Uploaded by restudamaru1998

Jurnal Proses Membran Ultrafiltrasi

advertisement
JMRTEC-1467; No. of Pages 10
ARTICLE IN PRESS
J MATER REs TECHNOL.
2 0 2 0;x x x(x x):xxx–xxx
Available online at www.sciencedirect.com
www.jmrt.com.br
Original Article
Synthesis of new low-cost organic ultrafiltration
membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes
removal
Said Benkhayaa,∗, Souad M’rabetb, Rachid Hsissoua, Ahmed El Harfia
a
b
LABORATORy Agro-Resources, OrGANIC Polymers AND Process Engineering, FACULTY of Sciences, University Ibn TofAIL, KenitrA, Morocco
LABORATORy of Geosciences AND Environment, FACULTY of Sciences, University Ibn TofAIL, KenitrA, Morocco
A R T I C L E
I N F O
A B
s
T R A C T
Article history:
This study concerns the synthesis and characterization of a new organic ultrafiltration
Received 19 January 2020
membrane by phase inversion process based on the physical blends of Polysul-
Accepted 29 February 2020
fone/Polyetherimide (PS/PEI). The organic membranes we have synthesized are innovative;
Available online xxx
their first performance is to process soluble azoic dyes. They also have a low preparation
cost and easy handling. We varied from 5 to 20 wt.% of PEI compared to the wt.% of PS.
Keywords:
The microstructural membrane structure obtained was characterized by Fourier Transform
Organic ultrafiltration membrane
Infrared spectroscopy (FTIR) and by nuclear magnetic resonance spectroscopy (NMR). The
Phase inversion
Polysulfone/Polyetherimide
measurement of the contact angle characterized the surface properties of the membrane
Azoic dyes
Acid orange 74
Rejection rate
pared membrane was evaluated by ultrafiltration of solutions of acid orange 74 (AO-74) and
and the morphology was evaluated using the scanning electron microscope (SEM). The premethyl orange (MO). The effect of adding polyetherimide in addition to the polysulfone
increases the retention rate of the two dyes at a pressure of 4 bar. The results of the experimental filtration of the optimized membrane (prepared with 20 wt.% of PEI) have a water
permeability of 22.66 L/hm2 bar, a maximum rejection rate of RM20 (AO-74) = 68.5% and RM20
(OM) = 64.7% respectively and a good high dye flow of around 10,08 L/h m 2 for orange-74
acid AO-74 and 16,30 L/h m2 for methyl orange (OM).
© 2020 The Authors. Published by Elsevier B.V. This is an open access article under the
CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
AbbrEVIATIONS: PS, Polysulfone; PEI, Polyetherimide; AO-74, acid orange 74; MO, methyl orange; FTIR, Transform Infrared Spectroscopy;
NMR, nuclear magnetic resonance spectroscopy; SEM, scanning electron microscope; RM, rejection rate; Jw, permeate flux; Jp, water
permeability.
∗ Corresponding AUTHOR.
E-mail: [email protected] (B. Said).
https://doi.org/10.1016/j.jmrt.2020.02.102
2238-7854/© 2020 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
2
1.
ARTICLE IN PRESS
J MATER REs TECHNOL.
Introduction
The surface of the terrestrial globe is 70% covered by the waters
of the oceans and seas and only a proportion of 3% represents the fresh water reserves which are distributed in the
glaciers, the underground waters and a tiny quantity in the
lakes and the rivers. As a result, the shortage of drinking water
and wastewater treatment have attracted a lot of attention
because of their importance in daily life and the growth of use
in various industrial fields. Besides, the reuse of wastewater
has become a considerable challenge, especially in countries
where water resources are limited as in our country (Morocco)
[1]. Accelerated water quality change, due to industrial pollution, is one of the major environmental concerns throughout
the world [2].
