science, 2008, 2016, 499, 201–233.
319, 1229–1232. 63 R. Bernstein, S. Belfer and V. Freger, Environ. Sci. Technol.,
43 R. Sekiya, Y. Uemura, H. Murakami and T. Haino, Angew. 2011, 45, 5973–5980.
Chem., Int. Ed., 2014, 53, 5619–5623. 64 J. Sun, L.-P. Zhu, Z.-H. Wang, F. Hu, P.-B. Zhang and
44 L. Wang, Y. Wang, T. Xu, H. Liao, C. Yao, Y. Liu, Z. Li, B.-K. Zhu, Sep. Purif. Technol., 2016, 157, 112–119.
Z. Chen, D. Pan, L. Sun and M. Wu, Nat. Commun., 2014, 65 G.-D. Kang, C.-J. Gao, W.-D. Chen, X.-M. Jie, Y.-M. Cao and
5, 5357. Q. Yuan, J. Membr. Sci., 2007, 300, 165–171.
45 J. Zhao, Y. Zhu, F. Pan, G. He, C. Fang, K. Cao, R. Xing and 66 M. Liu, Z. Chen, S. Yu, D. Wu and C. Gao, Desalination, 2011,
Z. Jiang, J. Membr. Sci., 2015, 487, 162–172. 270, 248–257.
46 X. Song, L. Wang, C. Y. Tang, Z. Wang and C. Gao, 67 N. Rakhshan and M. Pakizeh, Korean J. Chem. Eng., 2015, 32,
Desalination, 2015, 369, 1–9. 2524–2533.
47 Y. Zhou, Q. Bao, L. A. L. Tang, Y. Zhong and K. P. Loh, Chem. 68 M. Safarpour, A. Khataee and V. Vatanpour, J. Membr. Sci.,
Mater., 2009, 21, 2950–2956. 2015, 489, 43–54.
48 S. N. Ibragimova, K. B. Stibius, P. P. Szewczykowski, M. Perry, 69 S. G. Kim, D. H. Hyeon, J. H. Chun, B. H. Chun and
H. Bohr and C. H. Nielsen, Biophys. J., 2010, 98, 604A. S. H. Kim, Desalin. Water Treat., 2013, 51, 6338–6345.

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. A, 2016, 4, 16896–16905 | 16905

You might also like