You are on page 1of 39

Journal Pre-proofs

Fabrication of high-performance composite nanofiltration membranes for dye


wastewater treatment: mussel-inspired layer-by-layer self-assembly

Dongxue Guo, Yirong Xiao, Tong Li, Qingfeng Zhou, Liguo Shen, Renjie Li,
Yanchao Xu, Hongjun Lin

PII: S0021-9797(19)31266-4
DOI: https://doi.org/10.1016/j.jcis.2019.10.078
Reference: YJCIS 25572

To appear in: Journal of Colloid and Interface Science

Received Date: 20 August 2019


Revised Date: 18 October 2019
Accepted Date: 20 October 2019

Please cite this article as: D. Guo, Y. Xiao, T. Li, Q. Zhou, L. Shen, R. Li, Y. Xu, H. Lin, Fabrication of high-
performance composite nanofiltration membranes for dye wastewater treatment: mussel-inspired layer-by-layer
self-assembly, Journal of Colloid and Interface Science (2019), doi: https://doi.org/10.1016/j.jcis.2019.10.078

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Inc.


Fabrication of high-performance composite nanofiltration
membranes for dye wastewater treatment: mussel-inspired
layer-by-layer self-assembly

Dongxue Guo, Yirong Xiao, Tong Li, Qingfeng Zhou, Liguo Shen, Renjie Li, Yanchao

Xu*, Hongjun Lin*

College of Geography and Environmental Sciences, Zhejiang Normal University, Jinhua

321004, China

*Corresponding author. Tel.: +86 0579 82282273, Email address: ycxu@zjnu.edu.cn

*Corresponding author. Tel.: +86 0579 82282485. E-mail address: hjlin@zjnu.cn

1
Abstract

Inspired by the mussel adhesion mechanism, plant polyphenol tannic acid (TA) with abundant

catechol groups and hydrophilic Jeffamine (JA) containing amino groups were used in a

layer-by-layer (LBL) process to fabricate composite nanofiltration (NF) membranes in this

study. Alternately immersing a polyacrylonitrile substrate into individual TA and JA buffer

solutions could readily construct a NF membrane selective layer without any pre-treatment to

the substrate. The optimised membrane showed a high pure water permeance of 37 L m-2 h-1

bar-1 whilst maintaining rejections higher than 90% towards various dyes with molecular

weights ranging from 269 to 1017 g moL-1. Particularly, the obtained membrane exhibited

excellent anti-fouling and long-term performance attributed to the hydrophilic membrane

surface and covalent bonds in the selective layer. The novel strategy inherited the advantages

of a mussel-inspired dopamine material but overcame its disadvantages. The results disclosed

in this study not only provide a novel strategy to prepare composite NF membranes, but also

facilitate the mussel-inspired LBL design of advanced materials for environmental

applications.

Keywords: tannic acid, hydrophilicity, layer-by-layer self-assembly, nanofiltration, dye

wastewater treatment

2
1. Introduction

The ever-rising global demand for high quality and sustainable water accelerates the

concerns about the eco-friendly purification and reuse of wastewater [1-4]. Dye wastewater

generated by the textile industry is rated as the most aggressive pollutions in all the industrial

sectors [5-7]. Many dyes in the wastewater are extremely toxic and even carcinogenic and can

cause permanent damage to the receiving waters if discharged untreated. Various technologies,

such as biological degradation [8], adsorption [9], coagulation [10] and advanced oxidation

[11], have been developed to treat dye wastewater. Of these technologies, biological

degradation has a poor effect on stable aromatic and azo dyes. Meanwhile, adsorption and

coagulation usually result in an intractable sludge that requires significant land use, and

advanced oxidation tends to use oxidizing agents resulting in secondary pollution.

As a typical pressure driven membrane separation process, nanofiltration (NF) can readily

fractionate solutes with molecular weights in the range of 200 to 1000 g mol-1 from aquatic

systems, and is especially suitable for dye wastewater treatment [5, 12-14]. Compared to the

conventional technologies, NF has the advantages of low separation cost, low energy

consumption, and no secondary contaminants in the environment [15, 16]. Thin film composite

(TFC) membranes consist of a dense and ultrathin selective layer on a porous ultrafiltration

(UF) support [17-19]. Due to the fact that the selective layer and support are prepared

separately, it allows the independent optimisation of both parts to obtain an excellent

comprehensive membrane performance and has been widely studied in the field of NF. Several

approaches such as interfacial polymerization, dip-coating, surface grafting, and layer-by-layer

3
(LBL) self-assembly, have been utilised to prepare TFC NF membranes. Of these various

approaches, LBL assembly is known as a simple and versatile strategy to construct ultrathin

selective layers with precise control of layer composition and thickness, and thus has attracted

much attention [20, 21]. During the preparation process, a charged UF support is alternately

immersed into solutions containing cationic and anionic polyelectrolytes, and a selective layer

is deposited onto the support surface driven by electrostatic interaction. The support surface

has to be pre-treated with an extra complicated modification to generate the requisite net

charge before assembly, so as to adsorb the first layer. Besides, the formation of LBL selective

layers driven by covalent bonding has also been developed. Gu et al. [22] dissolved

m-phenylenediamine and trimesoyl chloride in toluene respectively, and alternately immersed

a modified polyacrylonitrile (PAN) support into the above solutions to fabricate a

high-performance polyamide TFC NF membrane. However, it still requires the hydrolysis of

the support before LBL assembly and involves a toxic solvent during assembly. In addition, the

scale of the LBL process is a challenge in applying this technology to real applications due to

the substantial fabrication time for multiple layers. One potential solution to overcome this is to

decrease the multiple layer number via optimising the preparation process.

Natural materials have always been a source of inspiration for innovation and provide

solutions to deal with tough, practical issues. Interface interaction plays an important role in

membrane surface engineering [23, 24]. The mussel-inspired dopamine material is especially

attractive due to the potential for fabricating multifunctional coatings onto nearly any substrate

surface using a simple and effective approach [25]. Since then, the mussel-inspired chemistry

4
has been widely used in membrane surface engineering [26]. The exact reaction mechanism

underlying the dopamine self-polymerization has long been the topic of scientific debate due to

the complex oxidation-reduction process, and the involvement of a series of intermediates [27,

28]. However, it is generally accepted that the covalent reactions between quinone

intermediates and amine groups via Michael addition and Schiff base reactions, the

self-crosslinking reaction of quinone, and the non-covalent interactions such as π-π stacking,

hydrogen bonding, and charge transfer interactions, contribute to the deposition of a

polydopamine coating onto material surfaces [29-32]. It has been proved that the catechol and

amine groups in the dopamine molecule play crucial roles during dopamine

self-polymerisation, and any system containing both catechol and amine groups can mimic

dopamine and be co-deposited onto various substrates by simulating a similar polymerization

mechanism [33, 34]. Plant-derived polyphenols such as tannic acid (TA), gallic acid (GA), and

epigallocatechin gallate (EGCg), extracted from various plant tissues, have also been

demonstrated to be capable of constructing a substrate-independent coating [35, 36]. Due to the

abundant catechol groups in their structure, plant polyphenols have been mixed with polymer

or organic containing amine groups to mimic dopamine self-polymerisation and fabricate NF

membrane selective layers via a co-deposition strategy. For example, Cheng et al. have

fabricated a loose NF membrane selective layer via the co-deposition of GA and branched

polyethylenimine (PEI) [37]. Xu’s group reported the preparation of TFC NF membranes via

tea catechins and chitosan [38]. Zhang et al. have demonstrated the co-deposition of EGCg and

PEI to prepare TFC NF membrane selective layers [39]. However, the co-deposition process is

5
uncontrollable due to the rapid reactions in the mixed system. In addition, only a few raw

materials are deposited on the substrate surface, while the rest are aggregated and precipitated

from solution as particles.

