Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Chemical Engineering & Processing: Process Intensification 158 (2020) 108191

Contents lists available at ScienceDirect

Chemical Engineering and Processing - Process


Intensification
journal homepage: www.elsevier.com/locate/cep

Antifouling and photocatalytic properties of 2-D Zn/Al layered double


hydroxide tailored low-pressure membranes
Yuvaraj Mutharasi a, Noel Jacob Kaleekkal a, *, Thanigaivelan Arumugham b, Fawzi Banat b,
MSR Sridhar Kapavarapu a
a
Membrane Separation Group, Department of Chemical Engineering, National Institute of Technology Calicut (NITC), Kozhikkode, 673601, Kerala, India
b
Department of Chemical Engineering, Khalifa University, 127788, Abu Dhabi, United Arab Emirates

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

Keywords: In this study, Zinc-Aluminium layered double hydroxide (LDH) was synthesized to fabricate membranes with
Zinc-Aluminium layered double hydroxide superior hydrophilic and photocatalytic characteristics. The prepared LDH was extensively characterized and
Mixed matrix membranes employed for the preparation of polyetherimide mixed matrix ultrafiltration (UF) membranes, which displayed >
Thin-film nanocomposite
96 % BSA rejection and a flux recovery ratio of 88 % even after 3-cycles. M-1 (UF, 1 wt. %) displayed a 94 %
Photocatal
degradation of organic dye (methylene blue) in 120 min when irradiated with UV-light. Further, nanofiltration
Ytic activity
Hydrophilicity (NF) membranes were prepared by incorporating Zn/Al LDH into the polyamide (PA) layer during the in-situ
interfacial polymerization to prepare thin-film nanocomposite membranes. MT-2 (NF, 0.4 wt.% LDH in PA
layer) displayed improved hydrophilicity (lower water contact angle), > 94.5 % rejection of cationic dye
(Methylene blue), > 91 % rejection of inorganic salt (CaSO4) with an 80 % flux recovery after long-term
filtration. The LDH holds an auspicious potential to tune the properties of UF and NF membranes to be used
in the treatment of industrial effluents.

1. Introduction high cost of recovery of the nanoparticles used in the adsorption or


photocatalysis severely limit their largescale application [6]. For the
In India, groundwater provides 80 % of drinking water and two- removal of these contaminants, membrane-based separations are being
thirds of its irrigation needs. A sharp rise in the number of industries – widely accepted over other competing technologies due to their attri­
textile, petrochemical, food, pharmaceuticals, paper, and pulp, etc. butes such as higher efficiency, ability to be scaled up and integrated
contributes to contamination of the groundwater resources [1]. Textile into existing systems, cost-effectiveness, robustness, no/low require­
effluents contain organic pollutants, toxic dyes, inorganic compounds, ment of chemicals, etc. [7]. Highly selective, low-pressure membrane
and heavy metals, which, when discharged to the environment, can processes such as ultrafiltration (UF) and nanofiltration (NF) play a vital
directly harm the ecosystem and diminish the quality of human life [1, role in the recovery and reuse of water to alleviate the current water
3]. Methylene blue is a cationic dye that is widely used in several ap­ stress [8].
plications (medicinal, dying, etc.) and, when discharged into waste­ Nanomaterials have been widely used in the preparation of ultrafil­
water, can pose a significant burden to the environment [2]. Calcium tration membranes, and this class of membranes is referred to as mixed
sulphate used in primary water treatment In addition to these, the matrix membranes/nanocomposite membranes. These mixed matrix
presence of proteins in the effluents can pose a threat if not treated membranes are employed for the separation of humic substances, pro­
before discharge [5]. teins, oil-water emulsion, and other macromolecular substances [9,10].
Conventional technologies such as adsorption, coagulation, bio- Recent reports on mixed matrix membranes show that nanomaterials
mediated process, advanced oxidation processes (photocatalysis) have viz. multi-walled carbon nanotubes [11], graphene oxide [12], percar­
been explored for the removal of these contaminants. However, (i) boxylic acid-functionalized SiO2 bound Fe3O4 nanoparticle [13], MIL-53
limited efficiency, (ii) inability to degrade or remove all classes of the [14], ZnO [15] have been incorporated into the polymer matrix to
contaminants, (iii) creation of sludge/secondary pollutants, and (iv) improve the antifouling property, provide adsorption sites, aid in

* Corresponding author.
E-mail addresses: noeljacob89@gmail.com, noel@nitc.ac.in (N.J. Kaleekkal).

https://doi.org/10.1016/j.cep.2020.108191
Received 6 June 2020; Received in revised form 10 September 2020; Accepted 15 October 2020
Available online 19 October 2020
0255-2701/© 2020 Elsevier B.V. All rights reserved.
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

catalytic degradation, and improve membrane longevity, etc. by incorporating LDH into the polyetherimide (PEI) matrix, which, to
Nanofiltration (NF) membranes possessing smaller pore sizes (0.5 the best of our knowledge, has not been reported earlier. The MMMs
− 2 nm) have also received widespread attention owing to their ability to were characterized using FTIR, TGA, AFM, SEM, Tensile strength, Water
remove low molecular weight organic compounds (dyes, etc.) as well as contact angle, etc. The protein rejection and antifouling ability of the
dissolved divalent/trivalent salts and having a lower energy require­ MMMs were evaluated. Further, the photocatalytic degradation of
ment [16]. However, adsorption of dyes can cause potential fouling due methylene blue (MB) dye using the MMMs of the membranes was
to hydrophobic interactions and electrostatic interactions [17]. New explored using a photocatalytic reactor, and the leaching of the LDH
generation NF membranes are fabricated by an in-situ interfacial poly­ from the membrane matrix was analyzed. Thin-film nanocomposite
merization forming an ultrathin polyamide (PA) layer on the base membranes (TFNs) were prepared by the incorporation of the same LDH
UF/MF membrane. However, producing membranes with superior into the polyamide (PA) layer during the interfacial polymerization
water permeability with a high rejection remains a challenging task between an amine and an acyl chloride monomer on the PEI base
[18]. In this case, the introduction of nanomaterials into the PA active membrane and have not been investigated earlier. These nanofiltration
layer is a promising direction of research to prepare thin-film nano­ membranes were evaluated for pure water permeability, organic solute
composite membranes (TFNs) with superior separation ability. The rejection (Methylene blue dye rejection), divalent ion removal (Ca+2),
addition of functional materials possessing a variety of characteristics and flux recovery after long-term filtration, etc.
can tune the membrane surface property by introducing- hydrophilicity,
antifouling ability, surface charge, surface roughness, antibacterial 2. Experimental details
property, photocatalytic ability, etc. [19]. Nanomaterials such as CeO2,
TiO2, Ag, GO, CNTs have been explored for their potential to impart 2.1. Materials used
antifouling property to the membrane surface, and these membranes
have been explored for the removal of various contaminants from Zinc nitrate hexahydrate (reagent grade, 98 %), Aluminium nitrate
aqueous streams [20]. nonahydrate (EMPLURA grade), Sodium hydroxide (ACS reagent, ≥97.0
Two-dimensional (2D) nanomaterials obtained popularity with the %, pellets), and Sodium nitrate (99.9 % Assay) were purchased from
application and versatility of graphene/graphene oxide sheets. Some of Merck Life Sciences Ltd for the synthesis of the hydrotalcite-like LDH.
the other classes of 2D materials actively being explored are 2D metal- PEI (melt index9 g/10 min (337 ◦ C/6.6 kg)) manufactured by Sigma
organic frameworks, MXenes, layered metal carbide/nitride, layered Aldrich was used and Solvent N-Methyl Pyrrolidone (A.R. grade) man­
dichalcogenides, and layered double hydroxides [21]. Layered double ufactured by Sisco Research laboratory, India were used in the mem­
hydroxides are a class of two dimensional (2D) nanomaterials commonly brane preparation. m-Phenylenediamine (MPD flakes, 99 %) 1,3,5-
referred to as hydrotalcite-like compounds with flexibility in composi­ Benzenetricarbonyl trichloride (TMC 98 %) from Sigma-Aldrich, USA,
tion, low manufacturing cost, and ease of preparation. These are anionic n-Hexane (EMPLURA) from Merck was used for the interfacial poly­
layered materials composing of brucite like 2D sheets with alternating merization reaction to prepare the active PA layer. Methylene Blue
divalent and trivalent cations possessing an exchangeable anion with (373.9 g/mol), Bovine Serum Albumin (BSA molecular biology grade),
water molecules in the interlayer spaces [22]. These LDH have atomic Calcium Suphate (A.R grade, dried) were procured from Himedia Lab­
layer thickness, which (1) the cation layers can be tuned easily, (ii) can oratories, India. Deionized water (R = 18.2 MΩ cm) obtained from a
host a variety of anions by intercalation, and (iii) has a distinctive Milli-Q system was employed throughout this study.
memory effect [23]. There is an increase in the number of reports of LDH
being employed for the removal of contaminants from aqueous systems: 2.2. Synthesis of Zinc-Aluminium layered double hydroxide
for adsorption for phosphate recovery from urine [24], removal of tar­
trazine dye from aqueous solutions [25], nanofillers for preparing TFN A facile one-pot synthesis of Zn(II) and Al (III) layered double hy­
membranes [26], antibacterial and photocatalytic microfiltration droxide was prepared by co-precipitation followed by aging. Briefly,
membranes [27]. aqueous solutions of Al(N03)3.9H2O (Merck, emplura) and Zn
In this study, a layered anionic clay Zinc -Aluminium LDH possessing (NO3)2.6H2O (Merck, reagent grade, 98 %) were mixed by the dropwise
dual functionality – hydrophilicity and photocatalytic ability is explored addition of one solution into the other using a magnetic stirrer in an inert
as nanofillers to tailor the performance of UF and NF polyetherimide (N2) atmosphere. The concentrations of the salts were calculated, such
(PEI) membranes. PEI is a hydrophobic polymer having excellent film- that the molar ratio of Zn2+/Al3+ was in the ratio 2:1. To this solution,
forming properties, possessing high mechanical strength/ chemical 50 mL of NaNO3 was added, followed by the dropwise addition of 1 M
resistance, and is available commercially at competitive prices [28]. NaOH. The dropwise addition of NaOH resulted in the precipitation of
Moreover, PEI is a UV resistant polymer which is attributed to the high the metal oxides at a pH ~10. The slurry was then transferred into a
dissociation energy of the C– – C (145 Kcal/mol) and C–– O (84 Kcal/mol) Teflon lined autoclave and aged at 110 ⁰C for 18 h. The precipitate ob­
bonds present in the polymer backbone [29]. tained was filtered, washed with water and acetone and dried in a
The Zn/Al LDH possesses − OH groups at its surface, which impart vacuum oven for 48 h. The dried sample was ground and stored dry for
hydrophilicity, and the positive charge can be traced to the isomorphous further use [33].
substitution of the divalent cations by the trivalent ones [30]. The
photocatalytic ability of the LDH can be associated with the − OH− 2.3. Fabrication of Low-pressure membranes (UF and NF)
groups of the brucite layer, which combine with the valence band holes
yielding hydroxyl radicals. Moreover, the presence of a suitable anion 2.3.1. Mixed matrix ultrafiltration (UF) membranes (MMMs)
(here NO−3 ) in the interlamellar space quenches the rate of charge The ultrafiltration membranes were prepared by the phase inversion
recombination as a result of interaction between the photo-generated technique, in particular, the non-solvent induced phase separation
charge carrier and the anion [31]. Introducing a photocatalytic nano­ method (NIPS) using DI water as the non-solvent. The determined
material can not only improve the membrane antifouling ability but also quantity of LDH (Table 1) was dispersed in the solvent, n-methyl pyr­
enhances the catalytic degradation of the adsorbed organic foulants, rolidone (SRL, India) using an ultrasonic batch (40 kHz) for 30 min to
improving the lifetime of the membrane. This could impart a ensure complete dispersion of the nanofiller. To this, the polymer, PEI
self-cleaning property to the membrane, thereby reducing the time and (Sigma Aldrich, melt index9 g/10 min) was added and stirred using an
solvents needed for backwash and membrane regeneration [32]. overhead mechanical stirrer for 8 h until a homogenous dope solution
The synthesized LDH is characterized extensively by XRD, FTIR, was obtained. The degassed solution was cast onto a clean glass sub­
TGA, SEM, TEM, etc. Mixed matrix membranes (MMMs) were fabricated strate using a thin-film applicator (Elcometer 4340) while maintaining a