The objective of this treatment is to obtain an environmentally friendly quality of the treated effluents and sludge for
disposal and/or reuse [3]. Most of these industries discharge
discharges rich in organic and inorganic micropollutants without prior treatment. The textile industry generates a huge
volume of wastewater which contains a variety of toxic agents
like synthetic dyes, wetting agents, surfactants, etc. Among
the dyes synthesized, there are azo dyes. These are dyes widely
used in the manufacture of textiles and other consumer industries. They are dangerous for health (carcinogenic to humans)
and the environment for people at low concentrations [4,5].
These dyes have been used in the textile industry for the dyeing of textile products [6–9]. They are generally applied in
a solution that is aqueous [10]. The existence of these dyes
in textile wastewater has an undesirable impact on aquatic
life, human health and the environment due to their complex aromatic molecular structure [11,12]. In addition to that,
raw effluent textiles show significantly higher cytogenetic and
mutagenic effects [13].
Current techniques for removing dyes from industrial
textile wastewater include conventional treatment methods including, adsorption [14], chemical oxidation [15],
coagulation/flocculation [16,17], ozonation [18–21], biological
treatment [13,22], electrocoagulation [23], advanced oxidation
process [24], Photocatalytic degradation and the Fenton type
process [25], etc. The latter are often used to treat organic pollutants including synthetic organic dyes [25]. Chham et al.
[26,27], to use a natural adsorbent based on Moroccan bituminous shale in an adsorption process to treat dyes in the
aqueous solution (case of methylene blue). Adsorption on bentonite followed by electroflotation. While Ayach is working
on adsorption of methylene blue on bituminous schists from
Tarfaya–Boujdour [28], other work is done by Djehaf et al. [29],
using a combined adsorption treatment on bentonite followed
by an electroflotation step. In addition to that, carbon black
has many pore structures and is easy to aggregate, which provides a hierarchical porous structure. This is beneficial for the
adsorption of dyes, which contributes to the catalytic reaction
[30]. According to Iqbal [13] the ozonation is an effective treatment method for reducing the toxicity of textile wastewater
while V. faba bioassay was found to be sensitive and reliable in
determining the toxicity of raw and treated textile effluents.
According to the same author [13], biologically treated
effluents are safe, suggesting that the performance of biolog-
2 0 2 0;x x x(x x):xxx–xxx
ical treatment reduces the toxicity and sensitivity of V. faba
bioassay to reveal the reduction in contaminant load. Abbas
et al. [31] have been confirmed that the evaluation of the toxicity of different types of effluents from the textile industry
(dyeing and textile finishing plant in Ayazaga, Istanbul, Turkey)
by V. fischeri in LUMIStox 300 is efficient and useful for measuring the toxicity of chemicals used in textile finishing. These
procedures were used to remove azo dyes from wastewater
[32]. According to Daij et al. [23] has been confirmed that, the
electrocoagulation process and an efficient and inexpensive
method for treating drinking water and textile wastewater.
However, they have been shown to be inadequate due to the
molecular structures of these dyes, which resist biodegradation [33].
Membrane separation technology is one of the most
promising means in terms of wastewater treatment as well
as the production of high-quality water [34]. This importance
of membrane separation lies in several advantages such as
high efficiency, low environmental footprint and the nonuse of chemicals, as well as the ease of handling [1,35]. This
technology has various applications concerning the pharmaceutical industries [36], the agro-food industries [37,38], water
treatment [1,39–41], and industrial effluents, the chemical,
electronic and nuclear industries, desalination [42,43], gas
separation [44], medicine [45], as well as seawater desalination
[42,46].
Many organic membranes are currently available [47–51].
The most widely used organic materials are polymers of synthetic or natural origin, cross-linked or not [23,26]. According
to Iqbal et al. [52], a series of mixed membranes based on
chitosan, guar gum and polyvinyl alcohol were prepared by
the solution casting method. These membranes were used to
control the release of the drug. Organic membranes based on
Polysulfone and/or Polyetherimide have been widely used as
an ultrafiltration membrane material because of their excellent properties, such as good chemical stability, good thermal
stability, low resistivity and unique mechanical properties
[53–57]. Therefore, studies related to blends of Polysulfone
with other commonly used polymers such as Polyetherimide
are reported [58]. Hence, studies on combinations of Polyetherimide (PEI) with other widely used polymers like cellulose
acetate are published [59].