Herein, we demonstrated the fabrication of TFC NF membranes based on the low cost and

environmental-friendly TA and a hydrophilic Jeffamine (JA) via LBL assembly on PAN

supports. PAN supports can be directly utilised for the current LBL assembly without any

pre-treatment due to the substrate-independent coating behaviour of TA. JA is a polyethylene

glycol (PEG) like water-soluble compound possessing excellent hydrophilicity and flexible

long chains, with amino groups at both ends. The middle polymer segment in JA not only

interacts with the hydroxyl groups in TA via hydrogen bonding and drives the LBL process

[40], but also improves the surface hydrophilicity and prevents the adhesion of hydrophobic

molecules on the resultant membrane surface. The amino groups at both ends of JA can react

with the quinone intermediate from TA oxidation, via Michael addition or Schiff base

reactions, to form covalent bonds and render the JA molecules firmly embedded in the

selective layer. In addition, the application of the TFC NF membrane in dye wastewater

separation is investigated in detail.

2. Materials and methods

2.1 Materials

Polyacrylonitrile (PAN) powder containing methyl methacrylate sequence (Mw= 75,000

g/mol, 96%) was obtained from the Shanghai petrochemical company. Jeffamine (JA,

6
O,Oˊ-Bis(2-aminopropyl) polypropylene glycol-block-polyethylene

glycol-block-polypropylene glycol 400, 99%) was bought from Sigma-Aldrich. Tannic acid

(TA, 99%) was received from the Aladdin Industrial Corporation. Polyethylene glycol (PEG,

800 g/mol, 99%) was purchased from the Xilong Chemical Industrial Company.

1-methyl-2-pyrrolidone (NMP, 98%), NaH2PO4 (99%), Na2HPO4 (99%), bovine serum

albumin (BSA), ethanol (99.7%) and n-hexane (97%) were bought from the Sinopharm

Chemical Reagent Co., Ltd. Methyl orange (MO, 96%), Rose bengal sodium salt (RB, 90%),

Methylene blue (MB, 99%), Congo red (CR, 99%), Rhodamine B (RhB, 99%) and Methyl red

(MR, 99%), were purchased from the Tianjin Jinbei Fine Chemical Co., Ltd. Ultrapure water

was used throughout the experiments.

2.2 Preparation of PAN UF membranes

PAN UF membranes were prepared by a non-solvent induced phase inversion process.

The PAN powder was dried in an oven at 60 oC under vacuum for 12 hours to remove the

moisture. 19 g PAN powder and 1 g PEG 800 were dissolved in 80 g NMP and stirred at 60 oC

for 12 hours. Then the solution was allowed to stand at 60 oC for 10 hours to remove any

bubbles in the solution. After cooling to room temperature, the polymer solution was poured on

a glass plate and the membrane was scraped off, using a 200 μm scraper, and immediately

placed in deionised water at room temperature. After the phase inversion was completed, the

membrane was transferred into a second deionised water to remove any residual solvent and

was then stored for further modification.

7
2.3 Preparation of (TA/JA)n/PAN membrane

10 - 50 mg of TA and JA were dissolved in a 100 mL NaH2PO4/Na2HPO4 buffer solution

with designed pH value, respectively. A PAN UF membrane was immersed in the TA solution

for 5 - 45 min. Then, the membrane was removed and washed with deionised water for 5min.

After that, the membrane was immersed in the JA solution for 5 - 45 min. Again, the

membrane was taken out and washed with deionised water for 5min. Thus, one layer of LBL

self-assembly was achieved. The immersion and washing steps were cycled so as to deposit the

desired number of bilayers on the PAN membrane surface. Note that the solute concentration,

buffer solution pH, and membrane deposition time in the JA solution was identical with that in

the TA solution for each self-assembly. For convenience, the resultant membranes were

designated as (TA/JA)n/PAN membranes, where n is the number of bilayers. The schematic

diagram of the composite NF membranes via the LBL assembly process is shown in Fig. 1.

Fig. 1. The schematic diagram of LBL assembly process of the multilayer membrane.

2.4 Characterisations of PAN and modified membranes

The chemical structure of the membrane was characterised by using Fourier transform

infrared spectroscopy (FT-IR, NEXUS 670, USA). The chemical element composition of the

8
membrane surface was measured by an X-ray photoelectron spectrometer (XPS, ESCALAB

250Xi, Thermo Fisher Scientific, USA). The surface morphology of the membrane was

characterised by a scanning electron microscope (SEM, S-4800, Japan). The contact angles of

the different membranes were tested by a contact angle meter (Kino Co., Ltd., USA).

Membrane surface-energies were calculated according to a reported method [32].

2.5 Separation performance of the membranes

The dye solution was used as the feed solution to evaluate the separation performance of

the modified membranes. The filtration experiments in this experiment were carried out on a

dead-end stirred cell filtration device at room temperature, and a low pressure of 2 bar. The

effective area of each membrane was 44 cm2 and at least three tests of each membrane were

performed. The membranes were pre-pressed for 30 min, under a pressure of 2 bar, to achieve

a steady stable pure water permeance. Then, a 35 μM dye solution was poured into the cell as a

feed and stirred at 700 rpm to diminish the possible concentration polarisation. The membrane

permeance was calculated as shown in Eq. (1):

V
P (1)
A  t  p

where P is the permeance, V is the solution volume (L) through the membrane in time t, A is

the effective area (m2) of the membrane, t is the filtration time (h), and Δp is the filtration

pressure (bar).

The rejection of the dye was determined as shown in Eq. (2):

9
Cp
R  (1  ) 100% (2)
Cf

where R is the rejection, Cp is the concentration in the permeate, and Cf is the concentration of

the feed solution. The concentration of the dye solution was measured by using an ultraviolet

spectrophotometer (Beijing Persee, T6).

2.6 Anti-fouling performance experiment

The anti-fouling performance of the pristine membrane and the modified membrane were

evaluated by using BSA as a model protein. In this study, three cycles of filtration experiments

were carried out. Each cycle filtration experiment was divided into three steps: (1) permeation

of deionised water for 30 min at 2 bar, P1; (2) filtration of BSA solution (1 g/L) for 1 h at 2 bar,

P2; (3) the membrane was rinsed with pure water for 20 min, then permeation of pure water for

30 min, P3. The flux recovery rate (FRR) was calculated using Eq. (3):

P3
FRR(%)  100% (3)
P1

Generally, the FRR is used to characterise the anti-fouling properties of the membrane. A

higher FRR value means a stronger membrane anti-fouling ability.