2
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

Table 1
Composition and Physico-chemical properties of the prepared membranes.
Porosity Mechanical Properties
Membrane Weight % of LDH used in (%) Average surface roughness Viscosity of the Casting Solution
Codes preparation Tensile Strength % Elongation at Ra (nm) (mPa.s)
(MPa) Break

Mixed matrix Ultrafiltration Membranes (16 wt.% Polyetherimide)


M0 0 53.6 ± 3.0 3.55 (±0.01) 6.7 (±2.00) 46 ± 12 862.4
M-1 1 68.4 ± 2.4 3.61 (±0.05) 12.4 (±1.80) 131 ± 29 1024.6
M-2 1.5 62.5 ± 1.8 3.68 (±0.07) 15.8 (±2.15) 86 ± 29 1261.5

Thin-film nanocomposite membranes (16 wt.% Polyetherimide +1 wt. % PVP K-90)


MT 0 28.2 ± 1.4 6.21 (±1.05) 8.2 (±0.55) 72 ± 4 3340.1a
MT-1 0.2 36.4 ± 2.5 6.18 (±0.05) 11.6 (±2.00) 81.5 ± 3 3340.1a
MT-2 0.4 35.3 ± 1.2 5.89 (±1.00) 9.8 (±1.20) 88 ± 5 3340.1a
a
Viscosity of the base membrane.

thickness of 150 ± 5 μm. The glass substrate with the polymer solution 2.6. Membrane performance evaluation
was immediately immersed in DI water (non-solvent) to obtain the
polymeric membrane. 2.6.1. Performance evaluation of the mixed matrix ultrafiltration
membranes
2.4. Thin-film nanocomposite (TFN) nanofiltration (NF) membranes The filtration experiments were carried out in a self-designed cross-
flow filtration set up using a test cell (Sterlitech Corporation, USA)
The pristine PEI UF membranes fabricated (as discussed previously) having an effective membrane area of 42 cm2. The MMMs were com­
was used as the base membrane for the preparation of the thin-film pacted at 5 bar for 2 h before carrying out the filtrations at an operating
nanocomposite membranes. The polyamide layer was prepared by the pressure of 3 bar.
in-situ interfacial polymerization (IP) of m-Phenylenediamine (MPD, The pure water flux and the solution flux of the membranes were
Sigma, 99 %) with 1,3,5-Benzenetricarbonyl trichloride (TMC, Sigma, evaluated at 25 ⁰C using:
98 %) [34]. The base membrane was affixed onto an acrylic frame with
V
the top layer facing upwards. Briefly, 100 mL of 1 wt.% aqueous MPD J (LMH) = (1)
A ×t
solution with LDH was poured onto the membrane surface and held for
120 s to ensure maximum penetration into the pores, after which the where V is the volume of permeate in Litres, A is the effective area of the
excess solution was removed using filter paper and a rubber roller. A membrane used (m2), and t is the time taken to collect permeate in
0.1 wt. % solution of TMC dissolved in n-hexane was then poured over hours.
the membrane surface and held for 60 s to aid in the polycondensation The protein BSA (1000 mg/l), MB dye (20 mg/l) were used as feed
reaction between the monomers. The excess, unreacted monomers were solutions to investigate the rejection performance of the MMMs (pH ~
removed by rinsing using DI water, and membranes were dried at 60 ⁰C 7.5). The concentrations of BSA and MB dye were A UV–vis spectro­
for 20 min for curing to stop the reaction altogether. The membranes photometer (PerkinElmer-Lambda 650) at a wavelength of 280 nm and
(Table 1) were stored in DI water until further use. 665 nm, respectively, using the following equation
[ ]
2.5. Characterizations Rejection (%) = 1 −
Cpermeate
× 100 (2)
Cfeed
The X-ray diffraction patterns of the LDH were obtained from The antifouling properties of the membranes were determined, as
2Θ = 3− 90◦ using Rigaku Miniflex 600; Japan operated at 40 kV and discussed in previous reports. Initially, the pure water flux of the
15 mA with CuKαas, the radiation source. Attenuated total reflectance- membrane (Jw1) was determined using DI water as feed at an operating
Fourier transform infrared spectroscopy (ATR-FTIR/FTIR) spectra in pressure of 3 bar for 60 min using eq (1). An aqueous solution of BSA was
the range of 4000–400 cm− 1 were recorded using an Agilent Technol­ used as the feed solution, and the filtration was carried out under the
ogies Cary 630, USAFTIR Spectrometer. The cross-sectional and top same pressure for 60 min. The membrane was backwashed and then
surface morphology of the membranes were investigated using Scanning flushed with DI water for 10 min. The PWF of the cleaned membrane
Electron Microscopy (Hitachi SU6600, Japan) and atomic force micro­ (Jw2) was determined similarly. The experiment was repeated for a total
scopy (Park Systems XE-100, Korea). Transmission electron microscopy of 3 cycles, and the flux recovery ratio (FRR) was determined using the
(TEM) equipped with an Energy Dispersive X-Ray Spectroscopy (EDS) given equation
was done using the FEI-Tecnai G2 20 Twin instrument. The thermo­
grams were realized using a thermogravimetric analyzer (Hitachi FRR(%) =
JW2
× 100 (3)
STA7200), and the samples were scanned from room temperature to JW1
800 ◦ C at a rate of 10 ◦ C/min. The surface area and porosity of the LDH These experiments were repeated for 3 such runs, and the FRR values
were calculated from the BET equation using Belsorp Max (Japan) in­ for each run is determined.
strument. The tensile strength was obtained using the Universal Testing
Machine (UTM), Shimadzu AG-X plus, Japan at 10 kN, and a draw rate 2.6.2. Photocatalytic degradation of methylene blue using the mixed matrix
of 0.5 mm/min (n = 3). The water contact angle was determined by the membranes
sessile drop method (GBX technologies, Germany) using a volume of Initially, the adsorption efficiency of the pristine PEI and MMMs
5 μL, and measurement was carried out on at least 5 different locations were evaluated by placing a 10 × 15 cm membrane sample into a
on the membrane sample. 400 mL solution of Methylene Blue dye (MB, 20 mg/L). The adsorption
experiment was carried out under dark conditions for 120 min, and a
2 mL aliquot was collected every 10 min for analysis. The concentration
of the dye was calculated using a UV–vis Spectrophotometer (664 nm).