According to numerous studies were carried out to modify
the PS membranes to improve the separating performance,
the surface properties (hydrophilicity and roughness), the
porosity as well as the mechanical resistance [61]. The membranes used in water treatment are porous or dense and must
allow the removal of contaminants, under transmembrane
pressure or a constant permeate flow [13–16]. The various
membrane separation techniques [62–65], in particular ultrafiltration, were found to be most effective in the treatment
of textile wastewater compared to other membrane processes
with elevated removal efficiencies [35,41]. The UF membrane
process is widely used for the removal of organic soluble dyes
[66,67].
This work reports the preparation and the characterization
of new low-cost ultrafiltration organic membranes according
to a scientific approach which consists in gradually changing
the components of the membranes and testing their performance, particularly in terms of high-water permeability
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
ARTICLE IN PRESS
J MATER REs TECHNOL.
3
2 0 2 0;x x x(x x):xxx–xxx
Table 1 – Structure of each monomer used.
Chemical structure of PS and PEI
performance, flux as well as retention of dyes in solution. This
direction was developed from the mixtures of PS and PEI by
the phase inversion technique. The influence of PEI by weight
has been studied. It should be noted that the microstructural
characterization (NMR and FTIR) of physical blends of PS/PEI
has been described elsewhere [1]. The membranes prepared
were characterized by the contact angle and the SEM. The
membranes obtained were applied for the ultrafiltration of
solutions of two textile dyes of orange acid 74 (AO-74) and
methyl orange (OM) [1].
2.
Material and methods
2.1.
Chemical materials
Fig. 1 – Phase-inversion process.
Fig. 2 – Ultrafiltration set up.
The granules of Polysulfone (PS, P-1700) and of Polyetherimide
(trade name PEI-ULTEM 1000) were respectively supplied by
Solvay and General Electric Plastics. The two polymers were
dried in an oven at 100 ◦C for one hour, then put in a dryer
before use. Table 1 shows the structure of each monomer used
in the preparation of the membranes.
Dichloromethane (CH2C12, 99. 80 wt.%), acid orange
74 (C16H11CrN5NaO8S, 95 wt.%) and methyl orange
(C14H12N3O3NaS, 95 wt.%) were supplied by Sigma Aldrich
Chemical. The additional characteristics of acid orange 74
(AO-74) and methyl orange (MO) are shown in Table 2. Fresh
ultrapure water was used in all stages of the preparation of
the aqueous solutions.
2.2.
Preparation of membrane
In this work, we have synthesized in increasing order in wt.%
of PS ranging from 80 to wt.% 95 in addition with wt.% PEI, four
homogeneous membranes named (MH5, MH10, MH15, MH20) in
the objective to assess their separating performance. These
membranes were developed by the phase inversion process
[68–70].
The blends PS/PEI was taken up in a dry flat-bottomed glass
flask in dichloromethane with magnetic stirring for 1 h at room
temperature until solutions of homogeneous polymers were
obtained. The prepared solution was spread on a glass plate
to form a thin film. After that, the whole was evaporated in
the air for a few minutes before being immersed in a bath of
non-solvent, Fig. 1 [70–73].
2.3.
Ultrafiltration study
All ultrafiltration experiments were carried out using a laboratory Amicon made of stainless steel and connected to a
solution tank under nitrogen pressure. The effective volume
of the whole system is 0.5 L. The schematic ultrafiltration system is shown in the Fig. 2. The transmembrane pressure used
varied from 0 to 4 bar and controlled by a pressure regulator.
It should be noted that the four membranes were used for the
same filtration test and that the values obtained are the arithmetic mean of the four values. During the UF tests, the feed
solution was kept stirring at room temperature.
The performance of the membranes prepared was evaluated by the discolouring of solutions containing AO-74 and
OM with a concentration of 50 ppm. The filtration test was
carried out at a constant transmembrane pressure (4 bar) for 2
h. During filtration, the permeate was sampled every 20 min to
measure the permeate flux noted Jw (L h−1 m−2), water permeability Jp (L h−1 m−2 bar) and dye rejection R (%) are calculated
using the Eqs. (1)–(3).