In addition, the anti-fouling parameters, including total fouling ratio (Rt), reversible

fouling ratio (Rr) and irreversible fouling ratio (Rir), were calculated according Eqs. (4)-(6), to

study the anti-fouling performance of the modified membrane [41]:

P2
Rt  (1  ) 100% (4)
P1

10
P3  P2
Rr  ( ) 100% (5)
P1

P1  P3
Rir  ( ) 100% (6)
P1

where Rt is the sum of Rr and Rir, Rir is the fouling caused by the adsorption or deposition of

macromolecules on the membrane surface, and Rr is the fouling caused by concentration

polarisation.

3. Results and discussions

3.1 Chemical characterisation of membranes

As shown in Fig. 2(a), the sole TA treated membrane TA/PAN is white and appears

unchanged after TA deposition. However, it changed to dark upon treatment with aqueous

AgNO3 as a result of silver nanoparticle formation via a redox couple between Ag+ and the

phenol groups, demonstrating the presence of polyphenolic coating on the membrane surface

[35, 36]. This polyphenolic coating provides a secondary interaction platform for the following

JA assembly onto the membrane surface. Fig. 2(b) shows the ATR-FTIR spectra of the pristine

PAN membrane and the (TA/JA)2/PAN membrane. Compared with the pristine PAN

membrane, a broad peak around 3300 cm-1 attributed to the stretching vibration of the hydroxyl

groups can be found in the (TA/JA)2/PAN membrane [42, 43]. Meanwhile, the intensity of

peaks at 1190 cm-1, assigned to the Ar-OH stretching vibration and 1370 cm-1 related to the

ester C-O stretching vibration, is significantly enhanced [44, 45]. These results further confirm

the successful deposition of TA on the membrane surface. In addition, the peaks at 2926 cm-1

11
and 2858 cm-1, related to -CH3, -CH2- and -CH- groups, become more intense and a

characteristic peak at 1505 cm-1, associated with the deformation vibration of N-H, emerges

after the self-assembly suggesting success of the deposition of JA [46]. The enhanced peak

intensity at 1080 cm-1 related to the C-O stretching vibration could be due to a combined

contribution of Ar-OH in TA and C-O-C in JA [47]. Moreover, the enhanced peak at 1630 cm-1

due to the -C=N stretching vibration indicates the Schiff base reaction between TA and JA

[48]. The construction of covalent bonds would favour the membrane stability.

(a) (b) PAN


(d) 12000 (f) 5000

10000 4500
TA/PAN
Transmittance (a.u.)

8000 4000 -C=O


(TA/JA)2/PAN -C ≡ N
-C-OH
3500
Counts

Counts
6000
1630 3000
758 4000
3300
2926 1505 2500
1080
2000
1190
2000
1370
0
4000 3500 3000 2500 2000 1500 1000 404 402 400 398 396 538 536 534 532 530 528 526
Wavenumber (cm )
-1
Binding Energy (eV) Binding Energy (eV)
Ag/TA/PAN
(c) (e) 6000 (g) 20000
C 1s
O 1s 5000 16000
-C≡N
(TA/JA)2/PAN -C-OH
N 1s 12000
4000 -C=O
Counts

Counts

-C=N-
3000 8000
-NH-
PAN
4000
2000

0
900 800 700 600 500 400 300 200 100 404 402 400 398 396 538 536 534 532 530 528 526
Binding Energy (eV) Binding Energy (eV) Binding Energy (eV)

Fig. 2. (a) Digital images of TA/PAN membrane and AgNO3 treated TA/PAN membrane, (b)

ATR-FTIR and (c) XPS spectra of pristine PAN membrane and (TA/JA)2/PAN membrane, (d)

N 1s fitting spectra of PAN membrane, (e) N 1s fitting spectra of (TA/JA)2/PAN membrane, (f)

O 1s fitting spectra of PAN membrane, and (g) O 1s fitting spectra of (TA/JA)2/PAN

membrane.

XPS was further used to quantitatively characterise the membrane surface chemistry. As

shown in Fig. 2(c) and Table 1, after the LBL self-assembly, the O content increased from

6.74% to 27.14% while the N content decreased from 20.05% to 8.08%. Since both TA and JA
12
have a relatively higher O content than PAN, it is rational that the O content increased after TA

and JA were self-assembled onto the PAN membrane surface. Similarly, TA contains no N

content, and JA has a much lower N content than that of PAN, so the N content significantly

decreases after self-assembly. The N 1s deconvolution in the PAN membrane shows a single

peak at 399.04 eV, which is related to the -C≡N groups (Fig. 2(d)), while two new peaks at

399.4 eV and 401.7 eV emerge in the (TA/JA)2/TA membrane, as shown in Fig. 2(e). The

former is attributed to the -C=N- bonds, while the latter is assigned to -NH- bond. The

existence of the -C=N- bond on the membrane surface corresponds to the ATR-FTIR results

and further confirms the Schiff base reaction between TA and JA. The PAN membrane shows

two O 1s fitting peaks at 531.09 ev and 532.3 ev, which can be assigned to -C=O and -C-O-,

respectively (Fig. 2(f)). For the (TA/JA)2/PAN membrane, the ratio of the peak assigned to

-C-O- increases significantly owing to the abundant -C-OH groups in TA and the -C-O-C

groups in JA, as shown in Fig. 2(g).

Table 1. Surface chemical composition of PAN membrane and (TA/Jeffamine)2/TA membrane

obtained from XPS spectra. Three tests of each membrane were performed to obtain an average

value.

Composition (atomic%)
Membrane
C N O

PAN 73.2±1 20.1±0.2 6.7±0.4

(TA/JA)2/PAN 64.8±2 8.1±0.2 27.1±0.3

13
Accordingly, we suggested a possible reaction mechanism of TA and JA, as shown in Fig.

3. Under weak alkaline conditions, the TA was oxidised to quinoid form and gradually

deposited onto the PAN membrane surface via the plant polyphenol inspired coating ability.

Then, the covalent reactions between the amino groups in JA and the quinoid form in TA via

the Michael addition reaction or Schiff base reaction, as well as the hydrogen bonding reaction

between the phenolic hydroxy group and the ether group, contribute to the deposition of JA

onto the membrane surface.


OH OH OH OH
OH O O O
OR OR OR OR
O OR O O OR O O OR O O RO O
O OH Oxidation O O O O O O
O O
Self-polymerization O O
RO O RO O RO O O OR
OR OR OR OR
OH O O O
OH O O O

R
TA JA NH2
O NH2
O O

OH OH OH
O O O
OR OR OR
O OR O O OR O O OR O
O O O OH O O
O O O
RO O RO O RO O
OR OR OR
O O OH
O OH

Schiff base reaction Mechael addition reaction Hydrogen bonding interaction

Fig. 3. Possible reaction mechanism between TA and JA.

3.2 Surface properties of membranes

Hydrophilic membrane surfaces have been proved to form a strong interaction with water

molecules, conducive to the formation of a tightly bound water molecular layer and fouling

repellence during filtration applications. The pristine PAN membrane showed a water contact

angle of 55o (Table 2). In comparison, the water contact angle of the (TA/JA)2/PAN membrane

was reduced down to 44.5o, indicating the improved hydrophilicity after the LBL self-assembly

process. This could be attributed to the synergic effect of hydrophilic groups, such as the
14
hydroxyl and ether groups on the membrane surface, and the improved membrane surface

roughness. Membrane surface hydrophilicity was further elucidated by surface energy, and the

results can be found in Table 2. It can be seen that the (TA/JA)2/PAN membrane possesses an

increased total surface energy than that of the pristine PAN membrane, and the main

contribution to the increment is from the polar component, highlighting the role of polar groups

in the improved hydrophilicity. Fig. 3 shows that the membrane water contact angle gradually

decreases with the increase in bilayer number. Since both TA and JA are highly hydrophilic

molecules, the decreased water contact angle could be mainly related to the increased

membrane roughness, as proved in Fig. 6 and will be discussed later.

