You are on page 1of 12

Desalination 543 (2022) 116116

Contents lists available at ScienceDirect

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

PES mixed-matrix membranes incorporating ZIF-8@MXene nanocomposite


for the efficient dye/salt separation
Yu-yu Yao , Ting Wang , Li-guang Wu *, Hua-li Chen
School of Environmental Science and Engineering, Zhejiang Gongshang University, China

H I G H L I G H T S

• ZIF-8@Ti3C2Tx composites were prepared by reverse microemulsion as a nanoreactor.


• Ti3C2Tx could improve the dispersion of ZIF-8 in PES MMMs.
• Blending with ZIF-8@Ti3C2Tx enhanced the antifouling property of PES MMMs.
• PES MMMs incorporating ZIF-8@Ti3C2Tx have high flux in high salinity wastewater.
• ZIF-8@Ti3C2Tx/PES MMMs could effectively remove dye from high salinity wastewater.

A R T I C L E I N F O A B S T R A C T

Keywords: To improve the flux, anti-fouling and dye separation performance of PES-based membranes, PES mixed-matrix
PES mixed-matrix membrane membranes (MMMs) blended with ZIF-8@Ti3C2Tx composites were fabricated via a facile one-pot micro­
ZIF-8 emulsion strategy. The introduction of hydrophilic components in the microemulsion into the PES matrix
MXene
enhanced the pore structures and surface polarity of the resulting MMMs, causing them to exhibit high
Anti-fouling properties
Dye removal from high-salinity wastewater
permeation and anti-fouling performance compared to both neat PES and MMMs with ZIF-8. The optimal
membrane flux and anti-fouling performance were found to be 280 L⋅m− 2⋅h− 1 and 91 % (bovine serum albumin
rejection), respectively, at 3 % microemulsion content. The well-distributed ZIF-8@Ti3C2Tx composites in the
MMMs interact with dye molecules, which strengthens the dye retention from high-salinity wastewater. Further,
the pore size of the MMMs was positively correlated with the flux but negatively correlated with dye retention. In
particular, the M-3 membrane exhibited high flux (140 L⋅m− 2⋅h− 1) and up to 98 % dye rejection in high-salinity
wastewater, but its rejection of salt ions was <5 %. In addition, the dye rejection recovery rate of M-3 after each
washing exceeded 93 % over 5 fouling/backwashing cycles. These results strongly suggest that MMMs incor­
porated with ZIF-8@Ti3C2Tx composites have promising application prospects for removing trace dye pollutants
from high-salinity wastewater.

1. Introduction offers high separation efficiency towards trace pollutants. In particular,


due to ease of operation, ultrafiltration membrane separation processes
Over the last few decades, environmental pollution has increased are widely used in the advanced treatment of drinking water. For such
dramatically due to rapid industrial development and population processes, ultrafiltration membranes with high flux, good solute reten­
growth. Water pollution in particular is a cause for grave concern, as this tion, and excellent antifouling properties are required. The polymers
directly impacts human health [1,2]. The improvement of water quality, polyvinylidene fluoride (PVDF) and polyethersulfone (PES) are widely
especially for potable water, is highly dependent on the removal of trace used to construct ultrafiltration membranes owing to their good chem­
pollutants. As such, the effective treatment of trace pollution is a ical and mechanical stability, oxidation resistance, and biocompatibility
pressing issue in environmental and water supply and drainage engi­ [3]. However, as both PVDF and PES are typical hydrophobic polymers,
neering. Currently, one of the most promising methods for removing their pores often adsorb organic or protein pollutants, which causes
pollutants in water is through membrane separation, as this technique deposition and blockage during the ultrafiltration process and

* Corresponding author.
E-mail address: wulg64@hotmail.com (L.-g. Wu).

https://doi.org/10.1016/j.desal.2022.116116
Received 8 July 2022; Received in revised form 22 August 2022; Accepted 10 September 2022
Available online 21 September 2022
0011-9164/© 2022 Elsevier B.V. All rights reserved.
Y.-y. Yao et al. Desalination 543 (2022) 116116