3
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

To carry out the photocatalytic degradation study, a photocatalytic concentrations of MB dye and CaSO4 was analyzed using a UV–vis
reactor (Lelesil innovative systems, India) was employed, which uses a spectrophotometer (λmax =665 nm) and conductivity meter (Systronics-
125 W medium pressure mercury lamp as a UV-source placed in the 304, India) respectively, using eq (2).
inner-well. A 400 mL of dye solution was placed in the outer well The flux recovery after long-term use was carried using CaSO4 feed
equipped with a cooling jacket. Fresh membrane samples with an solution for a total of 12 h with a 10 min backwash after 6 h.
exposed area of 150 cm2 were introduced into the outer well containing
the dye solution. The solution with the MMMs was stirred using a 3. Results and discussion
magnetic stirrer and kept in the dark for 30 min to allow it to reach
adsorption equilibrium before the UV lamp was turned on. 2 mL aliquot 3.1. Characterization of the zinc-aluminium layered double hydroxide
were collected every 10 min until the end of the experiment (120 min)
using a silicon tube connected to a disposable syringe. The concentration The XRD diffraction pattern of the Zn/Al LDH synthesized by a one-
of the MB dye in the solution was measured as before. The membranes pot co-precipitation technique is shown in Fig. 1a. The sharp diffraction
were tested for 2 such runs. peaks were indexed to a hexagonal lattice with rhombohedral 3R sym­
metry as expected for these ionic class of solids (JCPDS 48–1023). The
2.6.3. Performance evaluation of the thin-film nanocomposite (TFN) basal peaks at 11.54◦ , 23.39◦ can be indexed to the (003) and (006)
nanofiltration membranes which provides the evidence of parallel orientation of the nanosheets.
The same experimental set-up was used to carry out the nano­ The non-basal peaks (012), (015), (018), (110), (113) are observed at
filtration experiments at an operating pressure of 6 bar, and these TFN 36.23◦ , 39.17◦ , 47.55◦ , 56.63◦ and 67.97◦ , respectively. The (003) re­
membranes were pre-compacted at 10 bar for 2 h before the start of the flections can be attributed to the anion present in the interlayer of the
experiments. The pure water flux and the solution flux of the membranes LDH incorporated [33]. The first peak of the doublet at a 2Θ of around
were evaluated at 25 ⁰C using eq (1). 60◦ is because of the alignment of the (110) planes, which matches with
MB dye (20 mg/l) and CaSO4 (1200 mg/l) were used as feed solu­ the closest interatomic distance in the brucite-like layers. The average
tions to investigate the rejection performance of the membranes. The size of the particle is calculated using Bragg’s equation and found to be

Fig. 1. X-ray diffraction spectra (a), FT-IR spectra (b), TGA curve (c), SEM image (d) and TEM images (e,f) of the Zinc-Aluminium layered double hydroxide.

4
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

60 nm.
Fig. 1b displays the FTIR spectrum of the prepared LDH and confirms
the presence of the absorption bands inherent in hydrotalcite com­
pounds. The broad distinct band at 3400 cm− 1 can be assigned to the
O–H stretching vibration of the interlamellar H2O moieties. The weak
band at 1680 cm− 1 can be assigned to the distorting vibration of the
interlayer water [35]. The bands recorded at 460, 550, and 790 cm− 1
can be attributed to the Al-O condensed groups, the Zn/Al− OH trans­
lation, and the Al− OH deformation, respectively. The well-defined ab­
sorption peak at 1385 cm− 1 confirms the stretching vibration of the
anion, and the weak band at 765 cm− 1 is another stretching mode of the
nitrate anion.
The total mass loss of 22.5 % was calculated from the thermogram
(Fig. 1c); when the LDH was heated to 800 ◦ C. The weight loss at <
100 ◦ C can be attributed to the removal of the physisorbed water mol­
ecules. Two distinct losses of removal of bound water and interlamellar
water molecules can be observed at 180 ◦ C and 200 ◦ C [35]. The
noticeable mass loss from 300− 450 ◦ C could be explained by the
dehydroxylation of the metal oxide layers and degradation of the
anionic (NO−3 1) interlayers.
SEM and TEM images confirm the morphology of the prepared LDH.
SEM images display agglomeration of flake-like crystallites in Fig. 1d.
The expected 2-D sheets having a hexagonal structure with well-defined
edges can be confirmed from the TEM images (Fig. 1e,f).

3.2. Characterization of the UF mixed matrix membranes

3.2.1. Chemical structure and surface wettability


The chemical structure of the prepared MMMs can be deduced
through the FT-IR spectra displayed in Fig. 2a.The FTIR spectra of the
pristine PEI membrane (M0) displayed absorption bands at 1780 cm− 1
and 1718 cm− 1, which corresponds to the asymmetric and symmetric
– O of the imide groups. The bands at 1356 cm− 1 and
stretching of the C–
745 cm− 1 can be attributed to the C– – N stretching and bending vibra­

tions. The C–O–C vibrations of the PEI are ascribed to the band at
1234 cm-1 [36]. The membrane M-2 displayed all the significant peaks of
corresponding to PEI polymer and additionally exhibits an absorption
peak at 1380 cm− 1, which can be accounted for the anti-symmetric
stretching mode of the nitrate anion present in Zn-Al-NO3 layers. The
incorporation of LDH can be confirmed by the increase in the stretching
vibration of the − OH bonds at 3500 cm− 1. An increase in the intensity of
the absorption bands at 790, 460 and 550 cm-1 correspond to the Al-O
condensed groups of the LDH. As the percentage of LDH on the mem­
brane surface is quite small, no major changes in the absorption spectra
were recorded.
The droplets of DI water are dispensed onto the clean membrane
surface using a Hamilton syringe, and the average contact angle of 5
such drops was determined, and their average plotted in Fig. 2b. The
surface of the M0 membrane displayed the lowest wetting by the water Fig. 2. The surface chemical composition (a), the surface wettability (b) and
drop (high water contact angle), and this indicated a more hydrophobic the TGA (c) of the LDH incorporated polyetherimide mixed matrix membranes.
surface. Incorporation of LDH into the membrane matrix provided
greater − OH sites for the water molecules to attach via H-bonding. This and bound water molecules.M-2 displays a further weight loss between
plays an essential role in the spreading of the water droplet, as 300− 450 ◦ C could be due to the loss of anions from the interlamellar
confirmed by the lower water contact angle measured. The porosity structures of the incorporated LDH nanomaterials. A temperature
(Table 1) and chemical structure account for the decrease in WCA from greater than 450 ◦ C leads to the scissoring of polymer chains, ring fusion,
87.2◦ (±1.1) for M-0 to 60.4◦ (±2.0) for M-2. The water molecules and finally, carbonization of the polymer membrane. The thermal sta­
surrounding the membranes form H-bonding with the − OH groups of bility of the MMMs is satisfactory for them to be employed in a photo­
the LDH present in the membrane matrix. This can prevent the catalytic reactor for the degradation of the MB dye [39].
adsorption of hydrophobic foulants onto the membrane surface [37], There is only an incremental increase in the tensile strength of the
and similar results are reported when hydrophilic halloysite nano­ membranes incorporated with LDH nanoparticles (Table 1). This could
tube/Fe3O4 were incorporated into hydrophobic PVDF membranes [38]. be explained by the following factors: LDH bearing a copious number of
functional groups (OH− , NO−3 , etc.) penetrates between the polymer
3.2.2. Thermal and mechanical stability chains and produces cross-linking, which could be responsible for the
The TGA of the mixed matrix membranes is displayed in Fig. 2c. The increase in % elongation. The crystalline nanoparticles embedded
pristine and the MMMs did not show appreciable weight loss up to within the matrix improve the overall crystallinity of the mixed matrix
250 ◦ C; the small decrease in weight could be due to the loss of unbound