Table 2. Contact angle and surface-energy data of the PAN membrane and the (TA/JA)2/PAN

membrane

Contact angle (°) Surface-energy components (mJ m-2)


Membrane
Pure water Glycerol Diiodomethane γp γd γ

PAN 55±1.2 39.6±0.6 13.2±1.4 8.2 49.5 57.7

(TA/JA)2/PAN 44.5±0.5 65.2±1.2 23.5±0.4 15.9 46.7 62.6

15
65

60

Water contact angle (°)


55

50

45

40

35

30
0.5 1.0 1.5 2.0 2.5
Number of bilayer

Fig. 3. Water contact angle of modified membranes with different bilayer numbers. Three tests

of each membrane were performed to obtain an average value.

3.3 Morphologies of membranes

Fig. 4 shows the SEM images of the pristine PAN membrane and the self-assembled

membranes prepared under different buffer solution pH values. The pristine PAN membrane

shows a clear surface with obviously visible pores (Fig. 4(a)). After LBL self-assembly by TA

and JA at various pHs, the pores on the membranes disappear, indicating a new layer fully

covering the PAN membrane surface. As shown in Fig. 4 (b)-(d), the self-assembled membrane

surfaces prepared at pH 6 and 7 are relatively clear and smooth, while the membrane surface

obtained at pH 8 shows massive nanoaggregates. Under weak alkaline conditions, the phenols

in TA molecules can be readily oxidised to highly reactive quinones. Meanwhile, the

self-crosslinking reactions of quinones contributes to the covalent binding among aryl rings

and leads to the formation of nanoaggregates [44]. As shown in Figs. 4(f)-(h), cross-sectional

16
morphologies show that a top layer has formed on the PAN membrane surface after

self-assembly, and the top layer thickness is strongly influenced by the buffer solution pH

during the self-assembly process. The top layer obtained under pH 7 exhibits a thickness of

260 nm, which is higher than that prepared under pH 6 (120 nm) and pH 8 (220 nm). In fact,

the TA coating process is highly dependent on the solution pH values and the optimal value,

under which TA would form a thicker coating than other pH values, is pH 7 [36]. Considering

this, its rational that the thickest top layer was formed under a pH of 7 during the LBL

self-assembly process.

(a) PAN (b) pH 6 (c) pH 7 (d) pH 8

1 ηm 1 ηm 1 ηm 1 ηm

(e) PAN (f) pH 6 (g) pH 7 (h) pH 8


220 nm
120 nm

260 nm
1 ηm 1 ηm 1 ηm 1 ηm

Fig. 4. Surface and cross-sectional SEM images of membranes prepared under different buffer

solution pH values: (a), (e) the pristine PAN membrane; (b), (f) pH 6; (c), (g) pH 7; (d), (h) pH

8.

Figs. 5(a)-(d) shows the morphologies of membranes prepared at various TA and JA

concentrations. It can be seen that with the increase of monomer concentration, the membrane

surface became rough, and some significant nanoaggregates appeared on the membrane surface

when the concentration reached 0.3 g/L. Cross-sectional images of the selective layers in Figs.

17
5(e)-(h) indicate that the layer thickness increased with the monomer concentration. This could

be due to the deposition behaviour being enhanced at higher monomer concentrations, leading

to rougher surfaces and thicker top layers.

(a) 0.1 g/L (b) 0.2 g/L (c) 0.3 g/L (d) 0.5 g/L

(e) 0.1 g/L (f) 0.2 g/L (g) 0.3 g/L (h) 0.5 g/L
200 nm
220 nm
110 nm

1 ηm 1 ηm 1 ηm 1 ηm

Fig. 5. Surface and cross-sectional SEM images of membranes prepared with different

monomer concentrations: (a), (e) 0.1 g/L; (b), (f) 0.2 g/L; (c), (g) 0.3 g/L; (d), (h) 0.5 g/L.

The membrane surface nanoaggregates structure became more significant and the top

layer thickness inceased with the increase of bilayer number, as shown in Fig. 6. The

nanoaggregates increase the membrane roughness and contribute to the membrane

hydrophilicity, which is in agreement with the water contact angle results.

18
(a) 1.0 (b) 2.0 (c) 2.5

(d) 1.0 (e) 2.0 (f) 2.5


200 nm
100 nm 250 nm

1 ηm 1 ηm 1 ηm

Fig. 6. Surface and cross-sectional SEM images of membranes prepared with different bilayer

numbers: (a), (d) 1.0; (b), (e) 2.0; (c) (f) 2.5.

3.4 Separation performance of membranes

The effect of deposition time on the LBL self-assembly membrane separation

performance is shown in Fig. 7(a). The monomer concentration was fixed at 0.5 g/L, the pH of

the phosphate buffer solution was fixed at 7, and the bilayer number was 2. It can be found that

with the increase of deposition time, the water permeance of the assembled membrane

decreased, while the rejections of MO and RB gradually increased. These could be due to the

fact that the thickness of the assembled selective layer increases with the deposition time,

which results in an increased membrane permeation resistance. When the coating time was

prolonged from 5 min to 15 min, the water permeance of the membrane decreased from 40.7 L

m-2 h-1 bar-1 to 4.5 L m-2 h-1 bar-1, while the MO rejction increased from 24.5% to 71.2%, and

the RB rejection increased from 83.5% to 95.8%. Further increases in deposition time would

significantly decrease the water permeance while the dye rejection remained nearly unchanged.
19
So the coating time was fixed at 15 min for the next tests.

(a) 50
(b) 50
100 100
Water Permeance (L m h bar )
-1

Water Permeance (L m h bar )


-1
40 80 40 80
-1

-1
-2

-2
Rejection(%)

Rejection (%)
30 60 30 60
Water Permeance
MO Rejection
20 RB Rejection 40 20 40

Water permeance
10 20 10 MO rejection 20
RB rejection
0 0 0 0
0 10 20 30 40 50 6.0 6.5 7.0 7.5 8.0
Coating Time (min) pH

(c) (d)
80 100 150 100
Water Permeance (L m h bar )

Water Permeance (L m h bar )


-1

-1

80 120 80
-1

-1

60

Rejection(%)
-2

-2
Rejection(%)

60 90 Water Permeance 60
MO Rejection
40 40 60 RB Rejection
40
Water Permeance
MO Rejection 20 30 20
20
RB Rejection
0 0 0
0.1 0.2 0.3 0.4 0.5 0.5 1.0 1.5 2.0 2.5 3.0
Monomer Concentration (g/L) Bilayer Number

Fig. 7. Membrane separation performance in terms of (a) coating time, (b) solution pH, (c)

monomer concentration, and (d) bilayer number. Three tests of each membrane were

performed to obtain an average value.