ultimately leads to the fouling and flux attenuation of the ultrafiltration 2. Experimental
membrane. Therefore, it is common to add adding hydrophilic materials
to fabricate blended or modified membranes with improved hydro­ 2.1. Materials
phobic properties compared to unmodified PVDF or PES [4,5]. Some of
the common additive nanomaterials used to fabricate blended mem­ PES (Ultrason E6020P, Mw = 58,000 Da) was obtained from BASF
branes are inorganic metal nanoparticles [6,7], carbon-based nano­ (Germany) and dried for 24 h at 60 ◦ C prior to use. Ti3C2Tx was pur­
materials [8], porous materials (zeolites and molecular sieves) [9], and chased from Xianfeng Nano Co., Ltd. Analytical-grade Zn(NO3)2⋅6H2O,
metal-organic frameworks (MOFs) [10]. ethanol, and N,N-dimethylacetamide (DMAC, AR) were purchased from
Due to their large hydrophilic specific surface, MOFs have been the Sinopharm Chemical Reagent Co., Ltd. 2,2-Azobisisobutyronitrile
widely used in the construction of blended membranes, and have been (AIBN, >98 %), 1-vinyl-3-ethylimidazolium tetrafluoroborate
shown to significantly enhance the anti-fouling and separation proper­ (VeimBF4, ≥98.5 %), 1-butyl-3-methylimidazolium (BmimOTF, >97
ties of PVDF or PES ultrafiltration membranes. For instance, Moutloali %), Rhodamine B (RhB, AR), Congo Red (CR, AR), Methylene Blue (MB,
et al. [11,12] prepared Ag@ZIF-8@GO, Cu@ZIF-8@GO, Ag-Cu@ZIF- AR) and Methyl Orange (MO, AR) were purchased from Macklin. Bovine
8@GO, and Cu(tpa)@GO/PES mixtures via solution blending, finding Serum Albumin (BSA, Mw = 66,500 Da), triethylamine (AR), and 2-
that the hydrophilicity of the resulting mixed-matrix ultrafiltration methylimidazole (2-MeIM, 98 %) were provided by Aladdin. Na2SO4
membranes were significantly enhanced due to the hydrophilicity and was purchased from Shanghai Lingfeng Chemical Reagent Co., Ltd.
porous structures of GO, ZIF-8, and Cu(tpa). As a result, the pure water
flux (Jw) increased by over 2.5-fold compared with the pure membrane. 2.2. Synthesis of ZIF-8 and ZIF-8@Ti3C2Tx nanocomposites
In a previous study [13], ZIF-8 was prepared using a polymerizable ionic
liquid as the oil phase in a reverse microemulsion, which was then Ionic liquid microemulsions of Zn2+ and 2-MeIM solution were
blended with PES to form mixed-matrix membranes (MMMs). In inverse prepared following a method from the literature [14] and were denoted
microemulsions, the water core in the micelle acts as a micro- as microemulsion A (composed of VeimBF4, BmimOTF, and Zn
nanoreactor to generate ZIF-8 nanoparticles. An initiator is then added (NO3)2⋅6H2O) and microemulsion B (composed of VeimBF4, BmimOTF,
to the inverse microemulsion to form a latex via in-situ prepolymeri­ and 2-MeIM). The concentrations of Zn(NO3)2⋅6H2O and 2-MeIM were
zation, which maintains the ZIF-8 nanoparticles in the micelles and 1 mol⋅L− 1 and 4 mol⋅L− 1, respectively. ZIF-8 nanoparticles were
ensures their even dispersion in the latex. In the subsequent preparation generated in microemulsion D through mixing microemulsions A and B,
process, the latex can be uniformly blended with PES to ensure that the while holding the mass percentages of all other components in the in­
ZIF-8 nanoparticles in the latex also maintain good dispersibility in the verse microemulsion constant, as follows: water, 1 wt%; BmimOTF, 4 wt
PES matrix. Hence, the addition of only a small amount of ZIF-8 material %; and VeimBF4, 6 wt% [13].
(1.25 % or less) was found to significantly improve the water flux and ZIF-8@Ti3C2Tx composites were synthesized as follows: first, Ti3C2Tx
anti-fouling properties of the MMMs. The highest water flux and flux (270 mg) was dispersed into 50 mL of DMAC, before adding micro­
recovery rate of MMMs were 10 times of those of neat PES, at 105 emulsion A to the Ti3C2Tx solution and stirring at room temperature for
L⋅m− 2⋅h− 1 and 98 %, respectively. However, it was also found that if the 1 h. Microemulsion B was then added, and, after stirring for 3 h at room
amount of ZIF-8 added was >1.25 %, agglomeration occurred due to temperature. ZIF-8@Ti3C2Tx nanocomposites were generated in micro­
insufficient affinity between the surface hydrophilic ZIF-8 and the hy­ emulsion C through mixing microemulsions A, B, and Ti3C2Tx solution,
drophobic PES, thereby worsening the water flux and anti-fouling and the resulting composite was centrifuged and washed several times
properties of the resulting membranes. with fresh methanol. Finally, the product was dried in a vacuum oven at
Against this backdrop, the present study demonstrates the design and 60 ◦ C. Fig. 1 shows a schematic of the microemulsion C synthesis
construction of a PES mixed-matrix ultrafiltration membrane blended process.
with ZIF-8@Ti3C2Tx. To this end, the 2D material Ti3C2Tx was added
after the preparation of a microemulsion containing ZIF-8, with its 2.3. Fabrication of MMMs
lamellar structure used to anchor the ZIF-8 nanoparticles and ensure
good dispersibility in the PES polymer. The introduction of large The polymerization process was initiated through the addition of
numbers of hydrophilic groups through blending with well-distributed AIBN (2,2-Azobisisobutyronitrile, 0.5 wt% of total oil phase) to micro­
ZIF-8@Ti3C2Tx composites is expected to significantly improve the hy­ emulsion C with ZIF-8@Ti3C2Tx nanoparticles at 60 ◦ C [13]. Once the
drophilicity and polarity of the resulting membranes, thereby increasing viscosity of microemulsion C reached 0.50 Pa⋅s, it was determined to
their water flux and anti-fouling properties. The performance of the have been polymerized and was labelled as microemulsion PC. The
prepared membranes towards the removal of trace dye pollutants from polymerization process is illustrated in Fig. 2.
high-salinity wastewater was also investigated. It is generally well Next, the ZIF-8@Ti3C2Tx/PES MMM samples were fabricated by
accepted that, although nanofiltration can effectively separate trace blending PES with the microemulsion containing ZIF-8@Ti3C2Tx nano­
organic pollutants in high-salinity systems, such processes typically composites. To prepare the different compositions of membrane-casting
require high operating pressures and increased costs due to the solution (as listed in Table 1), the different amounts of the micro­
extremely small pore structures of nanofiltration membranes. Although emulsion were added dropwise under stirring for 6 h at 60 ◦ C to obtain a
using ultrafiltration separation for dye removal can potentially over­ membrane-casting solution. After complete degassing under a vacuum
come the problems of operating pressure and cost, improving the for 24 h, the solution was uniformly cast onto clean glass plates which
retention rate of dye molecules is also key to successfully applying ul­ were subsequently immersed in deionized water for 30 min. The mem­
trafiltration to the advanced treatment and reuse of high-salinity branes were peeled from the glass plates and immersed in distilled water
wastewater. Therefore, in the present study, it is expected that the for 24 h to remove the residual DMAC. Fig. 3 shows a schematic of the
presence of well-distributed ZIF-8@Ti3C2Tx in the PES mixed-matrix membrane preparation process.
ultrafiltration membranes will enhance the retention for dye pollut­
ants through interactions with the dye molecules. 2.4. Characterization

Transmission electron microscopy (TEM) (JEM-1230, Jeol, Japan)


and X-ray diffraction (XRD) (D8 Advance, Germany), with a Cu Kα
source and a wavelength of 1.5406 Å, were used to analyze the mor­
phologies and crystallinity of ZIF-8@Ti3C2Tx. X-ray photoelectron

2
Y.-y. Yao et al. Desalination 543 (2022) 116116

Fig. 1. Flow chart of the microemulsion synthesis of ZIF-8@Ti3C2Tx.

Fig. 2. Schematic diagram of microemulsion polymerization.

conducted using a contact angle goniometer (SL200B, Kono, USA) to


Table 1
characterize the hydrophilicity of the ultrafiltration membranes. The
Compositions of the different MMM blends.
overall porosity (ε) and mean pore radius (rm) of the membranes were
Membranes Amount of Amount of PES/ Amount of calculated using the gravimetric method [15] and Guerout–Elford–Ferry
microemulsion/g g DMAC/g
equation [16] based on the Jw and porosity data, respectively.
Neat PES 0 15 85
M-1 0.25 15 84.75
2.5. Penetration, antifouling and separation properties of MMMs
M-2 0.5 15 84.5
M-3 1 15 84
M-4 2 15 83 The pure water permeation and antifouling properties of the MMMs
M-5 3 15 82 were tested using an MSC-300 ultrafiltration cup with a volume of 0.3 L
M-6 4 15 81 and an effective area of 34.2 cm2. Each membrane was compacted at
0.15 MPa for 30 min to achieve a steady flux, before being operated at
spectroscopy (XPS) (ESCALAB 250, Thermo Fisher Scientific, USA) and 0.1 MPa. The pure water flux, Jw, BSA permeation flux, Jp, and recovery
scanning electron microscopy (SEM) (S-4700, Hitachi, Japan) were used water flux, JR, were calculated as follows:
to characterize the morphology, structure, surface roughness, and VW
elemental composition of the membranes. Fourier transform infrared JW = (1)
A×t
(FT-IR) spectroscopy was conducted using a Nexus-670 spectrometer
(Nicolet, USA). Water contact angle (WCA) measurements were JP =
VP
(2)
A×t

3
Y.-y. Yao et al. Desalination 543 (2022) 116116

Fig. 3. Schematic diagram of MMM preparation.

Vr where Vx is the volume of the permeated solution, A is the area of the


JR = (3)
A×t membrane (m2), and t is the operating time (h).
To evaluate the anti-fouling properties of the membranes, the BSA

ƺ a b

ZIF-8@Ti3C2Tx
ƺ
Transmittance (%)
Intensity (a.u)

ƺ ƺ
Ti3C2Tx
C=N
ZIF-8 C-H C-N
ƺ SO3H
ƿ ZIF-8@Ti3C2Tx ZIF-8
ƿ ƿ ƿ ƿ
ƿ ƿ Ti3C2Tx
ƿ ƿ ƿ

4000 3500 3000 2500 2000 1500 1000 500


10 20 30 40 50 60 70
2θ (°) Wavenumber (cm-1)

ZIF-8@Ti3C2Tx Zn2p3/2 c Ti3C2Tx d


Ti2p3/2
ZIF-8 ZIF-8@Ti3C2Tx Ti2p1/2
Zn2p1/2
Ti3C2Tx ZIF-8
Intensity (a.u.)
Intensity (a.u.)