5
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

membranes, which could be responsible for the small increase in the finger-like pores with the formation of a considerable quantity of
tensile strength [40]. There are reports of a decrease in mechanical interconnected pores. This could explain the greater pure water flux
properties of the membrane with the incorporation of nano-fillers due to described in the later section. The higher viscosity of the casting solution
lower interaction between polymer and nanoparticles or inhomoge­ (Table 1) leads to a sponge-like morphology for M-2, which is distinct in
neous dispersion density [41]. However, in this case, this phenomenon the SEM images.
was not observed due to the low loading ratio of the LDH. This indicates The AFM was used in the non-contact mode to investigate the to­
that LDH nanomaterial was relatively well dispersed within the mem­ pology of other membrane surfaces. It was observed that the average
brane matrix with an excellent affinity to the PEI matrix. surface roughness was most significant for the 1 wt% membrane (M-1)
followed by M-2 and least was for the pristine PEI membrane. This could
3.2.3. Analysis of the membrane architecture be explained as follows: the presence of LDH in the dope solution (up to
Fig. 3 shows the top surface and cross-sectional morphologies of the 1 wt.%) could increase the rate of solvent and non-solvent exchange,
prepared MMMs. All the MMMs possess a typical asymmetric structure thereby giving rise to a larger number of peaks and valleys. Above this
with a thin top layer with surface pores and a porous sublayer. The concentration (at 1.5 wt.%), the nanomaterial increased the viscosity of
incorporation of the hydrophilic nanomaterial, LDH, affects the kinetics the casting solution, thereby suppressing the rate of solvent and non-
and thermodynamics of the phase inversion process, thereby altering the solvent exchange, leading to a surface with lower roughness (Table 1).
membrane vitrification pathway in the ternary phase diagram [42]. The increase in surface roughness can also explain improved water
Both of these factors play opposing roles in the membrane formation permeability as it increases the number of surface pores and provides a
process. The pristine PEI membranes exhibited finger-like pores in the greater contact area (hydrophilic sites), which enhances the water pas­
porous sublayer and with a few surface pore openings. The top surface sage [43].
image of the membrane M-1 reveals the presence of a more significant The improvement in membrane porosity, which is evident in the SEM
number of pore openings on the denser top surface due to the increase in and AFM images are quantified using the gravimetric technique, and the
hydrophilic content in the polymer solution. The substructure remained results are given in Table 1.

Fig. 3. The cross-sectional SEM image, the top surface SEM image and the topography from AFM imaging for M0 (a, a’, a’’); M-1 (b, b’, b’’) and M-2 (c, c’, c’’)
respectively.

6
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

3.3. Pure water flux and rejection studies of the MMMs %), which rapidly declined in successive cycles (Fig. 5). Membrane M-1
possessed the most excellent antifouling property evidenced by a 92 %
The pure water flux of the prepared mixed matrix membranes FRR, which showed only a small decline even after two cycles. The
(MMM) is calculated and plotted in Fig.4. The water flux was obtained membrane hydrophilicity, optimum loading, and distribution of the
using pre-compacted membranes of higher pressure to obtain consistent LDH contributed to its highest FRR. This indicates a low level of irre­
results. The pure water flux of M1 and M2 is greater than the pristine PEI versible fouling even after three cycles, which was reported with the
membrane. This could be explained as follows: the incorporation of LDH incorporation of hydrophilic filler GO in the polyethersulfone/sulfo­
containing large no. of − OH groups improve the porosity of the mem­ nated polysulfone ultrafiltration membranes [50]. During the filtration
branes due to its preferential migration towards the non-solvent (water) studies, Zn or Al was not detected (ICP-OES, Perkin Elmer Optima 5300
[44]. The addition of LDH improves the hydrophilicity of the mem­ DV) in the permeate side, indicating that the leaching of the LDH from
branes, as evident from the water contact angle results [45]. Apart from the MMMs was negligible.
this, the membrane surface roughness (rough surface of hydrophilic
membranes) provide a greater area of contact to the water molecules
and thus enhance the permeation rate. When the LDH loading was 3.5. Photocatalytic Degradation of Methylene Blue Dye using the mixed
increased to 1.5 wt%, the pure water flux decreased slightly to 77 LMH; matrix membranes
this contrast in behavior could be associated with the agglomeration or
aggregation of LDH [46], which causes blocking and channeling of the The photocatalytic degradation of the MB dye is a complex multi-
pores or the pores of smaller diameter and spongier substructure as step process involving several intermediate steps, which finally yields
confirmed by the SEM images Fig. 3 non-toxic carbon-di-oxide and water as by-products. Here the double-
The rejection of the BSA (66.5 kDa) was found to be > 85 % for the well enclosed chamber is equipped with 125-Watt high-pressure mer­
UF membranes indicating that the pore sizes had not significantly cury as a UV source to evaluate the photocatalytic degradation effi­
changed with the addition of the Zn/Al LDH. The slight increase in ciency of the prepared membranes. The MMMs containing the LDH
rejection from 85.3 % (M0) to 96 % (M-1) can be explained as follows:
above the isoelectric point, BSA is negatively charged and can be
effectively hindered due to the combination of steric effect and elec­
trostatic effects (adsorption) as observed in earlier reports [47].
The rejection of the methylene blue dye, having a molecular weight
373.9 Da, was low, as expected when filtered using ultrafiltration
membranes. There is a slight increase in rejection due to the increased
repulsion of the cationic dye with the less-negatively charged mixed
matrix membranes. However, no significant results are obtained as total
rejection is less than 20 %.

3.4. Antifouling ability and the Flux recovery ratio of the MMMs

The membranes having high rejection of BSA cause an increase in


BSA concentration at the membrane surface, which can attach to the
membrane causing an irreversible loss in permeate quantity as well as
quality [48].
The decrease in flux when the BSA solution was used can be because
of the formation of a gel layer as BSA molecules interact with the
membrane surface, thereby increasing the membrane resistance. The
presence of LDH on the surface and inside the pores of the mixed matrix
membranes improves the membrane hydrophilicity, which lowers the
rate of adsorption of the hydrophobic BSA molecules, thereby
decreasing the membrane fouling [49].
The pristine PEI membranes exhibited the least flux recovery (68.5

Fig. 4. The pure water flux and the rejection performance of the mixed ma­ Fig. 5. (a) Performance of the MMM for a single cycle and (b) Flux Recovery
trix membranes. Ratio (%) for 3 cycles.