The effect of phosphate buffer solution pH on the LBL self-assembly membrane

separation performance is shown in Fig. 7(b). The monomer concentration was fixed at 0.5

g/L, coating time was fixed at 15 min, and the number of bilayers was 2. As shown in Fig. 7,

the water permeance of the assembled membrane tends to decrease at first, and then increase

with the pH rise. The rejection of MO and RB shows a reverse trend. The lowest water

permeance is obtained at pH 7. The optimal coating pH of TA, under which the polyphenol
20
coating is thicker than that formed at other solutions, is pH 7. This results in the thickest

selective layer at pH 7 (as proved by the SEM images in Fig. 3) and the lowest permeance, due

to the highest permeate resistance. When the solution pH is enhanced to 8, the water

permeance of the membrane increases to 20 L m-2 h-1 bar-1, and the rejection of MO and RB

increases to 91% and 98%, respectively. Therefore, the solution pH was fixed at 8 for the

following tests.

Figs. 7(c) and 7(d) shows the increase of the MO and RB rejections, and the decrease of

water permeance with the increase in monomer concentration and bilayer number. This

phenomenon can be attributed to the fact that a low monomer concentration, or few bilayer

numbers, will restrict the formation of an integrated selective layer, while an excessive

monomer concentration, or bilayer numbers, may result in too thick a selective layer. Herein,

the optimal membrane was fabricated using a monomer concentration of 0.3 g/L and a bilayer

number of 2. This membrane showed an MO rejection of 90.5% and RB rejection of 99.5%,

with a water permeance up to 36 L m-2 h-1 bar-1.

The separation performance of the optimal membrane towards different dyes was further

investigated, and the results can be seen in Fig. 8(a). It shows that the dye rejection increases

with its molecular weight. The dyes with relatively higher molecular weight were almost

completely rejected. Some dyes with lower molecular weight, such as MR and MO, were

partly rejected, but still held rejections higher than 90%. These results demonstrate the sieving

effect in filtration processes. However, it cannot exclude the existence of charge effects during

filtration.
21
(a) 100 (b) 50
100
45
80

Permeance (L m h bar )
-1
90

RB Rejection(%)
40
Rejection (%)

-1
60 80

-2
35
40 70
30
Permeance
20 RB Rejection 60
25

0 20 50
MR MO MB RhB CR RB 0 10 20 30 40 50
269 327 374 479 697 1018 Time (hour)
-1
Dye and its molecular weight (g mol )

Fig. 8. (a) Rejection of (TA/JA)2/PAN for various dyes; (b) long-term separation performance

of (TA/JA)2/PAN membrane. Three tests of each membrane were performed to obtain an

average value.

The stability of the (TA/JA)2/PAN membrane was characterised by a long-term separation

performance using RB solution as a feed. As shown in Fig. 8(b), the RB rejection was stable all

through the filtration, and the water permeance did not change after an initial stage of

membrane compaction. These results demonstrate that the membrane has good stability over a

long-term filtration while still maintaining a remarkable separation performance.

Membrane fouling is one of the most trickiest problems in membrane processes [49-54]

and it results in many drawbacks such as permeance decline, increase in operational costs, and

membrane degeneration. Therefore, lots of efforts have been made to enhance the membrane

anti-fouling ability. In this study, the anti-fouling ability of both the pristine PAN membrane

and the (TA/JA)2/PAN membrane was evaluated with three cycle filtration tests using BSA as

a model protein. Two kinds of filtration models, time-dependent filtration, and filtrate

volume-dependent filtration, were used to measure the membrane anti-fouling performance, as

22
shown in Figs. 9(a) and 9(b). It can be seen that both membranes exhibit a lower BSA flux than

the pure water permeance due to protein fouling, while in both models, the water permeance of

the (TA/JA)2/PAN membrane can be completely recovered after cleaning.

(a) (b) 120


120 PAN
(TA/JA)2/ PAN PAN
100

)
-1
)

(TA/JA)2/ PAN

Permeance (L m h bar
-1

100
Permeance (L m h bar

80

-1
-1

80

-2
-2

60
60

40 40

20 20

0 0
0 40 80 120 160 200 240 280 0 100 200 300 400 500 600
Time (min) Filtrate volume (mL)

(c) 100 PAN (d) 100 PAN


(TA/JA)2/PAN (TA/JA)2/PAN
80 80
Percentage (%)
Percentage (%)

60 60

40 40

20 20

0 0
FRR Rt Rr Rir FRR Rt Rr Rir

Fig. 9. (a) time-dependent permeance and (b) permeate volume-dependence of the pristine

PAN membrane and the (TA/JA)2/PAN membrane under three cycles of BSA solution

filtration tests, (c) FRR, Rt, Rr and Rir from time-dependent filtration, and (d) FRR, Rt, Rr and

Rir from permeate volume-dependent filtration. Three tests of each membrane were performed

to obtain an average value.

To further investigate the anti-fouling performance of the membranes, four parameters

including FRR, Rt, Rr and Rir were calculated and are shown in Figs. 9(c) and 9(d). It is

apparent that the FRR value of the (TA/JA)2/PAN membrane is higher than that of the pristine

23
PAN membrane, indicating its stronger anti-fouling ability. The (TA/JA)2/PAN membrane has

a much lower Rir than the pristine PAN membrane, suggesting that the pristine PAN membrane

is prone to irreversible fouling. The Rr of the (TA/JA)2/PAN membrane is relatively higher,

indicating that reversible fouling is dominant for the membrane. In fact, the reversible fouling

can be readily removed by simple physical cleaning, which is beneficial for efficient filtration

processes. It has been shown that the hydrophilic surface contributes to the resistance to BSA

absorption [55]. The previous analysis (Table 2) shows that the (TA/JA)2/PAN membrane is

more hydrophilic than the pristine PAN membrane, and the highly hydrated polypropylene

glycol segments from the JA structure would take up a large amount of free water to form a

water molecule layer on the (TA/JA)2/PAN membrane outer surface, and so reduce, or even

prevent, the direct contact of BSA with the membrane [56]. This characteristic is helpful to

mitigate membrane fouling, which is generally considered as the major limitation of membrane

technology [57-62].

Table 3 lists the separation performance of some TFC NF membranes reported in the

literature and in this study. (TA/JA)2/PAN membranes prepared in this study had a much

higher pure water permeance whilst possessing a similar, or higher, dye rejection than the other

TFC NF membranes including dopamine, TA, and EGCg based membranes fabricated via

interfacial polymerisation or co-deposition methods. The significantly higher permeance of the

(TA/JA)2/PAN membrane could be mainly attributed to the fact that the hydrophilic segment of

the selective layer facilitates water molecule permeation.