6eV
23eV

1050 1045 1040 1035 1030 1025 1020 1015 475 470 465 460 455 450
Binding Energy (eV) Binding Energy (eV)
Fig. 4. XRD, XPS and FT-IR analysis of ZIF-8@Ti3C2Tx nanoparticles: (a) XRD patterns of ZIF-8, Ti3C2Tx and ZIF-8@Ti3C2Tx; (b) FT-IR spectra; (c) Zn 2p spectra; (d)
Ti 2p spectra.

4
Y.-y. Yao et al. Desalination 543 (2022) 116116

rejection rate, R, flux recovery ratio (FRR) of pure water, total pollution Fig. S1 shows the XPS spectrum of the ZIF-8@Ti3C2Tx composites,
rate, Rt, reversible pollution rate, Rr, and irreversible pollution rate, Rir, along with those of single ZIF-8 and Ti3C2Tx for comparison. The results
were also calculated, as follows: show that both Zn and Ti elements were present in the ZIF-8@Ti3C2Tx
composites. In the Zn2p and Ti2p spectra of ZIF-8@Ti3C2Tx composites
C0 − C1
R= × 100% (4) (Fig. 4c and d), the peaks at 1023 eV (Zn2p3/2) and 1046 eV (Zn2p1/2)
C0
correspond to Zn2+ (Fig. 4c), and the two main peaks at 455 eV and 461
JR eV correspond to the Ti2p3/2 and Ti2p1/2 spin orbit peaks of Ti4+ in
FRR = × 100% (5) Ti3C2Tx (Fig. 4d) [21].
JW
The SEM and TEM images of ZIF-8@Ti3C2Tx nanocomposites are
JW − JP shown in Figs. 5 and S2, respectively. For comparison, the SEM and TEM
Rt = × 100% (6)
JW images of original Ti3C2Tx are also shown in Figs. 5 and S2. The results
show that Ti3C2Tx has a dense multilayer lamellar structure, which is
Rr =
JR − JP
× 100% (7) consistent with the structure of MXene in most studies from the litera­
JW ture. The images also show that the ZIF-8 particles were anchored on the
surface or interior of the lamellar structure of Ti3C2Tx, and these appear
JW − JR
Rir = × 100% (8) as bright white particles in the SEM images and small black dots in the
JW
TEM images. This further supports the idea that the addition of Ti3C2Tx
in the microemulsion containing ZIF-8 resulted in chemical interactions
where C0 and C1 are the feed and permeate concentrations of BSA so­
between the two substances.
lution, respectively.
The cation dye Rhodamine B (RhB) (at a concentration of 10 mg/L)
and the salt Na2SO4 (at concentrations of 1 %, 2.5 %, 5 %, and 10 %) 3.2. Morphology of different MMMs containing ZIF-8@Ti3C2Tx
were used to determine the separation performance of the membrane
towards dye and salt, respectively. After filtration, the absorbance of Figs. S3 and 6 show the XPS spectra and corresponding high-
RhB was measured at a wavelength of 554 nm using the UV photometer, resolution Ti2p and Zn2p profiles of the M-3 and M-5 samples, respec­
and the concentrations of the feed and permeate solutions were calcu­ tively. The results show that the characteristic peaks of Zn and Ti ele­
lated [17], allowing the rejection of RhB to be determined. The retention ments in the XPS spectra in Fig. S3 are very weak due to the small
rate of salt, meanwhile, was determined using conductometers. amount of ZIF-8@Ti3C2Tx present in the two MMMs. As shown by the
high-resolution XPS profiles for Zn2p and Ti2p in M-3 and M-5, the weak
3. Results and discussion peaks at 1022 eV (Zn2p3/2) and 1045 eV (Zn2p1/2) corresponded to Zn2+
[23]. The characteristic peaks of Ti were clearly present in the MMMs,
3.1. Microstructures of ZIF-8@Ti3C2Tx and the two main peaks in the Ti 2p profiles, at binding energies of 455
and 461 eV, correspond to the Ti2p3/2 and Ti2p1/2 of Ti4+ in Ti3C2Tx,
Fig. 4a shows the XRD patterns of the ZIF-8@Ti3C2Tx composite, respectively [22]. It is evident that with the addition of microemulsion,
along with those of single ZIF-8 and Ti3C2Tx for comparison. The results the content of ZIF-8 and Ti3C2Tx in the film increased, which was also
showed that both of the main crystalline diffraction peaks of ZIF-8 and reflected by an increase in the corresponding characteristic peak
Ti3C2Tx were present in the patterns for the ZIF-8@Ti3C2Tx composites, strength of the two elements.
which indicates the successful preparation of the composites [18]. In Figs. S4 and 7 show SEM images of neat PES and the different MMMs.
addition, the diffraction peak intensities of ZIF-8@Ti3C2Tx were much The images indicate that all the samples exhibited a typical asymmetric
weaker than those of single ZIF-8 prepared under the same conditions, structure composed of a top skin layer, finger-like pores, and sponge-like
which suggests that chemical interactions occurred between Ti3C2Tx and pores. After the addition of the ZIF-8@Ti3C2Tx microemulsion, the mean
ZIF-8 [18], as opposed to a purely physical blending of the two pore size of the membrane effectively increased. This phenomenon may
substances. be attributed to the introduction of hydrophilic components through the
Fig. 4b shows that the FT-IR spectrum of Ti3C2Tx has no distinct addition of the microemulsion (consisting of BmimOTF, VeimBF4,
characteristic peak, which is consistent with previous studies [19,20]. In Ti3C2Tx, and ZIF-8 nanoparticles), which accelerated the exchange of
the spectra of both ZIF-8 and ZIF-8@Ti3C2Tx, peaks at 2916, 1557, and the solvent (DMAC) with the non-solvent (deionized water) during
1134 cm− 1 were observed, corresponding to the absorption peaks of the membrane fabrication [24]. As the content of microemulsion in the
C–H, C– – N, and C–N bonds of ZIF-8, respectively [19]. The vibrations membranes increased, however, the sponge-like pores decreased and the
of symmetrically-bonded sulfonate groups were also observed at 1190 uniform finger-like pores increased. At a microemulsion content of 4 wt
and 1149 cm− 1 [13], which further suggests that the Ti3C2Tx was % (Fig. 7f), ZIF-8@Ti3C2Tx nanoparticle agglomerates were observed
chemically blended with ZIF-8 in the ZIF-8@Ti3C2Tx composites. within the membrane, which may be attributed to the excess polymer

a b c

Fig. 5. SEM images for (a) Ti3C2Tx and (b), (c) ZIF-8@Ti3C2Tx.

5
Y.-y. Yao et al. Desalination 543 (2022) 116116

a
b
Ti2p3/2
Ti2p1/2
Zn2p3/2
Intensity (a.u.)

Intensity (a.u.)
Zn2p1/2

470 465 460 455 450 1050 1045 1040 1035 1030 1025 1020
Binding Energy (eV) Binding Energy (eV)

c d
Ti2p3/2
Zn2p3/2

Ti2p1/2
Intensity (a.u.)

Intensity (a.u.)

Zn2p1/2

470 465 460 455 450 1050 1045 1040 1035 1030 1025 1020 1015
Binding Energy (eV) Binding Energy (eV)

Fig. 6. High resolution XPS spectra of MMMs: (a), (c) Ti 2p spectra and (b), (d) Zn 2p spectra.

being unable to blend uniformly with PES. nanocomposites and PES.