7
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

when irradiated with low wavelength UV light (higher intensity than the photocatalytically degrade an organic compound (Methylene Blue Dye)
band gap), can engender high-energy states of electrons and hole pairs. under UV –light in a batch mode. M-1 holds great potential to be further
These electron/hole pairs function as powerful oxidizing/reduction explored for UF applications and has been compared with recent liter­
species or can generate new active species that can aid in the degrada­ ature reports (Table 2) [54–58].
tion of organic compounds in aqueous solutions [51].
The photolysis (photo-degradation) of the MB dye without the 3.6. Characterisation of thin-film nanocomposite NF membranes
MMMs is evaluated prior to the photocatalytic experiments using the
MMM. Furthermore, the removal of MB dye by adsorption is also 3.6.1. SEM of TFN membranes
determined to confirm the photocatalytic property of the LDHs. The cross-sectional SEM images (Fig. 7a,b,c) show a highly porous
Initially, the feed solution (MB- 20 mg/mL, pH 7.5) was fed into the PEI sublayer with well-developed finger-like pores having a high degree
photocatalytic reactor, and the chamber was closed, and the UV source of interconnectivity. The incorporation of 0.4 wt%. LDH is seen to lower
turned on. In this case, no LDH or MMMs were placed in the reactor. It the thickness (300 nm) of the PA layer compared to the MT and MT-1
was noticed that there was no significant decrease (< 3 %) in dye con­ membranes (420 nm).
centration at the end of 60 min, which implies that direct photo- A characteristic ridge-and-valley like topography is evident in the
degradation of MB did not occur at these conditions (A cooling water top-surface images of TFC as well as the nanocomposite modified
jacket did not allow the temperature in the reactor to rise more than 4 membranes (Fig. 7a’,b’,c’). The LDH increases the formation of the
⁰C). The irradiation of dye molecules by the photon bombardment nodular structures in the PA layer, which could be due to the adsorption
provided some excitation energy, and subsequent dispersion of this of water molecules during the interfacial polymerization. The hydration
energy did not decolourize the MB solution. of these water molecules of LDH introduced in the aqueous amine so­
The absorption kinetics of M-1 (Fig.6a), carried out under dark lution could cause an augmentation in its size, forming the nodule-like
conditions, confirms a 25 % reduction in MB concentration. It is inter­ structure [59]. The dispersed LDH could be observed as white clusters
esting to note that the total adsorption was completed in the initial in the top layer of the MT-2, which was absent in the MT-1. The presence
20 min, which could be due to the affinity for the cationic dye to the of an optimum concentration of LDH can improve the morphology of the
hydroxyl and nitrate anions present in the LDH. Moreover, there was no membrane surface to ensure that there is no pore blocking or chan­
further change in concentration, even up to 100 min [52]. neling, which is present in MT-2. Even in higher loading, only a few
After initial equilibration in dark-condition, the MB feed solution nanomaterial clusters can be observed as most of the particles are
with the MMMs immersed was placed in the photocatalytic chamber, embedded in the PA layer [60]. The EDX (Fig. 7d) mapping confirms the
and the UV source turned on. The solution was continuously stirred to incorporation of Zn/Al LDH on the membrane surface.
provide in-situ aeration and to maintain a homogenous mixture
throughout the experiment. M-1 and M-2 membranes exhibited a dye 3.6.2. AFM of TFN membranes
removal of 94 % and 92 %, respectively, during the initial run [53]. This The membrane surface roughness investigated using the AFM is
could be explained by the larger surface area available due to the well displayed in Fig. 8, and the surface roughness increased as follows:
dispersed LDH in the matrix, which increased the membrane surface MT < MT-1< MT-2. The higher content of the LDH in the active PA layer
roughness. M-2 exhibited a decline in dye degradation during the second was responsible for the increase in the membrane surface roughness. The
run, which could be due to the larger number of LDH particles that could LDH can absorb water molecules and swell during the IP, leading to the
irreversibly bind the dye or intermediate degradation products (Fig. 6b). formation of nodules, as explained earlier. The increased roughness of
The degradation of the dye is aided by the holes (conduction band) the membranes can also improve membrane permeability as it offers
and hydroxyl radicals generated; the mechanism can be postulated as greater contact area and more hydrophilic pathways through the PA
follows. layer, which is also seen to impart antifouling property to the TFNs [61].
LDH + h ν (<400 nm) → (e−CB)+ (h+
VB)
3.6.3. Surface wettability of the TFN membranes
(h+
VB) + R → intermediates →CO2+ H2O The water contact angles (Fig. 9) measured were found to decrease
→ with the increase in LDH loading, from ~ 64.6◦ for the TFC membrane to
H2O + (h+
VB) OH⋅ + H
+
~ 52.6◦ for MT-1. An optimum concentration of hydrophilic nano­
OH⋅+ R → intermediates → CO2+ H2O. materials in the surface is thought to be responsible for the decline in the
WCA to its due hydrogen bonding with water molecules [62]. This,
The ultrafiltration studies and photocatalytic experiments concluded coupled with the increase in porous channels due to the LDH, can attract
that M-1 (1 wt.% loading of Zn/Al LDH) displayed > 96 % BSA rejection water molecules to the membrane surface and improve its water
with a flux recovery ratio of 88 % even after 3-cycles and was able to permeability. MT-2 displays a slightly higher contact angle due to the

Fig. 6. The photocatalytic degradation a), reusability of the MMMs b) for degradation of MB dye (initial concentration 20 mg/l).

8
Y. Mutharasi et al.
Table 2
A comparison with recent state-of-the-art literature.
Mixed Matrix Ultrafiltration Membranes

Sl. Polymer (wt.%) Filler (%) PWF (LMH) @ operating pressure BSA (rejection %) FRR% Other Specific Properties Ref.
No.
1 Cellulose Acetate (10.5 PEG 316 LMH 91.9 91 – [54]
%) 5 wt% (3 bar)
2 PES 1 wt% PVP and 1 wt% PA-6 175 LMH 96 55 – [55]
(18 %) (4 bar)
3 PES (16.5 %) +PEG (8 Iron-tanin-framework 319.4 LMH 95.9 88.29 – [56]
%) (ITF) (1 bar)
0.3 wt%
4 PVDF (13.3 %) Cellulose Nanocrystal 40 LMH 88.2 80 – [57]
(0.7 wt%) (1 bar)
5 PES (15 %) Bentonite (2%) 53.6 LMH 73.7 84.6 % – [58]
(2 bar)
6 PEI (16 %) 1 % Zn-Al-LDH 89 LMH 96.5 92 Able to Photocatalytically degrade organic compounds Current
in UV-Light Work
Nanofiltration membranes
9

Sl. Base Polymer (wt.%) Monomers used in IP Nanofillers (%) PWF @ pressure Solutes removed & % Other Specific Properties Ref
No. removal
1 PolyImide (PI) MPD and TMC Carbon Quantum Dots (CQD) 42.1 LMH 93.6 % Na2SO4 Antifouling performance [68]
(6 bar)
− 6
2 Polysulfone MPD and TMC – 17.02 LMH 88.75 % rejection of Permeability coefficient of NaCl 0.98 × 10 m/s [69]

Chemical Engineering and Processing - Process Intensification 158 (2020) 108191


(15 wt.%) (14 bar) hydroquinone
3 Polysulfone PIP and TMC ZIF-8 55.02 LMH 95 % [66]
(commercial) (6 bar) Na2SO4
4 Polysulfone PIP and TMC Graphitic carbon nitride (g-C3N4) 82 LMH 94.5 % [67]
(commercial) 16.4 μg cm− 2 g (4 bar) Na2SO4
2
and Halloysite nanotubes 19.7 μg cm−
(HNTs)
5 PAN (phase inversion) – Citric Acid and TiO2 58− 136 LMH 66− 86 % Calcium Suphate – [70]
(10 bar)
6 PEI MPD and TMC Zn-Al-LDH (0.1 wt %) MT-2 Pure water flux Antifouling property after long-term filtration (12 h) Current
18 LMH 91 % Zn/Al LDH possesses photocatalytic abilities. Work
(6 bar) Calcium Suphate
95 %
Methylene Blue Dye
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

Fig. 7. Cross-Section and top Surface SEM micrographs MT (a, a’), MT-1 (b, b’), MT-2 (c, c’) and the EDX mapping of the MT-2 membrane (d).

Fig. 8. Surface roughness using atomic force microscopy of MT (a,), MT-1 (b,), MT-2 (c).

LMH in MT-0 to 26.4 LMH in MT-1. This increase is due to the incor­
poration of the hydrophilic LDH that improves the membrane porosity
and hydrophilicity. The interruption in the PA layer due to an optimum
concentration of LDH can also explain the result. A slight decrease in
PWF is observed in MT-2, which can be attributed to the agglomeration
of particles on the top surface, which causes partial or complete blocking
of pores evident in the SEM images. However, it still exhibits greater flux
than pristine TFC. The TFN membranes possess better antifouling
properties, and an increase in flux by almost 2.5 times was recorded.
The permeate flux of the MB feed solution was lower than the PWF as
expected (Fig. 10a), and the rejection of the dye is due to mainly due to
size exclusion. The presence of the LDH tunes the PA layer forming
smaller pore openings. MT-2 membranes display the most outstanding
rejection (94.5 %) (Fig. 10b) could be due to the pore-blocking, as dis­
cussed in the previous section. The LDH can act to prevent the adsorp­
tion of the cationic dye on the membrane surface due to its hydrophilic
character, and this does not allow a build-up of the dye molecule layer,
ensuring that the membranes can be used for a long time [64].
Fig. 9. The water contact angle indicating the surface wettability of the TFNs.
3.6.5. Rejection Studies of calcium sulphate using TFN membranes
increase in LDH concentration, which could lead to a greater degree of Here, CaSO4 is chosen as a model divalent salt to investigate the
clusters being formed. The agglomeration of the nanofillers can lead to rejection performance of the nanofiltration membranes, and it could be
the formation of H-bonding within itself, thus decreasing the − OH noted that the membranes exhibited lower salt rejection compared to
groups available, as observed in earlier reports [63]. MB dye as expected due to the difference in molecular weights. An in­
crease in rejection of CaSO4 salt was observed for the TFN membranes
3.6.4. Pure water flux and MB rejection of the TFN membranes (Fig.10 b), which could be explained as follows- (i) the hydrated radii of
The pure water flux of the membranes (Fig.10a) increased from 10.5 the ions was greater than the pathway formed in the PA layer due to the

10
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

Fig. 10. Pure water flux, solution permeate flux (a) and rejection performance (b) of the TFNs.

incorporation of LDH and (ii) the incorporation of LDH is seen to in­


crease the surface charge (less negative) [30] which improves the salt
rejection due to Donnan exclusion effect. Membrane MT-2 has greater
flux and better rejection than the pristine TFC, which is an indication
that the trade-off criterion could be surpassed by the incorporation of
Zn-Al LDH [59,65].