Jw =
V
(1)
S..At
Jw = Jp.AP
(2)
Where Jw is the permeate flux;
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
4
ARTICLE IN PRESS
J MATER REs TECHNOL.
2 0 2 0;x x x(x x):xxx–xxx
Table 2 – Physico-chemical characteristics of the dyes.
Dyes
Molecular formula
AO-74
OM
Chemical structure
hMAX(nm)
Molecular weight (g/mol)
C16H11CrN5NaO8S
455
508, 34
C14H14N3NaO3S
616
327, 33
Fig. 3 – SEM images, top view of organic membranes of PS/PEI with (a) 5 wt.% and (b) 10 wt.% of PEI.
In this work, the organic membranes of UF have been identified by fundamental techniques which are commonly used in
the characterization of membranes. The FTIR, NMR, SEM and
the contact angle make it possible to understand the chemical
composition, the morphology as well as the hydrophilicity of
the membranes prepared PS/PEI.
a view above the organic membrane prepared according to
various wt.% of PEI relative to the PS.
According to Figs. 3 and 4, the morphology of the UF
membranes made by the PS/PEI blends, revealed by the SEM,
indicates the presence of a slightly porous structure with a
few pores. In addition, these images reveal that the PS/PEI
membranes are homogeneous. It should be noted that the
same behavior was observed by Benkhaya et al., in the
case of depositing the same blend on ceramic support by a
spin/spray-coating method [75], which could confirm that PEI
is completely miscible with PS. The porous properties measured by SEM for a weight fraction of 5 and 15 wt.% by weight of
PEI seem similar to those obtained for macroporous polymers
with a porous structure.
Fig. 3(a) and (b) show that the surface of the organic membrane has a relatively smooth and dense structure with the
appearance of some randomly distributed pores. Increasing
the PEI concentration reduces pore diameters. For the membranes synthesized at 15% and 20% by weight of PEI (Fig. 4(c)
and (d)), they have the highest concentration of PEI polymer
and have the densest structure, in comparison with the membranes at 5% and 10% by weight of PEI.
3.
Results and discussion
3.2.
3.1.
Membrane morphology
The most important surface property of the membrane
is its hydrophilicity or its hydrophobic character [76]. The
hydrophilicity of the membrane can play a crucial role in filtration performance, in particular from the point of view of
V (m3) indicates the volume of the permeate over the area
of the effective membrane S (m2) and t is the time difference
(h).
.
R = 100* 1 −
CP
Ci
Σ
(3)
Where Ci is the initial feed concentration of dye (ppm), Cf is
the dye concentration in the permeate (ppm).
The concentration of dyes azoic in the permeate solution
was measured by spectrophotometer, at a wavelength corresponding to the maximum absorbance for the two dyes 455
and 616 nm respectively for AO-74 and MO [1,35,74].
2.4.
Membrane characterization
The SEM was used to assess the morphology of the organic
membrane. Figs. 3 and 4 show images taken by the SEM for
Hydrophobicity of the membrane
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
ARTICLE IN PRESS
J MATER REs TECHNOL.
2 0 2 0;x x x(x x):xxx–xxx
5
Fig. 4 – SEM images, top view of homogeneous membranes of PS/PEI with (c) 15 wt.%and (d) 20 wt.% of PEI.
Fig. 5 – Contact angles of organic membranes M5, M10, M15
and M20 as a function of the percentage by weight of PEI.
permeability. The increase in membrane hydrophilicity greatly
improves filtration performance in terms of permeate flow
as well as the reduction of the fouling phenomenon [61]. To
assess the influence of added PEI on the hydrophilicity of
the PS membrane, the contact angle of the prepared membranes was measured. Fig. 5 illustrates the results of the
measurement of the contact angle of organic membranes having different PEI contents.