To further confirm the conclusions drawn from the SEM images in Fig. 8 illustrates the changes in the pore structure and the water
Fig. 7, an SEM image of an MMM only incorporating Ti3C2Tx is shown in contact angle (WCA) of the MMMs. The results show that the addition of
Fig. S5. From this image, some aggregations were observed in the cross- VeimBF4 and ZIF-8@Ti3C2Tx to the membrane improved hydrophilicity
section of the sample, which demonstrates that ZIF-8 anchored in the and promoted the formation of large pores [26]. As the microemulsion
Ti3C2Tx can improve the distribution of Ti3C2Tx in the PES matrix. FTIR content increased, the mean pore size of the MMMs gradually increased,
analysis of neat PES, M-3, and M-5 was carried out to explore the in­ reaching a maximum value of ~45 nm for sample M-5. When the
teractions between the ZIF-8@Ti3C2Tx nanocomposites and PES, with microemulsion content exceeded 4 wt%, however (sample M-6), the
the results shown in Fig. S6. The FTIR spectra show that the character­ mean pore size decreased dramatically, to only 20 nm. Meanwhile, the
istic peaks of PES appear in all three samples. The vibrations of WCAs of the MMMs showed the opposite trend, reaching a minimum
symmetrically-bonded sulfonate groups appear at approximately 1190 value of 60.62◦ at 3 wt% microemulsion. At a microemulsion content of
and 1149 cm− 1, those of C–O–C bonds appear at 1238 and 1104 cm− 1, 4 wt%, the WCA increased due to ZIF-8@Ti3C2Tx agglomeration and its
and those of aromatic groups appear at 1012 cm− 1 [21]. The charac­ non-uniform blending with PES, which is consistent with the SEM image
teristic peaks of ZIF-8 nanoparticles or Ti3C2Tx were not observed in the shown in Fig. 7f. The change in pore size was further characterized by
FTIR spectra of M-3 and M-5, although the elements Ti (representing measuring the molecular weight cut-off (MWCO) of the blending
Ti3C2Tx) and Zn (representing ZIF-8 nanoparticles) were detected in the membranes based on the rejection coefficient R (%) of polyethylene
XPS analysis. This might be due to the low content and very good glycol (PEG) with average molecular weights (Mw) of 0.2, 0.4, 0.8, 1, 2,
dispersion of ZIF-8@Ti3C2Tx nanocomposites in the MMMs. However, 10, 20, 35, 70, 100, and 150 kDa [25]. The value of MWCO was
the FTIR spectra of M-3 and M-5 do contain a wide sawtooth peak at calculated from the molecular weights of the PEG molecules that were
~3400 cm− 1 (corresponding to the presence of –OH), and no obvious up to 90 % rejected by the membrane, with the results shown in Fig. S7.
absorption peak was present at ~3400 cm− 1 in neat PES [25], which The MWCO values of neat PES, M-3, and M-5 were found to be 1 kDa,
confirms the improvement in the hydrophilicity of the MMMs. This is 100 kDa, and 130 kDa, respectively. The mean pore sizes of the mem­
likely due to the interactions between the ZIF-8@Ti3C2Tx branes, as shown in Fig. 8, followed a similar trend. Fig. 9 shows the zeta

6
Y.-y. Yao et al. Desalination 543 (2022) 116116

a b c

d e f

Fig. 7. SEM images of MMMs blended with different microemulsion content: (a) 0 (neat PES); (b) 0.5 %; (c) 1 %; (d) 2 %; (e) 3 %; (f) 4 %.

120 120 120


Porosity Mean Pore Size neat PES
100 M-3
100 75.53° 70.40° 68.41° 66.84° 63.09° 60.62° 66.74° 100 M-5
80
Zeta Potential (mV)

60
80 80
Mean Pore Size (nm)
Porosity (%)

40

60 60 20

0
40 40
-20

-40
20 20
-60

0 0
neat PES M-1 M-2 M-3 M-4 M-5 M-6 3 5 7
pH
Fig. 8. Effect of microemulsion content on the porosity, average pore diameter,
Fig. 9. Zeta potentials of MMMs with different amounts of microemulsion at
and water contact angle of MMMs with different amounts of microemulsion.
different pH values.

potentials on the surface of the neat PES, M-3, and M-5 membranes. The
the water flux of neat PES is only around 15 L⋅m− 2⋅h− 1, the addition of
results show that adding the microemulsion significantly increases the
DMAC would not significantly affect its performance. In addition,
zeta potential of the MMMs due to the high polarity of the micro­
MMMs with pure microemulsion (without adding substances to the
emulsion components (including VeimBF4 as an oil phase, ZIF-8, and
water phase other than VeimBF4, BmimOTF, and water) were prepared,
Ti3C2Tx).
which had a flux of only around 25 L⋅m− 2⋅h− 1, suggesting that the
microemulsion composition has a weak effect on the membrane per­
3.3. Permeation and anti-fouling performance of different MMMs formance. Unlike DMAC and the microemulsion composition, ZIF-8,
containing ZIF-8@Ti3C2Tx Ti3C2Tx, and ZIF-8@Ti3C2Tx all have significant effects on the perfor­
mance of the MMMs. As Fig. 10a shows, the ZIF-8@Ti3C2Tx MMMs
To explore the key structural factors affecting membrane perfor­ performed the best (due to the synergy between ZIF-8 and Ti3C2Tx [23]),
mance, neat PES, ZIF-8 incorporated MMMs without Ti3C2Tx, and the ZIF-8 MMMs without Ti3C2Tx (1 % ZIF-8) were second best, and the
Ti3C2Tx incorporated MMMs without ZIF-8 (but with VeimBF4, Bmi­ Ti3C2Tx MMMs without ZIF-8 (1 % Ti3C2Tx) showed the worst
mOTF, and water added during preparation) were constructed, with all performance.
other preparation conditions being the same. The performance of these Fig. 10b shows that the flux and FRR of the MMMs both showed
membranes was then evaluated, with the results shown in Fig. 10a. As

7
Y.-y. Yao et al. Desalination 543 (2022) 116116

Fig. 10. Flux, rejection of BSA and flux recovery ratio of: (a) neat PES and various MMMs; (b) MMMs with different microemulsion content.

maximum values (of 280 L⋅m− 2⋅h− 1 and 96 %, respectively) with


Table 2
increasing microemulsion content, with the optimal amount of com­
Comparison between the performance of PES-based MMMs from the literature
posite added being 3 % (sample M-5), which is significantly higher than
with sample M-5 from the present study.
the optimal quantity of ZIF-8 microemulsion (1 %) determined in a
Fillers Flux/ Flux recovery ratio/ Rejection/
previous study. This further indicates that Ti3C2Tx tends to promote the
L⋅m− 2⋅h− 1
% %
distribution of ZIF-8 in the PES matrix. Further, these results clearly
demonstrate that the introduction of well-distributed hydrophilic com­ ZIF-8@GO [11] 68 88 /
ZIF-8 [22] 55 50 99
ponents (including ILs, ZIF-8, and Ti3C2Tx) into PES improves not only
UiO-66-NH2 [30] 22 95 99
the water flux but also the anti-fouling performance of the MMMs. TMU-5 [31] 61 98 98
However, the SEM results (Fig. 7) show that the presence of excess AuL-TiO2 [32] 250 / 80
microemulsion results in agglomeration, which explains the decreased CNC [33] ~50 71.6 83.3
GO@TiO2 [34] 108 86.1 99
performance of sample M-6 [27].
CA-Ag2O [35] 92 / 88.8
To evaluate the stability of the different MMMs, five cycles of BSA CS-APSGO [36] 123 90 98
solution/backwashing were carried out, as shown in Fig. 11a. It is ZIF-8 [13] 105 98 99
evident that all the MMMs exhibited better stability compared with neat ZIF-8@Ti3C2Tx (This 280 96 99
PES, which further confirms the enhancing effect of adding micro­ work)

emulsion on the anti-fouling performance of the MMMs [28,29]. Here


too, sample M-5 had an advantage over the others due to its superior some distance, further indicating an enhancement in properties due to
anti-fouling properties. After four pollution-backwashing cycles, the flux the addition of ZIF-8@Ti3C2Tx.
of M-5 was still as high as 217 L⋅m− 2⋅h− 1 and its FRR also exceeded 78 %
(Fig. 11b).
Table 2 compares the JW, flux recovery rate, and BSA rejection values 3.4. Removal of dye from saline wastewater by MMMs
of sample M-5 with those reported in some recently published articles.
The comparison shows that M-5 had a similar flux recovery rate and BSA Based on the results in Section 3.3, samples M-3 and M-5 were
rejection rate compared to other MMMs, while having the highest JW by chosen to further evaluate the separation performance of the MMMs
containing ZIF-8@Ti3C2Tx through the removal of trace dye pollutants