3.6.6. Flux recovery ratio of the TFN membranes for long-term filtration
The improvement of flux recovery is due to several factors, all
explained in previous sections. Improvement in hydrophilicity and an
increase in surface roughness prevents the molecules from adhering or
depositing to the membrane surface. The flux recovery of MT-2 (Fig. 11)
is compared to the pristine TFC membrane. The membranes were
challenged with 1200 ppm CaSO4 solution for 6 h, followed by a mem­
brane cleaning for 10 min and was repeated two times (12 h). The MT-2
membrane exhibited a flux recovery ratio of 0.8 after two successive
runs (total 12 h). The high flux recovery is attributed to the improved
surface hydrophilicity and the hydration layer formed at the membrane
interface. Our prepared TFN membranes are compared with recent
studies and given in Table 2 [66–70]. The membrane performance is Fig. 11. Flux recovery after long term filtration with 1200 ppm CaSO2 solution.
comparable to other works and holds the potential to be employed as
TFN nanofiltration membranes for the removal of dye and divalent salts. Writing - review & editing. Thanigaivelan Arumugham: Formal
analysis, Resources, Writing - original draft. Fawzi Banat: Writing -
4. Conclusion review & editing. MSR Sridhar Kapavarapu: Data curation, Investi­
gation, Validation.
Ultrafiltration using PEI mixed matrix membranes incorporated with
Zn2+/Al3+ holds the potential to be employed for pre-treatment of Declaration of Competing Interest
wastewater before the nanofiltration process. The MMMs exhibited an
impressive photocatalytic ability in addition to the antifouling ability. The authors declare that they have no known competing financial
High rejection of BSA and an excellent flux recovery after 3-cyles run interests or personal relationships that could have appeared to influence
proves that the MMMs are favorable to be used in the pre-treatment step. the work reported in this paper.
The incorporation of LDH into the active layer of the TFN membranes
produced nanofiltration membranes that displayed greater surface Acknowledgments
porosity, enhanced surface wettability, and improved antifouling abil­
ity. MT-1 displayed a 2.5 times increase in pure water flux while The authors like to acknowledge the partial financial support by the
showing no deviation in thermal or mechanical stability. However, the National Institute of Technology Calicut, India for the student innova­
optimum membrane MT-2 displayed a flux greater than the TFC with an tive project grant titled “Development of layered double hydroxide
MB dye rejection of 94.5 % and 91 % salt rejection with a modest flux incorporated PEI membranes with enhanced photocatalytic and anti­
recovery after long-term filtration studies. The results indicate that the fouling ability” to Dr. Noel Jacob Kaleekkal.
dual functionality of the Zn/Al LDH and hence these materials hold
considerable potential to be employed as fillers in low-pressure mem­ References
branes (UF and NF), which have to be investigated for industrial efflu­
ents before they can be employed commercially. [1] Y. Lv, C. Zhang, A. He, S.-J. Yang, G.-P. Wu, S.B. Darling, Z.-K. Xu, Photocatalytic
nanofiltration membranes with self-cleaning property for wastewater treatment,
Adv. Funct. Mater. 27 (2017) 1700251, https://doi.org/10.1002/
CRediT authorship contribution statement adfm.201700251.
[2] M. Contreras, C.D. Grande-Tovar, W. Vallejo, C. Chaves-López, Bio-removal of
Yuvaraj Mutharasi: Investigation, Data curation, Validation, methylene blue from aqueous solution by galactomyces geotrichum KL20A, Water.
11 (2019) 282, https://doi.org/10.3390/w11020282.
Writing - original draft. Noel Jacob Kaleekkal: Conceptualization, [3] D.A. Yaseen, M. Scholz, Textile dye wastewater characteristics and constituents of
Funding acquisition, Project administration, Visualization, Supervision, synthetic effluents: a critical review, Int. J. Environ. Sci. Technol. 16 (2019)
1193–1226, https://doi.org/10.1007/s13762-018-2130-z.