Table 3. Dye removal performance comparison of some TFC NF membranes in reference to


24
(TA/JA)2/PAN prepared in this study a

Pure water
Membrane Preparation method permeance (L m-2 Dye rejection (%) Reference
h-1 bar-1)

(PDDA/GO)4/PAN LBL 5.8 CR, 99.9 [ 63]

(PS/PDMAEMA)/PSf LBL 8 CR, 99 [ 64]

(CMCNa/PEI)/PP LBL 5.7 CR, 99.4; RhB, 98 [ 65]

(Dopamine/TMC)/PES Interfacial polymerization 30 CR, 95 [ 66]

(DETA/TMC)/PES Interfacial polymerization 17 MO, 81; CR, 97 [ 67]

(TA/TMC)/PES Interfacial polymerization 23.4 MB, 71.2 [ 68]

TiO2/Ceramic Atomic layer deposition 8 RB, 96 [ 69]

(EGCg/EPI)/PES Co-deposition 19 CR, 99 [ 39]


MO, 91; MB, 98; RhB, 99; CR,
(TA/JA)2/PAN LBL 37 This work
99.5; RB, 99.5

a PDDA, GO, PS, PDMAEMA, PSF, CMCNa, PEI, PES, DETA, TMC are abbreviations of

poly(diallyldimethyl ammoniumchloride), graphene oxide, polystyrene, poly(N,N-dimethylaminoethyl

methacrylate), polysulfone, sodiumcarboxymethylcellulose, polyethylenimine, polyethersulfone,

diethylenetriamine, trimesoyl chloride.

4. Conclusions

Inspired by the mussel adhesion mechanism, where catechol and amino groups interact via

various covalent and non-covalent bonds, a strategy which alternately assembles a plant

polyphenol tannic acid (TA) with abundant catechol groups and hydrophilic Jeffamine (JA)

containing amino groups, onto a polyacrylonitrile (PAN) substrate membrane, was developed
25
in this study. Accordingly, novel (TA/JA)n/PAN membranes were successfully fabricated via

this novel mussel-inspired LBL strategy. Compared with the previously reported

mussel-inspired co-deposition strategy, which is uncontrollable and generates a lot of waste in

solution [37-39], the LBL process in this study is controllable and has a high atomic economy.

The hydrogen bonding interactions and covalent reactions between TA and JA acted as the

driving force of the LBL process. The optimal membrane can be obtained under a deposition of

15 min, a buffer solution pH of 8, a monomer concentration of 0.3 g moL-1 and a bilayer

number of 2. The resulting membrane showed a high pure water permeance of 37 L m-2 h-1

bar-1 and dye rejection, toward various dyes, of higher than 90%. In addition, the membrane

exhibited an excellent anti-fouling performance due to the hydrophilic membrane surface. Most

importantly, the prepared membrane demonstrated good stability during a long-term separation

process which could be attributed to the covalent bonding between TA and JA in the selective

layer. This mussel-inspired LBL strategy also can be used to fabricate high performance

ultrafiltration or microfiltration in further work. The novel, mussel-inspired, LBL assembly

method described here offers a new incentive not only to nanofiltration membrane fabrication,

but also to the surface engineering of advanced materials in various applications.

Acknowledgements

This study was financially supported by the Natural Science Foundation of Zhejiang province,

China (No. LQ19B060008) and the National Natural Science Foundation of China (Nos.

51978628, 51578509).

26
References

[1] P.J.J. Alvarez, C.K. Chan, M. Elimelech, N.J. Halas, D. Villagran, Emerging opportunities

for nanotechnology to enhance water security, Nat. Nanotechnol. 13 (2018) 634-641.

[2] M.S. Mauter, I. Zucker, F.o. Perreault, J.R. Werber, J.-H. Kim, M. Elimelech, The role of

nanotechnology in tackling global water challenges, Nat. Sustain. 1 (2018) 166-175.

[3] X. Yang, Z. Wang, L. Shao, Construction of oil-unidirectional membrane for integrated oil

collection with lossless transportation and oil-in-water emulsion purification, J. Membr.

Sci. 549 (2018) 67-74.

[4] W. Pronk, A. Ding, E. Morgenroth, N. Derlon, P. Desmond, M. Burkhardt, B. Wu, A.G.

Fane, Gravity-driven membrane filtration for water and wastewater treatment: A review,

Water Res. 149 (2019) 553-565.

[5] H. Fan, J. Gu, H. Meng, A. Knebel, J. Caro, High-Flux Membranes Based on the Covalent

Organic Framework COF-LZU1 for Selective Dye Separation by Nanofiltration, Angew.

Chem. Int. Ed. 57 (2018) 4083-4087.

[6] Y. Zhang, W. Yu, R. Li, Y. Xu, L. Shen, H. Lin, B.-Q. Liao, G. Wu, Novel conductive

membranes breaking through the selectivity-permeability trade-off for Congo red

removal, Sep. Purif. Technol. 211 (2019) 368-376.

[7] Y. Xu, M. Tognia, D. Guo, L. Shen, R. Li, H. Lin, Facile preparation of

polyacrylonitrile-co-methylacrylate based integrally skinned asymmetric nanofiltration

membranes for sustainable molecular separation: An one-step method, J. Colloid Interf.

Sci. 546 (2019) 251-261.

27
[8] P. Aravind, V. Subramanyan, S. Ferro, R. Gopalakrishnan, Eco-friendly and facile

integrated biological-cum-photo assisted electrooxidation process for degradation of

textile wastewater, Water Res. 93 (2016) 230-241.

[9] P. Senthil Kumar, S.J. Varjani, S. Suganya, Treatment of dye wastewater using an

ultrasonic aided nanoparticle stacked activated carbon: Kinetic and isotherm modelling,

Bioresour. Technol. 250 (2018) 716-722.

[10] L. Zhou, H. Zhou, X. Yang, Preparation and performance of a novel starch-based

inorganic/organic composite coagulant for textile wastewater treatment, Sep. Purif.

Technol. 210 (2019) 93-99.

[11] J. Khatri, P.V. Nidheesh, T.S. Anantha Singh, M. Suresh Kumar, Advanced oxidation

processes based on zero-valent aluminium for treating textile wastewater, Chem. Eng. J.

348 (2018) 67-73.

[12] W. Ye, J. Lin, R. Borrego, D. Chen, A. Sotto, P. Luis, M. Liu, S. Zhao, C.Y. Tang, B. Van

der Bruggen, Advanced desalination of dye/NaCl mixtures by a loose nanofiltration

membrane for digital ink-jet printing, Sep. Purif. Technol. 197 (2018) 27-35.

[13] Z. Thong, J. Gao, J.X.Z. Lim, K.-Y. Wang, T.-S. Chung, Fabrication of loose

outer-selective nanofiltration (NF) polyethersulfone (PES) hollow fibers via single-step

spinning process for dye removal, Sep. Purif. Technol. 192 (2018) 483-490.

[14] M. Peydayesh, T. Mohammadi, O. Bakhtiari, Effective treatment of dye wastewater via

positively charged TETA-MWCNT/PES hybrid nanofiltration membranes, Sep. Purif.

Technol. 194 (2018) 488-502.

28
[15] N. Dizge, R. Epsztein, W. Cheng, C.J. Porter, M. Elimelech, Biocatalytic and salt

selective multilayer polyelectrolyte nanofiltration membrane, J. Membr. Sci. 549 (2018)

357-365.

[16] C. Completo, V. Geraldes, V. Semião, M. Mateus, M. Rodrigues, Comparison between

microfluidic tangential flow nanofiltration and centrifugal nanofiltration for the

concentration of small-volume samples, J. Membr. Sci. 578 (2019) 27-35.

[17] L. Bai, Y. Liu, A. Ding, N. Ren, G. Li, H. Liang, Fabrication and characterization of

thin-film composite (TFC) nanofiltration membranes incorporated with cellulose

nanocrystals (CNCs) for enhanced desalination performance and dye removal, Chem.