350
100 80
a neat PES M-3 M-5 Recovery Ratio Inrecovery Ratio b
M-2 M-4 M-6 90 Flux Recovery Ratio
300 70
80
250 60
Flux Recovery Ratio (%)

70
Fouling Ratio (%)
Flux (L/m 2·h)

200 50
60

150 50 40

40
100 30
30
50 20
20
10
0 10
0 60 120 180 240 300 360 420 480 540 600
0 0
Time (min) neat PES M-1 M-2 M-3 M-4 M-5 M-6

Fig. 11. Permeation and anti-fouling properties of different MMMs: (a) circulation flux; (b) reversible fouling, irreversible fouling, and flux recovery ratio.

8
Y.-y. Yao et al. Desalination 543 (2022) 116116

in wastewater samples of varying salinity. The effect of salt concentra­ M-3 was chosen to further evaluate the cyclic stability for separating
tion on the dye and salt rejection rates of the two MMMs over a period of dyes from high-salinity water over five fouling/backwashing cycles
350 min are shown in Figs. 12 and 13 and Table 3. (Fig. 14a). During each experimental cycle, and in accordance with in­
Fig. 12a shows that the fluxes of the M-3 and M-5 membranes are dustrial practice, deionized water was used to reverse wash the
much higher compared with neat PES, which is consistent with the water contaminated membranes five times once the reject decayed by 20–25
flux results from Fig. 10. From Fig. 12b, it is evident that the rejection %. The dye/salt solution was used as the feed to evaluate the reject
rates of the neat PES membrane as well as the MMMs for dyes were much decay of membranes. During each ultrafiltration process, the operating
higher than those for salt ions. In general, dye contaminants exist in pressure was 0.1 MPa and the concentrations of salt and dye in the so­
water as aggregates of 5 to 8 molecules [37], the size of which is larger lution were 2.5 % and 10 mg⋅L− 1, respectively. When backwashing the
than the pore size of the sponges in PES-based membranes, causing the membrane with deionized water, the operating pressure and back­
dye to be retained by the membranes. Fig. 12b also shows that the washing time were 0.2 MPa and 30 min, respectively [40]. As a com­
rejection rates for dyes of M-3 and M-5 were higher compared with neat parison, stability experiments were also carried out on the neat PES
PES under the same conditions. This is likely because the addition of ZIF- membrane, with the results shown in Fig. 14b.
8@Ti3C2Tx composite enhances the retention for dye through in­ Fig. 14 shows that the flux of M-3 was consistently higher than that of
teractions between the dye molecules and the well-distributed ZIF- neat PES over five fouling/backwashing cycles, with neither membrane
8@Ti3C2Tx. To verify these interactions, the adsorption of RhB in high- showing large changes in flux. However, the dye rejection performance
salinity wastewater by ZIF-8@Ti3C2Tx powder, neat PES, and M-5 was of the two membranes was significantly different. Compared with neat
measured, with the results shown in Fig. S9. From the adsorption curves, PES (Fig. 14b), the dye rejection of M-3 (Fig. 14a) decayed more slowly,
it is evident that ZIF-8@Ti3C2Tx powder showed the best selective and its dye rejection recovery rate after each washing remained above
adsorption performance towards RhB, while M-5 showed significantly 93 %. This may be because, although the interactions between ZIF-
higher selective adsorption towards RhB than the neat PES due to the 8@Ti3C2Tx and the dye molecules were to strengthen the dye retention
presence of well-distributed ZIF-8@Ti3C2Tx. These results further of the membrane, this effect was not permanent, and the retained dye
confirm the interactions between the dye molecules and the ZIF- molecules were easily removed by backwashing. To further verify this,
8@Ti3C2Tx. Further, M-5 was found to have a higher flux but a lower digital photographs of the neat PES and M-3 after pollution and after
rejection rate for dyes compared with M-3 under the same conditions, backwashing are shown in Fig. S8. It was found that the retained dye
which is likely due to the larger pore size in the former (as shown by the molecules on the neat PES after pollution were difficult to clean,
structural characterization results in Section 3.2). whereas those on M-3 could easily be removed by backwashing after
Fig. 13 and Table 3 show the effects of salt concentration on the pollution. Therefore, the dye rejection rate of the neat PES membrane
rejection rate for dyes and salts of M-3 and M-5 during 350 min of ul­ decayed rapidly and was difficult to recover, dropping to <78 % after
trafiltration (Fig. 13a and b). The results show that the rejection rate for four cycles. This further demonstrates the promising application pros­
dyes by both membranes decreased significantly with increasing salt pects of ZIF-8@Ti3C2Tx-blended MMMs.
concentration in the wastewater, and that the attenuation of the dye Table 4 lists the flux and the dye and salt concentrations and rejec­
rejection rate was more pronounced as the operating time increased. tion rates of M-3 compared with other membranes from recent studies.
This is due to the accumulation of large numbers of salt ions on the The results show that while the dye and salt rejection rates of M-3 were
membrane surface or in its pores, which weakens the interactions be­ close to the other membranes, its Jw (122 L⋅m− 2⋅h− 1) was much higher.
tween the ZIF-8@Ti3C2Tx composite and the dye molecules [38]. To further verify the separation performance of dye from high-
Fig. 13c and d also show that the rate of flux decay for both membranes salinity wastewater, M-3 was tested with several other dyes in 100
increased significantly with increased salt concentration in the waste­ g⋅L− 1 Na2SO4 solution, with the results listed in Table 5. The data show
water due to the different osmotic pressures on both sides of the mem­ that the MMMs fabricated in this work were able to effectively remove
brane. Further, the flux decay of M-5 is relatively slow, which is possibly all the dyes tested, which confirms their promising application prospects
due to its pore size being relatively large. In contrast, the pore size of M-3 for removing trace dye pollutants from high-salinity wastewater.
is relatively small, and the aggregated dye forms a filter cake layer on the
membrane surface which blocks the membrane pores, causing a large 4. Conclusions
flux attenuation [39].
Based on the fact that it showed the best dye rejection performance, In this study, PES mixed-matrix ultrafiltration membranes blended

350 100 100


a Neat PES b
M-3 90 90
300
M-5 80 80
250 70 70
Salt Rejection (%)
Dye Rejection (%)
Flux (L/m 2·h)

60 60
200 Neat PES
50 M-3 50
150 M-5
40 40

100 30 30

20 20
50
10 10

0 0 0
1% 2.5% 5% 10% 1% 2.5% 5% 10%
Salt concentration Salt concentration

Fig. 12. Performance of neat PES and different MMMs: (a) flux; (b) rejection of dye and salt.