11
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

[5] A.W. Mohammad, Y.H. Teow, W.L. Ang, Y.T. Chung, D.L. Oatley-Radcliffe, Ag3PO4/CuZnAl NLDH, Sep. Purif. Technol. 226 (2019) 218–231, https://doi.org/
N. Hilal, Nanofiltration membranes review: recent advances and future prospects, 10.1016/j.seppur.2019.05.104.
Desalination. 356 (2015) 226–254, https://doi.org/10.1016/j.desal.2014.10.043. [28] N.J. Kaleekkal, R. Radhakrishnan, V. Sunil, G. Kamalanathan, A. Sengupta,
[6] W. Raza, J. Lee, N. Raza, Y. Luo, K.-H. Kim, J. Yang, Removal of phenolic R. Wickramasinghe, Performance evaluation of novel nanostructured modified
compounds from industrial waste water based on membrane-based technologies, mesoporous silica/polyetherimide composite membranes for the treatment of oil/
J. Ind. Eng. Chem. 71 (2019) 1–18, https://doi.org/10.1016/J.JIEC.2018.11.024. water emulsion, Sep. Purif. Technol. 205 (2018) 32–47, https://doi.org/10.1016/
[7] M.C. Nayak, A.M. Isloor, Inamuddin, B. Prabhu, N.I. Norafiqah, A.M. Asiri, Novel J.SEPPUR.2018.05.007.
polyphenylsulfone (PPSU)/nano tin oxide (SnO2) mixed matrix ultrafiltration [29] N.S. Aseri, W.J. Lau, P.S. Goh, H. Hasbullah, N.H. Othman, A.F. Ismail, Preparation
hollow fiber membranes: fabrication, characterization and toxic dyes removal from and characterization of polylactic acid-modified polyvinylidene fluoride hollow
aqueous solutions, React. Funct. Polym. 139 (2019) 170–180, https://doi.org/ fiber membranes with enhanced water flux and antifouling resistance, J. Water
10.1016/j.reactfunctpolym.2019.02.015. Process Eng. 32 (2019), https://doi.org/10.1016/j.jwpe.2019.100912.
[8] R. Rodrigues, J.C. Mierzwa, C.D. Vecitis, Mixed matrix polysulfone/clay [30] M.H. Tajuddin, N. Yusof, N. Abdullah, M.N.Z. Abidin, W.N.W. Salleh, A.F. Ismail,
nanoparticles ultrafiltration membranes for water treatment, J. Water Process Eng. T. Matsuura, N.H.H. Hairom, N. Misdan, Incorporation of layered double
31 (2019) 100788, https://doi.org/10.1016/j.jwpe.2019.100788. hydroxide nanofillers in polyamide nanofiltration membrane for high performance
[9] D. Qadir, H. Mukhtar, L.K. Keong, Mixed matrix membranes for water purification of salts rejections, J. Taiwan Inst. Chem. Eng. 97 (2019) 1–11, https://doi.org/
applications, Sep. Purif. Rev. 46 (2017) 62–80, https://doi.org/10.1080/ 10.1016/j.jtice.2019.01.021.
15422119.2016.1196460. [31] L. Mohapatra, K. Parida, A review on the recent progress, challenges and
[10] D.S. Dlamini, J. Li, B.B. Mamba, Critical review of montmorillonite/polymer perspective of layered double hydroxides as promising photocatalysts, J. Mater.
mixed-matrix filtration membranes: possibilities and challenges, Appl. Clay Sci. Chem. A. 4 (2016) 10744–10766, https://doi.org/10.1039/C6TA01668E.
168 (2019) 21–30, https://doi.org/10.1016/j.clay.2018.10.016. [32] A.T. Kuvarega, N. Khumalo, D. Dlamini, B.B. Mamba, Polysulfone/N,Pd co-doped
[11] M. Chandrashekhar Nayak, A.M. Isloor, Inamuddin, B. Lakshmi, H.M. Marwani, TiO2 composite membranes for photocatalytic dye degradation, Sep. Purif.
I. Khan, Polyphenylsulfone/multiwalled carbon nanotubes mixed ultrafiltration Technol. 191 (2018) 122–133, https://doi.org/10.1016/J.SEPPUR.2017.07.064.
membranes: Fabrication, characterization and removal of heavy metals Pb2+, Hg2 [33] S. Iftekhar, M.E. Küçük, V. Srivastava, E. Repo, M. Sillanpää, Application of zinc-
+, and Cd2+ from aqueous solutions, Arab. J. Chem. 13 (2019) 4661–4672, aluminium layered double hydroxides for adsorptive removal of phosphate and
https://doi.org/10.1016/j.arabjc.2019.10.007. sulfate: equilibrium, kinetic and thermodynamic, Chemosphere. 209 (2018)
[12] M.S. Algamdi, I.H. Alsohaimi, J. Lawler, H.M. Ali, A.M. Aldawsari, H.M.A. Hassan, 470–479, https://doi.org/10.1016/J.CHEMOSPHERE.2018.06.115.
Fabrication of graphene oxide incorporated polyethersulfone hybrid ultrafiltration [34] S.H. Maruf, A.R. Greenberg, Y. Ding, Influence of substrate processing and
membranes for humic acid removal, Sep. Purif. Technol. 223 (2019) 17–23, interfacial polymerization conditions on the surface topography and permselective
https://doi.org/10.1016/j.seppur.2019.04.057. properties of surface-patterned thin-film composite membranes, J. Memb. Sci. 512
[13] V. Vatanpour, S. Shahsavarifar, S. Khorshidi, M. Masteri-Farahani, A novel (2016) 50–60, https://doi.org/10.1016/j.memsci.2016.04.003.
antifouling ultrafiltration membranes prepared from percarboxylic acid [35] Z. Karami, M. Jouyandeh, J.A. Ali, M.R. Ganjali, M. Aghazadeh, S.M.R. Paran,
functionalized SiO 2 bound Fe 3 O 4 nanoparticle (SCMNP-COOOH)/ G. Naderi, D. Puglia, M.R. Saeb, Epoxy/layered double hydroxide (LDH)
polyethersulfone nanocomposite for BSA separation and dye removal, J. Chem. nanocomposites: synthesis, characterization, and Excellent cure feature of nitrate
Technol. Biotechnol. 94 (2019) 1341–1353, https://doi.org/10.1002/jctb.5894. anion intercalated Zn-Al LDH, Prog. Org. Coatings. (2019), https://doi.org/
[14] Y. Ren, T. Li, W. Zhang, S. Wang, M. Shi, C. Shan, W. Zhang, X. Guan, L. Lv, 10.1016/j.porgcoat.2019.105218.
M. Hua, B. Pan, MIL-PVDF blend ultrafiltration membranes with ultrahigh MOF [36] N.J. Kaleekkal, A. Thanigaivelan, M. Durga, R. Girish, D. Rana, P. Soundararajan,
loading for simultaneous adsorption and catalytic oxidation of methylene blue, D. Mohan, Graphene oxide nanocomposite incorporated poly(ether imide) mixed
J. Hazard. Mater. 365 (2019) 312–321, https://doi.org/10.1016/j. matrix membranes for in vitro evaluation of its efficacy in blood purification
jhazmat.2018.11.013. applications, Ind. Eng. Chem. Res. 54 (2015) 7899–7913, https://doi.org/
[15] A.K.T. Taha, M. Amir, A. Demir Korkmaz, M.A. Al-Messiere, A. Baykal, S. Karakuş, 10.1021/acs.iecr.5b01655.
A. Kilislioglu, Development of novel Nano-ZnO enhanced polymeric membranes [37] M.S.S.A. Saraswathi, A. Nagendran, D. Rana, Tailored polymer nanocomposite
for water purification, J. Inorg. Organomet. Polym. Mater. 29 (2019) 979–988, membranes based on carbon, metal oxide and silicon nanomaterials: a review,
https://doi.org/10.1007/s10904-018-0988-3. J. Mater. Chem. A. 7 (2019) 8723–8745, https://doi.org/10.1039/C8TA11460A.
[16] Y. Kang, M. Obaid, J. Jang, I.S. Kim, Sulfonated graphene oxide incorporated thin [38] R. Zhang, X. Zhu, Y. Liu, Y. Cai, Q. Han, T. Zhang, Y. Li, The application of
film nanocomposite nanofiltration membrane to enhance permeation and halloysite Nanotubes/Fe3O4 composites nanoparticles in polyvinylidene fluoride
antifouling properties, Desalination. 470 (2019) 114125, https://doi.org/10.1016/ membranes for dye solution removal, J. Inorg. Organomet. Polym. Mater. 29
j.desal.2019.114125. (2019) 1625–1636, https://doi.org/10.1007/s10904-019-01125-z.
[17] J. Dasgupta, J. Sikder, S. Chakraborty, S. Curcio, E. Drioli, Remediation of textile [39] F.Z. Mahjoubi, A. Khalidi, M. Abdennouri, N. Barka, Zn–Al layered double
effluents by membrane based treatment techniques: a state of the art review, hydroxides intercalated with carbonate, nitrate, chloride and sulphate ions:
J. Environ. Manage. 147 (2015) 55–72, https://doi.org/10.1016/J. synthesis, characterisation and dye removal properties, J. Taibah Univ. Sci. 11
JENVMAN.2014.08.008. (2017) 90–100, https://doi.org/10.1016/J.JTUSCI.2015.10.007.
[18] A.M.A. Abdelsamad, M. Matthias, A.S.G. Khalil, M. Ulbricht, Nanofillers [40] Y. Zhao, N. Li, F. Yuan, H. Zhang, S. Xia, Preparation and characterization of
dissolution as a crucial challenge for the performance stability of thin-film hydrophilic and antifouling poly(ether sulfone) ultrafiltration membranes modified
nanocomposite desalination membranes, Sep. Purif. Technol. 228 (2019) 115767, with Zn-Al layered double hydroxides, J. Appl. Polym. Sci. 133 (2016), https://doi.
https://doi.org/10.1016/j.seppur.2019.115767. org/10.1002/app.43988.
[19] C.S. Ong, P.S. Goh, W.J. Lau, N. Misdan, A.F. Ismail, Nanomaterials for biofouling [41] S.V. Kononova, G.N. Gubanova, E.N. Korytkova, D.A. Sapegin, K. Setnickova,
and scaling mitigation of thin film composite membrane: a review, Desalination. R. Petrychkovych, P. Uchytil, Polymer nanocomposite membranes, Appl. Sci. Basel
393 (2016) 2–15, https://doi.org/10.1016/J.DESAL.2016.01.007. (Basel) 8 (2018) 1181, https://doi.org/10.3390/app8071181.
[20] C. Ursino, R. Castro-Muñoz, E. Drioli, L. Gzara, M.H. Albeirutty, A. Figoli, Progress [42] M.H.D.A. Farahani, V. Vatanpour, A comprehensive study on the performance and
of nanocomposite membranes for water treatment, Membranes (Basel). 8 (2018), antifouling enhancement of the PVDF mixed matrix membranes by embedding
https://doi.org/10.3390/membranes8020018. different nanoparticulates: clay, functionalized carbon nanotube, SiO2 and TiO2,
[21] M. Safarpour, S. Arefi-Oskoui, A. Khataee, A review on two-dimensional metal Sep. Purif. Technol. 197 (2018) 372–381, https://doi.org/10.1016/J.
oxide and metal hydroxide nanosheets for modification of polymeric membranes, SEPPUR.2018.01.031.
J. Ind. Eng. Chem. 82 (2020) 31–41, https://doi.org/10.1016/j.jiec.2019.11.002. [43] N. Abdullah, R.J. Gohari, N. Yusof, A.F. Ismail, J. Juhana, W.J. Lau, T. Matsuura,
[22] N. Basaki, A. Kakanejadifard, M. Shabanian, K. Faghihi, Synthesis and Polysulfone/hydrous ferric oxide ultrafiltration mixed matrix membrane:
characterization of a new photosensitive and electroactive polyamide/LDH preparation, characterization and its adsorptive removal of lead (II) from aqueous
nanocomposite containing azo groups, Polym. Bull. Berl. (Berl) (2019) 1–16, solution, Chem. Eng. J. 289 (2016) 28–37, https://doi.org/10.1016/J.
https://doi.org/10.1007/s00289-019-03048-8. CEJ.2015.12.081.
[23] P. Lu, Y. Liu, T. Zhou, Q. Wang, Y. Li, Recent advances in layered double [44] X.K. Lin, X. Feng, L. Chen, Y.P. Zhao, Characterization of temperature-sensitive
hydroxides (LDHs) as two-dimensional membrane materials for gas and liquid membranes prepared from poly(vinylidene fluoride)-graft-poly(N-
separations, J. Memb. Sci. 567 (2018) 89–103, https://doi.org/10.1016/j. isopropylacrylamide) copolymers obtained by atom transfer radical
memsci.2018.09.041. polymerization, Front. Mater. Sci. China 4 (2010) 345–352, https://doi.org/
[24] K. Dox, M. Everaert, R. Merckx, E. Smolders, Optimization of phosphate recovery 10.1007/s11706-010-0106-0.
from urine by layered double hydroxides, Sci. Total Environ. 682 (2019) 437–446, [45] Y. Zhao, N. Li, F. Yuan, H. Zhang, S. Xia, Preparation and characterization of
https://doi.org/10.1016/j.scitotenv.2019.05.181. hydrophilic and antifouling poly(ether sulfone) ultrafiltration membranes modified
[25] J. Li, S. Yuan, J. Zhu, B. Van der Bruggen, High-flux, antibacterial composite with Zn–Al layered double hydroxides, J. Appl. Polym. Sci. 133 (2016), https://
membranes via polydopamine-assisted PEI-TiO2/Ag modification for dye removal, doi.org/10.1002/app.43988.
Chem. Eng. J. 373 (2019) 275–284, https://doi.org/10.1016/j.cej.2019.05.048. [46] Y. Zhao, N. Li, B. Xu, B. Dong, S. Xia, Preparation and characterization of a novel
[26] G.S. Lai, W.J. Lau, P.S. Goh, M. Karaman, M. Gürsoy, A.F. Ismail, Development of hydrophilic poly(vinylidene fluoride) filtration membrane incorporated with Zn–Al
thin film nanocomposite membrane incorporated with plasma enhanced chemical layered double hydroxides, J. Ind. Eng. Chem. 39 (2016) 37–47, https://doi.org/
vapor deposition-modified hydrous manganese oxide for nanofiltration process, 10.1016/j.jiec.2016.05.006.
Compos. Part B Eng. 176 (2019) 107328, https://doi.org/10.1016/j. [47] X.T. Yuan, C.X. Xu, H.Z. Geng, Q. Ji, L. Wang, B. He, Y. Jiang, J. Kong, J. Li,
compositesb.2019.107328. Multifunctional PVDF/CNT/GO mixed matrix membranes for ultrafiltration and
[27] L. Ghalamchi, S. Aber, V. Vatanpour, M. Kian, Development of an antibacterial and fouling detection, J. Hazard. Mater. 384 (2020) 120978, https://doi.org/10.1016/
visible photocatalytic nanocomposite microfiltration membrane incorporated by j.jhazmat.2019.120978.