Eng. J. 358 (2019) 1519-1528.

[18] K.H. Mah, H.W. Yussof, M.N. Abu Seman, A.W. Mohammad, Optimisation of interfacial

polymerization factors in thin-film composite (TFC) polyester nanofiltration (NF)

membrane for separation of xylose from glucose, Sep. Purif. Technol. 209 (2019)

211-222.

[19] R. Zhang, M. He, D. Gao, Y. Liu, M. Wu, Z. Jiao, Y. Su, Z. Jiang, Polyphenol-assisted

in-situ assembly for antifouling thin-film composite nanofiltration membranes, J. Membr.

Sci. 566 (2018) 258-267.

[20] Q. Zhang, S. Chen, X. Fan, H. Zhang, H. Yu, X. Quan, A multifunctional graphene-based

nanofiltration membrane under photo-assistance for enhanced water treatment based on

layer-by-layer sieving, Appl. Catal. B-Environ. 224 (2018) 204-213.

[21] Y. Huang, J. Sun, D. Wu, X. Feng, Layer-by-layer self-assembled chitosan/PAA

29
nanofiltration membranes, Sep. Purif. Technol. 207 (2018) 142-150.

[22] J.E. Gu, S. Lee, C.M. Stafford, J.S. Lee, W. Choi, B.Y. Kim, K.Y. Baek, E.P. Chan, J.Y.

Chung, J. Bang, J.H. Lee, Molecular layer-by-layer assembled thin-film composite

membranes for water desalination, Adv. Mater. 25 (2013) 4778-4782.

[23] S.B. Darling, Perspective: Interfacial materials at the interface of energy and water, J.

Appl. Phys. 124 (2018).

[24] 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).

[25] Y. Xu, F. You, H. Sun, L. Shao, Realizing Mussel-Inspired Polydopamine Selective Layer

with Strong Solvent Resistance in Nanofiltration toward Sustainable Reclamation, ACS

Sustainable Chemistry & Engineering 5 (2017) 5520-5528.

[26] Z. Wang, H.-C. Yang, F. He, S. Peng, Y. Li, L. Shao, S.B. Darling, Mussel-Inspired

Surface Engineering for Water-Remediation Materials, Matter 1 (2019) 115-155.

[27] H.-C. Yang, R.Z. Waldman, M.-B. Wu, J. Hou, L. Chen, S.B. Darling, Z.-K. Xu,

Dopamine: Just the Right Medicine for Membranes, Adv. Funct. Mater. 28 (2018)

1705327.

[28] J. Li, S. Yuan, J. Wang, J. Zhu, J. Shen, B. Van der Bruggen, Mussel-inspired

modification of ion exchange membrane for monovalent separation, J. Membr. Sci. 553

(2018) 139-150.

[29] Y. Liu, K. Ai, L. Lu, Polydopamine and its derivative materials: synthesis and promising

30
applications in energy, environmental, and biomedical fields, Chem. Rev. 114 (2014)

5057-5115.

[30] M. Muller, B. Kessler, Deposition from dopamine solutions at Ge substrates: an in situ

ATR-FTIR study, Langmuir 27 (2011) 12499-12505.

[31] Y. Ding, L.T. Weng, M. Yang, Z. Yang, X. Lu, N. Huang, Y. Leng, Insights into the

aggregation/deposition and structure of a polydopamine film, Langmuir 30 (2014)

12258-12269.

[32] X. Zhao, N. Jia, L. Cheng, L. Liu, C. Gao, Dopamine-induced biomimetic mineralization

for in situ developing antifouling hybrid membrane, J. Membr. Sci. 560 (2018) 47-57.

[33] Y.C. Xu, Z.X. Wang, X.Q. Cheng, Y.C. Xiao, L. Shao, Positively charged nanofiltration

membranes via economically mussel-substance-simulated co-deposition for textile

wastewater treatment, Chem. Eng. J. 303 (2016) 555-564.

[34] Y.C. Xu, Y.P. Tang, L.F. Liu, Z.H. Guo, L. Shao, Nanocomposite organic solvent

nanofiltration membranes by a highly-efficient mussel-inspired co-deposition strategy, J.

Membr. Sci. 526 (2017) 32-42.

[35] T.S. Sileika, D.G. Barrett, R. Zhang, K.H. Lau, P.B. Messersmith, Colorless

multifunctional coatings inspired by polyphenols found in tea, chocolate, and wine,

Angew. Chem. Int. Ed. 52 (2013) 10766-10770.

[36] D.G. Barrett, T.S. Sileika, P.B. Messersmith, Molecular diversity in phenolic and

polyphenolic precursors of tannin-inspired nanocoatings, Chem. Commun. 50 (2014)

7265-7268.

31
[37] X.Q. Cheng, Z.X. Wang, J. Guo, J. Ma, L. Shao, Designing Multifunctional Coatings for

Cost-Effectively Sustainable Water Remediation, ACS Sustainable Chemistry &

Engineering 6 (2018) 1881-1890.

[38] W.-Z. Qiu, Q.-Z. Zhong, Y. Du, Y. Lv, Z.-K. Xu, Enzyme-triggered coatings of tea

catechins/chitosan for nanofiltration membranes with high performance, Green

Chemistry 18 (2016) 6205-6208.

[39] N. Zhang, B. Jiang, L. Zhang, Z. Huang, Y. Sun, Y. Zong, H. Zhang, Low-pressure

electroneutral loose nanofiltration membranes with polyphenol-inspired coatings for

effective dye/divalent salt separation, Chem. Eng. J. 359 (2019) 1442-1452.

[40] P. Shi, X. Hu, Y. Wang, M. Duan, S. Fang, W. Chen, A PEG-tannic acid decorated

microfiltration membrane for the fast removal of Rhodamine B from water, Sep. Purif.

Technol. 207 (2018) 443-450.

[41] V.V. M. Safarpour, A. Khataee, Preparation and characterization of graphene oxide/TiO2

blended PES nanofiltration membrane with improved antifouling and separation

performance, Desalination 393 (2016) 65–78.

[42] S. Kim, E. Shamsaei, X. Lin, Y. Hu, G.P. Simon, J.G. Seong, J.S. Kim, W.H. Lee, Y.M.

Lee, H. Wang, The enhanced hydrogen separation performance of mixed matrix

membranes by incorporation of two-dimensional ZIF-L into polyimide containing

hydroxyl group, J. Membr. Sci. 549 (2018) 260-266.

[43] R. Iwasa, T. Suizu, H. Yamaji, T. Yoshioka, K. Nagai, Gas separation in polyimide

membranes with molecular sieve-like chemical/physical dual crosslink elements onto the

32
top of surface, J. Membr. Sci. 550 (2018) 80-90.

[44] Z.J. G. Liu, C. Chen, L. Hou, B. Gao, H. Yang, H. Wu, F. Pan, P. Zhang, X. Cao,

Preparation of ultrathin, robust membranes through reactive layer-by-layer (LbL)

assembly for pervaporation dehydration, J. Membr. Sci. 537 (2017) 229–238.

[45] Y.-S.C. H.J. Kim, M.-Y. Lim, K. H. Jung, D.-G. Kim, J.-J. Kim, H. Kang, J.-C. Lee,

Reverse osmosis nanocomposite membranes containing graphene oxides coated by tannic

acid with chlorine-tolerant and antimicrobial properties, J. Membr. Sci. 514 (2016) 25–

34.