9
Y.-y. Yao et al. Desalination 543 (2022) 116116

110 110 110 110

100 a 100 100 b 100

90 90 90 90

80 80 80 80

Dye Removal (%)

Salt Removal (%)


Salt Removal (%)
Dye Removal (%)

70 70 70 70

60 60 60 60

50 50 50 50

40 40 40 40
R-dye (1.0%) R-dye (5.0%) R-dye (1.0%) R-dye (5.0%)
30 R-salt (1.0%) R-salt (5.0%) 30 30 R-salt (1.0%) R-salt (5.0%) 30
R-dye (2.5%) R-dye (10%) 20 R-dye (2.5%) R-dye (10%) 20
20 20
R-salt (2.5%) R-salt (10%) R-salt (2.5%) R-salt (10%)
10 10 10 10

0 0 0 0
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
Time (min) Time (min)
140
280
c d
130
270

120 260

110 250
Flux (L/m2·h)

Flux (L/m2·h)
100 240

230
90
220
80
210
1% 1%
70
2.5% 200 2.5%
60 5% 5%
190
10% 10%
50 180
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
Time (min) Time (min)

Fig. 13. Rejection rates for dye and salt and fluxes of the PES MMMs: (a), (c) M-3 and (b), (d) M-5.

Table 3
Dye rejection rates of M-3 and M-5 after different run times in wastewater samples of varying salinity.
Salt concentration/% 30 min run 180 min run 350 min run

M-3 M-5 M-3 M-5 M-3 M-5

1 97.17 % 96.03 % 89.44 % 77.68 % 87.82 % 74.9 %


2.5 93.72 % 95.7 % 84.06 % 73.73 % 80.75 % 68.26 %
5 91.68 % 89.25 % 75.63 % 63.13 % 70.51 % 59.13 %
10 83.82 % 76.82 % 58.12 % 49.02 % 56.68 % 42.02 %

with ZIF-8@Ti3C2Tx composites were constructed for removing dye negatively correlated with dye retention. In particular, the M-3 mem­
pollutants from high-salinity wastewater. Adding Ti3C2Tx into the brane exhibited high flux (140 L⋅m− 2⋅h− 1) and up to 98 % dye rejection
microemulsion used for preparing MMMs enhanced the distribution of in highly saline wastewater, but its rejection of salt ions was <5 %. In
ZIF-8 in PES polymeric matrix via the formation of ZIF-8@Ti3C2Tx addition, the dye rejection recovery rate of M-3 after each washing
composites. Due to the introduction of hydrophilic components in the exceeded 93 % over 5 fouling/backwashing cycles. These results
microemulsion (including BmimOTF, VeimBF4, Ti3C2Tx, and ZIF-8 strongly suggest that MMMs incorporated with ZIF-8@Ti3C2Tx com­
nanoparticles), the pore structures and surface polarity of the MMMs posites have promising application prospects for removing trace dye
were significantly improved, which in turn increased their permeation pollutants from high-salinity wastewater.
and anti-fouling performance. The M-5 membrane (3 % microemulsion
content) showed the highest flux and anti-fouling performance, at 280 CRediT authorship contribution statement
L⋅m− 2⋅h− 1 and 96 %, respectively. However, increasing the micro­
emulsion content above 3 % was found to decrease the flux and anti- Yu-yu Yao: Conceptualization, Methodology, Writing – original
fouling performance of the resulting MMMs, as this leads to agglomer­ draft. Ting Wang: Data curation, Writing – original draft. Li-guang Wu:
ation of the ZIF-8@Ti3C2Tx. Conceptualization, Supervision, Writing – review & editing. Hua-li
The interaction between the well-distributed ZIF-8@Ti3C2Tx com­ Chen: Data curation, Writing – review & editing.
posites in the PES polymeric matrix and dye molecules was found to
strengthen the dye retention from high-salinity wastewater. Further, the
pore size of the MMMs was positively correlated with the flux but

10
Y.-y. Yao et al. Desalination 543 (2022) 116116

Fig. 14. Flux and rejection rates for separating dye from high-salinity water over repeated fouling/backwashing cycles for (a) M-3 and (b) neat PES.

National Natural Science Foundation of China (Contracts 22078291 and


Table 4
42077178) and the Natural Science Foundation of Zhejiang Province
Comparison of the separation performance of membranes modified with
(Contracts LY20B060001). We would like to thank MogoEdit (https
different materials.
://www.mogoedit.com) for its English editing during the preparation
Fillers Flux/ Concentration Rejection of this manuscript.
L⋅m− 2⋅h− 1
Dye/ Salt/ Dye/ Salt/
1
mg⋅L− g⋅L− 1 % % Appendix A. Supplementary data
ES/GO-1hr [41] 100 10 (EB) 60 99 5
COF-LZU1 [17] 80 50 (DB- 2 99 6 Supplementary data to this article can be found online at https://doi.
38)
org/10.1016/j.desal.2022.116116.
Fe/GO-TAx [42] 55 50 (CR) 1 99 10
ZIF-8 [13] 55 10 (RhB) 100 99 5
TA-PEI [43] 40.6 100 (CR) 1 99 6 References
SMA-PEI [44] 23.6 100 (CR) 1 99 2.5
HFM [45] 10.83 50 (CR) 1 99 8 [1] N. Shamsuddin, D.B. Das, V.M. Starov, Filtration of natural organic matter using
SiO2 [46] 89 100 (CR) 1 90 7 ultrafiltration membranes for drinking water purposes: circular cross-flow
PVA-GA/Cu(OH)2 37 100 (CR) 1 98 13 compared with stirred dead end flow, Chem. Eng. J. 276 (2015) 331–339, https://
[47] doi.org/10.1016/j.cej.2015.04.075.
ZIF-8@Ti3C2Tx (This 122 10 (RhB) 100 98 7 [2] V. Jabbari, J.M. Veleta, M. Zarei-Chaleshtori, J. Gardea-Torresdey, D. Villagrán,
Green synthesis of magnetic MOF@GO and MOF@CNT hybrid nanocomposites
work)
with high adsorption capacity towards organic pollutants, Chem. Eng. J. 304
(2016) 774–783, https://doi.org/10.1016/j.cej.2016.06.034.
[3] J.G. Zhang, Z.W. Xu, W. Mai, C.Y. Min, B.M. Zhou, M.J. Shan, Y.L. Li, C.Y. Yang,
Z. Wang, X.M. Qian, Improved hydrophilicity, permeability, antifouling and
Table 5 mechanical performance of PVDF composite ultrafiltration membranes tailored by
Flux and rejection rates for separating different dyes from high-salinity oxidized low dimensional carbon nanomaterials, J. Mater. Chem. A 1 (2013)
wastewater. 3101–3111, https://doi.org/10.1039/c2ta01415g.
[4] S. Zinadini, A.A. Zinatizadeh, M. Rahimi, V. Vatanpour, H. Zangeneh, Preparation
Dye Concentration/ Flux/ Dye Salt of a novel antifouling mixed matrix PES membrane by embedding grapheme oxide
mg⋅L− 1 L⋅m− 2⋅h− 1
rejection/% rejection/% nanoplates, J. Membr. Sci. 453 (2014) 292–301, https://doi.org/10.1016/j.
Congo Red 10 122.81 99 5 memsci.2013.10.070.
[5] S.A. Kiran, Y.L. Thuyavan, G. Arthanareeswaran, T. Matsuura, A.F. Ismail, Impact
Methylene 10 129.82 99 3
of graphene oxide embedded polyethersulfone membranes for the effective
Blue
treatment of distillery effluent, Chem. Eng. J. 286 (2016) 528–537, https://doi.
Methyl 10 136.84 95 3
org/10.1016/j.cej.2015.10.091.
Orange [6] J. Chen, T. Shi, L.B. Duan, Z.K. Sun, E.J. Anthony, Microemulsion-derived,
Rhodamine B 10 133.33 96 3 nanostructured CaO/CuO composites with controllable particle grain size to
enhance cyclic CO2 capture performance for combined Ca/Cu looping process,
Chem. Eng. J. 393 (2020) 1–9, https://doi.org/10.1016/j.cej.2020.124716.
Declaration of competing interest [7] X.J. Huang, Y.B. Chen, X.H. Feng, X.Y. Hu, Y.F. Zhang, L. Liu, Incorporation of oleic
acid-modified Ag@ZnO core-shell nanoparticles into thin film composite
membranes for enhanced antifouling and antibacterial properties, J. Membr. Sci.
The authors declare that they have no known competing financial 602 (2020) 1–13, https://doi.org/10.1016/j.memsci.2020.117956.
interests or personal relationships that could have appeared to influence [8] F. Khoerunnisa, W. Rahmah, B. Seng Ooi, E. Dwihermiati, N. Nashrah, S. Fatimah,
the work reported in this paper. Y.G. Ko, E.-P. Ng, Chitosan/PEG/MWCNT/iodine composite membrane with
enhanced antibacterial properties for dye wastewater treatment, J. Environ. Chem.
Eng. 8 (2020) 1–10, https://doi.org/10.1016/j.jece.2020.103686.
Data availability [9] M. Safarpour, J. Barzin, A. Khataee, Z. Kordkatooli, Two-stage phase separation of
cellulose acetate membranes modified with plasma-treated natural zeolite:
response surface modeling, Polym. Adv. Technol. 30 (2019) 889–901, https://doi.
Data will be made available on request. org/10.1002/pat.4522.
[10] F. Xiang, A.M. Marti, D.P. Hopkinson, Layer-by-layer assembled polymer/MOF
Acknowledgment membrane for H2/CO2 separation, J. Membr. Sci. 556 (2018) 146–1533, https://
doi.org/10.1016/j.memsci.2018.03.081.