12
Y. Mutharasi et al. Chemical Engineering and Processing - Process Intensification 158 (2020) 108191

[48] E. Mahmoudi, L.Y. Ng, W.L. Ang, Y.T. Chung, R. Rohani, A.W. Mohammad, of salts rejections, J. Taiwan Inst. Chem. Eng. 97 (2019) 1–11, https://doi.org/
Enhancing morphology and separation performance of polyamide 6,6 membranes 10.1016/j.jtice.2019.01.021.
by minimal incorporation of silver decorated graphene oxide nanoparticles, Sci. [60] H. Dong, L. Wu, L. Zhang, H. Chen, C. Gao, Clay nanosheets as charged filler
Rep. 9 (2019) 1–16, https://doi.org/10.1038/s41598-018-38060-x. materials for high-performance and fouling-resistant thin film nanocomposite
[49] A. Karimi, A. Khataee, V. Vatanpour, M. Safarpour, High-flux PVDF mixed matrix membranes, J. Memb. Sci. 494 (2015) 92–103, https://doi.org/10.1016/j.
membranes embedded with size-controlled ZIF-8 nanoparticles, Sep. Purif. memsci.2015.07.049.
Technol. 229 (2019) 115838, https://doi.org/10.1016/j.seppur.2019.115838. [61] S. Ayyaru, Y.H. Ahn, Application of sulfonic acid group functionalized graphene
[50] M. Hu, Z. Cui, J. Li, L. Zhang, Y. Mo, D.S. Dlamini, H. Wang, B. He, J. Li, oxide to improve hydrophilicity, permeability, and antifouling of PVDF
H. Matsuyama, Ultra-low graphene oxide loading for water permeability, nanocomposite ultrafiltration membranes, J. Memb. Sci. 525 (2017) 210–219,
antifouling and antibacterial improvement of polyethersulfone/sulfonated https://doi.org/10.1016/j.memsci.2016.10.048.
polysulfone ultrafiltration membranes, J. Colloid Interface Sci. 552 (2019) [62] C.H. Koo, W.J. Lau, G.S. Lai, S.O. Lai, H.S. Thiam, A.F. Ismail, Thin-film
319–331, https://doi.org/10.1016/j.jcis.2019.05.065. nanocomposite nanofiltration membranes incorporated with graphene oxide for
[51] Y.-H. Chiu, T.-F.M. Chang, C.-Y. Chen, M. Sone, Y.-J. Hsu, Mechanistic insights into phosphorus removal, Chem. Eng. Technol. 41 (2018) 319–326, https://doi.org/
photodegradation of organic dyes using heterostructure photocatalysts, Catalysts. 9 10.1002/ceat.201700357.
(2019) 430, https://doi.org/10.3390/catal9050430. [63] J. Xue, Z. Jiao, R. Bi, R. Zhang, X. You, F. Wang, L. Zhou, Y. Su, Z. Jiang, Chlorine-
[52] K. Abderrazek, F.S. Najoua, E. Srasra, Synthesis and characterization of [Zn-Al] resistant polyester thin film composite nanofiltration membranes prepared with B-
LDH: Study of the effect of calcination on the photocatalytic activity, Appl. Clay cyclodextrin, J. Memb. Sci. 584 (2019) 282–289, https://doi.org/10.1016/j.
Sci. 119 (2016) 229–235, https://doi.org/10.1016/j.clay.2015.10.014. memsci.2019.04.077.
[53] H. Dzinun, M.H.D. Othman, A.F. Ismail, M.H. Puteh, M.A. Rahman, J. Jaafar, [64] L. Bai, Y. Liu, A. Ding, N. Ren, G. Li, H. Liang, Fabrication and characterization of
Photocatalytic degradation of nonylphenol by immobilized TiO2 in dual layer thin-film composite (TFC) nanofiltration membranes incorporated with cellulose
hollow fibre membranes, Chem. Eng. J. 269 (2015) 255–261, https://doi.org/ nanocrystals (CNCs) for enhanced desalination performance and dye removal,
10.1016/j.cej.2015.01.114. Chem. Eng. J. 358 (2019) 1519–1528, https://doi.org/10.1016/j.cej.2018.10.147.
[54] K.A. Gebru, C. Das, Effects of solubility parameter differences among PEG, PVP and [65] Y. Mo, A. Tiraferri, N.Y. Yip, A. Adout, X. Huang, M. Elimelech, Improved
CA on the preparation of ultrafiltration membranes: impacts of solvents and antifouling properties of polyamide nanofiltration membranes by reducing the
additives on morphology, permeability and fouling performances, Chinese J. density of surface carboxyl groups, Environ. Sci. Technol. 46 (2012) 13253–13261,
Chem. Eng. 25 (2017) 911–923, https://doi.org/10.1016/j.cjche.2016.11.017. https://doi.org/10.1021/es303673p.
[55] A. Shockravi, V. Vatanpour, Z. Najjar, S. Bahadori, A. Javadi, A new high [66] F. Xiao, B. Wang, X. Hu, S. Nair, Y. Chen, Thin film nanocomposite membrane
performance polyamide as an effective additive for modification of antifouling containing zeolitic imidazolate framework-8 via interfacial polymerization for
properties and morphology of asymmetric PES blend ultrafiltration membranes, highly permeable nanofiltration, J. Taiwan Inst. Chem. Eng. 83 (2018) 159–167,
Microporous Mesoporous Mater. 246 (2017) 24–36, https://doi.org/10.1016/j. https://doi.org/10.1016/j.jtice.2017.11.033.
micromeso.2017.03.013. [67] Y. Liu, X. Wang, X. Gao, J. Zheng, J. Wang, A. Volodin, Y.F. Xie, X. Huang, B. Van
[56] X. Fang, J. Li, X. Li, S. Pan, X. Sun, J. Shen, W. Han, L. Wang, B. Van der Bruggen, der Bruggen, J. Zhu, High-performance thin film nanocomposite membranes
Iron-tannin-framework complex modified PES ultrafiltration membranes with enabled by nanomaterials with different dimensions for nanofiltration, J. Memb.
enhanced filtration performance and fouling resistance, J. Colloid Interface Sci. Sci. 596 (2020) 117717, https://doi.org/10.1016/j.memsci.2019.117717.
505 (2017) 642–652, https://doi.org/10.1016/j.jcis.2017.06.067. [68] H. Sun, P. Wu, Tuning the functional groups of carbon quantum dots in thin film
[57] J. Lv, G. Zhang, H. Zhang, C. Zhao, F. Yang, Improvement of antifouling nanocomposite membranes for nanofiltration, J. Memb. Sci. 564 (2018) 394–403,
performances for modified PVDF ultrafiltration membrane with hydrophilic https://doi.org/10.1016/j.memsci.2018.07.044.
cellulose nanocrystal, Appl. Surf. Sci. 440 (2018) 1091–1100, https://doi.org/ [69] R. Modi, R. Mehta, H. Brahmbhatt, A. Bhattacharya, Tailor Made Thin Film
10.1016/j.apsusc.2018.01.256. Composite Membranes: Potentiality Towards Removal of Hydroquinone from
[58] R. Abedini, Enhanced antifouling properties of poly(ethersulfone) nano-composite Water, J. Polym. Environ. 25 (2017) 1140–1146, https://doi.org/10.1007/s10924-
membrane filled with nano-clay particles, Polym. Bull. 76 (2019) 1737–1753, 016-0887-z.
https://doi.org/10.1007/s00289-018-2464-1. [70] H.R. Shahriari, S.S. Hosseini, Experimental and statistical investigation on
[59] M.H. Tajuddin, N. Yusof, N. Abdullah, M.N.Z. Abidin, W.N.W. Salleh, A.F. Ismail, fabrication and performance evaluation of structurally tailored PAN nanofiltration
T. Matsuura, N.H.H. Hairom, N. Misdan, Incorporation of layered double membranes for produced water treatment, Chem. Eng. Process. - Process Intensif.
hydroxide nanofillers in polyamide nanofiltration membrane for high performance 147 (2020) 107766, https://doi.org/10.1016/j.cep.2019.107766.

13

You might also like