[46] Y.D. Y. Lv, Z.-X. Chen, W.-Z. Qiu, Z.-K. Xu, Nanocomposite membranes of

polydopamine/electropositive nanoparticles/ polyethyleneimine for nanofiltration, J.

Membr. Sci. 545 (2018) 99–106.

[47] H.J. D.L. Shaffer, S. R.-V. Castrillón, X. Lu, M. Elimelech, Post-fabrication modification

of forward osmosis membranes with a poly(ethylene glycol) block copolymer for

improved organic fouling resistance, J. Membr. Sci. 490 (2015) 209–219.

[48] G. Zin, J. Wu, K. Rezzadori, J.C.C. Petrus, M. Di Luccio, Q. Li, Modification of

hydrophobic commercial PVDF microfiltration membranes into superhydrophilic

membranes by the mussel-inspired method with dopamine and polyethyleneimine, Sep.

Purif. Technol. 212 (2019) 641-649.

[49] J. Teng, M. Zhang, K.-T. Leung, J. Chen, H. Hong, H. Lin, B.-Q. Liao, A unified

thermodynamic mechanism underlying fouling behaviors of soluble microbial products

(SMPs) in a membrane bioreactor, Water Res. 149 (2019) 477-487.

33
[50] H. Lin, W. Gao, F. Meng, B.-Q. Liao, K.-T. Leung, L. Zhao, J. Chen, H. Hong, Membrane

bioreactors for industrial wastewater treatment: a critical review, Crit. Rev. Environ. Sci.

Technol. 42 (2012) 677-740.

[51] M. Zhang, W. Peng, J. Chen, Y. He, L. Ding, A. Wang, H. Lin, H. Hong, Y. Zhang, H.

Yu, A new insight into membrane fouling mechanism in submerged membrane

bioreactor: Osmotic pressure during cake layer filtration, Water Res. 47 (2013)

2777-2786.

[52] J. Chen, M. Zhang, F. Li, L. Qian, H. Lin, L. Yang, X. Wu, X. Zhou, Y. He, B.-Q. Liao,

Membrane fouling in a membrane bioreactor: High filtration resistance of gel layer and

its underlying mechanism, Water Res. 102 (2016) 82-89.

[53] J. Teng, L. Shen, Y. He, B.-Q. Liao, G. Wu, H. Lin, Novel insights into membrane fouling

in a membrane bioreactor: Elucidating interfacial interactions with real membrane

surface, Chemosphere 210 (2018) 769-778.

[54] M. Zhang, H. Hong, H. Lin, L. Shen, H. Yu, G. Ma, J. Chen, B.-Q. Liao, Mechanistic

insights into alginate fouling caused by calcium ions based on terahertz time-domain

spectra analyses and DFT calculations, Water Res. 129 (2018) 337-346.

[55] E. Ostuni, R.G. Chapman, R.E. Holmlin, S. Takayama, G.M. Whitesides, A survey of

structure− property relationships of surfaces that resist the adsorption of protein,

Langmuir 17 (2001) 5605-5620.

[56] S. Kang, A. Asatekin, A. Mayes, M. Elimelech, Protein antifouling mechanisms of PAN

UF membranes incorporating PAN-g-PEO additive, J. Membr. Sci. 296 (2007) 42-50.

34
[57] Y. Chen, J. Teng, L. Shen, G. Yu, R. Li, Y. Xu, F. Wang, B.-Q. Liao, H. Lin, Novel

insights into membrane fouling caused by gel layer in a membrane bioreactor: Effects of

hydrogen bonding, Bioresour. Technol. 276 (2019) 219-225.

[58] L. Shen, Y. Zhang, W. Yu, R. Li, M. Wang, Q. Gao, J. Li, H. Lin, Fabrication of

hydrophilic and antibacterial poly(vinylidene fluoride) based separation membranes by a

novel strategy combining radiation grafting of poly(acrylic acid) (PAA) and electroless

nickel plating, J. Colloid Interf. Sci. 543 (2019) 64-75.

[59] W. Yu, Y. Liu, Y. Xu, R. Li, J. Chen, B.-Q. Liao, L. Shen, H. Lin, A conductive PVDF-Ni

membrane with superior rejection, permeance and antifouling ability via electric assisted

in-situ aeration for dye separation, J. Membr. Sci. 581 (2019) 401-412.

[60] Z. Zhao, Y. Lou, Y. Chen, H. Lin, R. Li, G. Yu, Prediction of interfacial interactions

related with membrane fouling in a membrane bioreactor based on radial basis function

artificial neural network (ANN), Bioresour. Technol. 282 (2019) 262-268.

[61] R. Li, Y. Lou, Y. Xu, G. Ma, B.-Q. Liao, L. Shen, H. Lin, Effects of surface morphology

on alginate adhesion: molecular insights into membrane fouling based on XDLVO and

DFT analysis, Chemosphere 233 (2019) 373-380.

[62] Y. Chen, G. Yu, Y. Long, J. Teng, X. You, B.-Q. Liao, H. Lin, Application of radial basis

function artificial neural network to quantify interfacial energies related to membrane

fouling in a membrane bioreactor, Bioresour. Technol. 293 (2019) 122103.

[63] L. Wang, N. Wang, J. Li, J. Li, W. Bian, S. Ji, Layer-by-layer self-assembly of

polycation/GO nanofiltration membrane with enhanced stability and fouling resistance,

35
Sep. Purif. Technol. 160 (2016) 123-131.

[64] J. Diep, A. Tek, L. Thompson, J. Frommer, R. Wang, V. Piunova, J. Sly, Y.-H. La,

Layer-by-layer assembled core–shell star block copolymers for fouling resistant water

purification membranes, Polymer 103 (2016) 468-477.

[65] Q. Chen, P. Yu, W. Huang, S. Yu, M. Liu, C. Gao, High-flux composite hollow fiber

nanofiltration membranes fabricated through layer-by-layer deposition of oppositely

charged crosslinked polyelectrolytes for dye removal, J. Membr. Sci. 492 (2015)

312-321.

[66] J. Zhao, Y. Su, X. He, X. Zhao, Y. Li, R. Zhang, Z. Jiang, Dopamine composite

nanofiltration membranes prepared by self-polymerization and interfacial

polymerization, J. Membr. Sci. 465 (2014) 41-48.

[67] Y. Li, Y. Su, Y. Dong, X. Zhao, Z. Jiang, R. Zhang, J. Zhao, Separation performance of

thin-film composite nanofiltration membrane through interfacial polymerization using

different amine monomers, Desalination 333 (2014) 59-65.

[68] Y. Zhang, Y. Su, J. Peng, X. Zhao, J. Liu, J. Zhao, Z. Jiang, Composite nanofiltration

membranes prepared by interfacial polymerization with natural material tannic acid and

trimesoyl chloride, J. Membr. Sci. 429 (2013) 235-242.

[69] H. Chen, S. Wu, X. Jia, S. Xiong, Y. Wang, Atomic layer deposition fabricating of

ceramic nanofiltration membranes for efficient separation of dyes from water, AIChE J.

64 (2018) 2670-2678.

36
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

37
Graphical abstract

dye solution

TA
JA

38

You might also like