The authors gratefully acknowledge the financial support from the

11
Y.-y. Yao et al. Desalination 543 (2022) 116116

[11] T.A. Makhetha, R.M. Moutloali, Incorporation of a novel Ag-Cu@ZIF-8@GO [30] M. Samari, S. Zinadini, A.A. Zinatizadeh, M. Jafarzadeh, F. Gholami, Designing of a
nanocomposite into polyethersulfone membrane for fouling and bacterial novel polyethersulfone (PES) ultrafiltration (UF) membrane with thermal stability
resistance, J. Membr. Sci. 618 (2021), 118733, https://doi.org/10.1016/j. and high fouling resistance using melamine-modified zirconium-based metal-
memsci.2020.118733. organic framework (UiO-66-NH2/MOF), Sep. Purif. Technol. 251 (2020), 117010,
[12] T.A. Makhetha, R.M. Moutloali, Antifouling properties of Cu(tpa)@GO/PES https://doi.org/10.1016/j.seppur.2020.117010.
composite membranes and selective dye rejection, J. Membr. Sci. 554 (2018) [31] F. Gholami, S. Zinadini, A.A. Zinatizadeh, A.R. Abbasi, TMU-5 metal-organic
195–210, https://doi.org/10.1016/j.memsci.2018.03.003. frameworks (MOFs) as a novel nanofiller for flux increment and fouling mitigation
[13] T.X. Wang, S.R. Chen, T. Wang, L.G. Wu, Y.X. Wang, PES mixed-matrix in PES ultrafiltration membrane, Sep. Purif. Technol. 194 (2018) 272–280, https://
ultrafiltration membranes incorporating ZIF-8 and poly(ionic liquid) by doi.org/10.1016/j.seppur.2017.11.054.
microemulsion synthetic with flux and antifouling properties, Appl. Surf. Sci. 576 [32] L. Zhang, X.X. Shi, M.H. Zhao, Z. Yin, J.P. Zhang, S.B. Wang, W. Du, J. Xiang, P.
(2021) 1–11, https://doi.org/10.1016/j.apsusc.2021.151815. G. Cheng, N. Tang, Construction of precisely controllable and stable interface
[14] W. Zheng, X. Hao, L. Zhao, W. Sun, Controllable preparation of nanoscale metal- bonding Au-TiO2/PVDF composited membrane for biofouling-resistant properties,
organic frameworks by ionic liquid microemulsions, Ind. Eng. Chem. Res. 56 Surf. Interfaces 24 (2021), 101152, https://doi.org/10.1016/j.
(2017) 5899–5905, https://doi.org/10.1021/acs.iecr.7b00694. surfin.2021.101152.
[15] J.F. Li, Z.L. Xu, H. Yang, Microporous polyethersulfone membranes prepared under [33] J.L. Lv, G.Q. Zhang, H.M. Zhang, C.Q. Zhao, F.L. Yang, Improvement of antifouling
the combined precipitation conditions with non-solvent additives, Polym. Adv. performances for modified PVDF ultrafiltration membrane with hydrophilic
Technol. 19 (2008) 251–257, https://doi.org/10.1002/pat.982. cellulose nanocrystal, Appl. Surf. Sci. 440 (2018) 1091–1100, https://doi.org/
[16] C.S. Feng, B.L. Shi, G.M. Li, Y.L. Wu, Preparation and properties of microporous 10.1016/j.apsusc.2018.01.256.
membrane from poly (vinylidene fluoride-co-tetrafluoroethylene) (F2.4) for [34] C.K. Ding, X.W. Qin, Y.Y. Tian, B.W. Cheng, PES membrane surface modification
membrane distillation, J. Membr. Sci. 237 (2004) 15–24, https://doi.org/10.1016/ via layer-by-layer self-assembly of GO@TiO2 for improved photocatalytic
j.memsci.2004.02.007. performance, J. Membr. Sci. 659 (2022), 120789, https://doi.org/10.1016/j.
[17] Y.Y. Su, X. Yan, Y. Chen, X.J. Guo, W.Z. Lang, Facile fabrication of COF-LZU1/PES memsci.2022.120789.
composite membrane via interfacial polymerization on microfiltration substrate for [35] S. Gulc, Z.A. Rehanbd, S.A. Khanab, K. Akhtarg, M.A. Khanc, M.I. Khanc, M.
dye/salt separation, J. Membr. Sci. 618 (2021), 118706, https://doi.org/10.1016/ I. Rashidef, A.M. Asiriab, S.B. Khanab, Antibacterial PES-CA-Ag2O nanocomposite
j.memsci.2020.118706. supported cu nanoparticles membrane toward ultrafiltration, BSA rejection and
[18] G. Chen, W.Q. Weng, L. Cong, P. Tan, Y. Jiang, Q. Zhang, X.Q. Liu, L.B. Sun, reduction of nitrophenol, J. Mol. Liq. 230 (2017) 616–624, https://doi.org/
Decorating MXene with tiny ZIF-8 nanoparticles: an effective approach to construct 10.1016/j.molliq.2016.12.093.
composites for water pollutant removal, Chin. J Chem. Eng. 42 (2021) 42–48, [36] S. Amiria, A. Asgharia, V. Atanpourb, M. Rajabia, Fabrication of chitosan-
https://doi.org/10.1016/j.cjche.2021.06.004. aminopropylsilane graphene oxide nanocomposite hydrogel embedded PES
[19] X.X. Wang, F.F. You, L.S. Wu, R. Ji, X.Y. Wen, B.X. Fan, G.X. Tong, D.B. Chen, W. membrane for improved filtration performance and lead separation, Sci. Direct,
H. Wu, Enhanced heat conductance and microwave absorption of 2D laminated J. Environ. Manag. 294 (2021), 112918, https://doi.org/10.1016/j.
Ti3C2Tx MXene microflakes via steering surface, defects, and interlayer spacing, jenvman.2021.112918.
J. Alloys Compd. 918 (2022), 165740, https://doi.org/10.1016/j. [37] M. Jiang, K.F. Y, J.J. Deng, J.Y. Lin, W.Y. Ye, S.F. Zhao, B.V.D. Bruggen,
jallcom.2022.165740. Conventional ultrafiltration as effective strategy for dye/salt fractionation in textile
[20] C. Gu, H.W. Mao, W.Q. Tao, Z. Zhou, X.J. Wang, P. Tan, S. Cheng, W. Huang, L. wastewater treatment, Environ. Sci. Technol. 52 (2018) 10698–10708, https://doi.
B. Sun, X.Q. Liu, J.Q. Liu, Facile synthesis of Ti3C2Tx-poly(vinylpyrrolidone) org/10.1021/acs.est.8b02984.
nanocomposites for nonvolatile memory devices with low switching voltage, ACS [38] J.B. Jin, X.L. Du, J. Yu, S.H. Qin, M. He, K.Z. Zhang, G.J. Chen, High performance
Appl. Mater. Interfaces 11 (2019) 38061–38067, https://doi.org/10.1021/ nanofiltration membrane based on SMA-PEI cross-linked coating for dye/salt
acsami.9b13711. separation, J. Membr. Sci. 611 (2020), 118307, https://doi.org/10.1016/j.
[21] F. Lessan, R. Foudazi, Effect of [EMIM][BF4] ionic liquid on the properties of memsci.2020.118307.
ultrafiltration membranes, Polymer 210 (2020), 122977, https://doi.org/10.1016/ [39] Y.F. Mi, G. Xu, Y.S. Guo, B. Wu, Q.F. An, Development of antifouling nanofiltration
j.polymer.2020.122977. membrane with zwitterionic functionalized monomer for efficient dye/salt
[22] L.Å. Naslund, N. Persson, XPS spectra curve fittings of Ti3C2Tx based on first selective separation, J. Membr. Sci. 601 (2020), 117795, https://doi.org/10.1016/
principles thinking, Appl. Surf. Sci. 593 (2022), 153442, https://doi.org/10.1016/ j.memsci.2019.117795.
j.apsusc.2022. [40] H.Q. Chang, H. Liang, F.S. Qu, B.C. Liu, H.R. Yu, X. Du, G.B. Li, S.A. Snyder,
[23] S.H. Zhang, Y. Liu, D. Li, Q. Wang, F. Ran, Water-soluble MOF nanoparticles Hydraulic backwashing for low-pressure membranes in drinking water treatment: a
modified polyethersulfone membrane for improving flux and molecular retention, review, J. Membr. Sci. 540 (2017) 362–380, https://doi.org/10.1016/j.
Appl. Surf. Sci. 505 (2020), 144553, https://doi.org/10.1016/j. memsci.2017.06.007.
apsusc.2019.144553. [41] N. Mehrabi, H.Q. Lin, N. Aich, Deep eutectic solvent functionalized graphene oxide
[24] L.G. Wu, L.L. Huang, Y. Yao, Z.H. Liu, T. Wang, X.Y. Yang, C.Y. Dong, Fabrication nanofiltration membranes with superior water permeance and dye desalination
of polyvinylidene fluoride blending membrane coupling with microemulsion performance, Chem. Eng. J. 412 (2021), 128577, https://doi.org/10.1016/j.
polymerization and their anti-fouling performance, Polymer 203 (2020) 1–9, cej.2021.128577.
https://doi.org/10.1016/j.polymer.2020.122767. [42] D.L. Xu, H. Liang, X.W. Zhu, L. Yang, X.S. Luo, Y.Q. Guo, Y.T. Liu, L.M. Bai, G.B. Li,
[25] M.F. Sun, T. Wang, L.G. Wu, Y.X. Wang, Enhancing the permeation and antifouling X.B. Tang, Metal-polyphenol dual crosslinked graphene oxide membrane for
performance of PVDF hybridmembranes by incorporating Co-Fe hydroxide desalination of textile wastewater, Desalination 487 (2020), 114503, https://doi.
nanoparticles in reverse microemulsion, J. Environ. Chem. Eng. 9 (2021), 106556, org/10.1016/j.desal.2020.114503.
https://doi.org/10.1016/j.jece.2021.106556. [43] Q. Li, Z.P. Liao, X.F. Fang, D.P. Wang, J. Xie, X.Y. Sun, L.J. Wang, J.S. Li, Tannic
[26] S.M. Maroofi, N.M. Mahmoodi, Zeolitic imidazolate framework acid-polyethyleneimine crosslinked loose nanofiltration membrane for dye/salt
polyvinylpyrrolidone-polyethersulfone composites membranes: from synthesis to mixture separation, J. Membr. Sci. 584 (2019) 324–332.
the detailed pollutant removal from wastewater using cross flow system, Colloids [44] J.B. Jin, X. Du, J. Yu, S.H. Qin, M. He, K.Z. Zhang, G.J. Chen, High performance
Surf. A Physicochem. Eng. Asp. 572 (2019) 211–220, https://doi.org/10.1016/j. nanofiltration membrane based on SMA-PEI cross-linked coating for dye/salt
colsurfa.2019.03.093. separation, J. Membr. Sci. 611 (2020), 118307, https://doi.org/10.1016/j.
[27] L.L.S. Silva, W. Abdelraheem, M.N. Nadagouda, A.M. Rocco, D.D. Dionysiou, F. memsci.2020.118307.
V. Fonseca, C.P. Borges, Novel microwave-driven synthesis of hydrophilic [45] H.Q. Liu, Y.B. Chen, K. Zhang, C.F. Wang, X.Y. Hu, B.W. Cheng, Y.F. Zhang, Poly
polyvinylidene fluoride/polyacrylic acid (PVDF/PAA) membranes and decoration (vinylidene fluoride) hollow fiber membrane for high-efficiency separation of dyes-
with nano zero-valent-iron (nZVI) for water treatment applications, J. Membr. Sci. salts, J. Membr. Sci. 578 (2019) 43–52, https://doi.org/10.1016/j.
620 (2021), 118817, https://doi.org/10.1016/j.memsci.2020.118817. memsci.2019.02.029.
[28] A. Karimi, A. Khataee, V. Vatanpour, M. Safarpour, High-flux PVDF mixed matrix [46] Z.T. Liu, R.R. Qiang, L.G. Lin, X.S. Deng, X. Yang, K.Y. Zhao, J. Yang, X.Y. Li, W.
membranes embedded with size-controlled ZIF-8 nanoparticles, Sep. Purif. S. Ma, M.N. Xu, Thermally modified polyimide/SiO2 nanofiltration membrane
Technol. 229 (2019), 115838, https://doi.org/10.1016/j.seppur.2019.115838. with high permeance and selectivity for efficient dye/salt separation, J. Membr.
[29] A. Karimi, A. Khataee, V. Vatanpour, M. Safarpour, The effect of different solvents Sci. 658 (2022), 120747, https://doi.org/10.1016/j.memsci.2022.120747.
on the morphology and performance of the ZIF-8 modified PVDF ultrafiltration [47] Y.D. Chen, R.Z. Sun, W.T. Yan, M.Y. Wu, Y. Zhou, C.J. Gao, Antibacterial polyvinyl
membranes, Sep. Purif. Technol. 253 (2020), 117548, https://doi.org/10.1016/j. alcohol nanofiltration membrane incorporated with Cu(OH)2 nanowires for dye/
seppur.2020.117548. salt wastewater treatment, Sci. Total Environ. 817 (2022), 152897, https://doi.
org/10.1016/j.scitotenv.2021.152897.

12

You might also like