The results show that the membranes prepared have relatively high contact angle values (>76◦), the increase in contact
angle means a decrease in the hydrophilicity of the membrane. This could indicate a decrease in the affinity of water to
wet the surface of the membrane. Moreover, as can be seen in
Fig. 5, the contact angle of the membranes increases slightly
from 76.70–80.80 ◦ with increasing PEI content of 5–20%.wt of
PEI.
This slight variation in the contact angle may be due to the
fact of the hydrophobic additive incorporated in the matrix
constituting the organic membranes. The same behavior has
been reported by Pagidi et al. [60]. The most striking result
which emerges from the measurement of the contact angle
is that the incorporation of PEI did not have much influence
on the reduction of the hydrophilicity of the organic membrane.
Fig. 6 – Distilled water flow from M5, M10, M15 and M20 as a
function of bP.
3.3.
Ultrafiltration study
3.3.1.
WATER PERMEABILITY
To assess the permeability of organic membranes (M5, M10,
M15 and M20) synthesized, we performed this test with distilled water at room temperature. The different membranes
were developed in our laboratory; they are described through
binary flow diagrams as a function of the transmembrane
pressure (Fig. 6).
The results obtained in Fig. 6 show a linear dependence
of the flow as a function of the pressure, in accordance with
Darcy’s law, [67]. After 2 h of filtration, the flow of distilled
water was quantified using equation next:
LP = J/OP, with : LPenL/h • m2•bar
These results highlight significant differences observed in
the permeability as a function of the decrease in the weight
percentage of the PEI in the matrix constituting the membranes.
We observed that the permeation through the membranes
M5, M10, M15 and M20 increased significantly from LM20 = 22.66
L m−2 h−1 bar to LM5 = 29.90 L m−2 h−1 bar. These values were
calculated from the slope of the line graphs. This discovery
can be explained by the thickening of the membrane with the
decrease in percentage by weight of PEI compared to PS.
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
6
ARTICLE IN PRESS
J MATER REs TECHNOL.
Fig. 7 – Variation of the flow of AO-74 through the M5, M10,
M15 and M20 as a function of time.
Fig. 8 – Variation of the OM flux through the M5, M10, M15
and M20 as a function of time.
As far as PS membrane concerned is alone has a low flux
value (water permeability) of the order of 16.7 L/h m2 bar [77].
So, in our membrane synthesized, the addition of PEI promotes
the improvement of the permeability of the membrane from
22.66 to 29.90 L h m2 bar. Because the membrane becomes
denser with the addition of PEI as already confirmed by the
SEM characterization [75]. Besides, the observed variation in
permeability can be mainly explained by the decrease in thickness of membrane layers as well as by the difference in pore
size [78].
3.3.2.
UltrAFILTRATION of AZOIC dye solutions
Filtration experiments are carried out using two solutions of
dyes AO-74 and OM to evaluate the flux of the permeation of
the resulting membranes. Figs. 7 and 8 show the changes in
permeate flow of the two acid dyes AO-74 and OM across the
membranes M5, M10, M15 and M20, throughout the filtration
test as a function of time.
The dye flow AO-74 stabilizes after 100 min through the
membranes and gives values of: M5 = 21.92 L/h m2; M10 = 20.10
L/h m2; M15 = 18.73 L/h m2 and M20 = 10.08 L/h m2. The initial
2 0 2 0;x x x(x x):xxx–xxx
Fig. 9 – Rejection rate of the dye AO-74 by M5, M10, M15 and
M20 as a function of time.
permeation flow dropped significantly from 21.92 to 10.08 L/h
m2 with an increase in weight of PEI from 5 wt.% to 20 wt.%.
The dye flow (OM) stabilizes after 100 min through the
membranes M5, M10, M15 and M20 and gives values of: M5 =
46.26 L/h m2; M10 = 36.13 L/h m2; M15 = 23.60 L/h m2 and M20
= 16.30 L/h m2. The initial permeation flow dropped significantly from 46.26 to 16.30 L/h m2 with an increase in weight
of PEI from 5 wt.% to 20 wt.%.
The effect of the addition of PEI on the performance of
the PS membrane is that the more the PEI added increases,
the more the permeability decreases, because the membrane
becomes more and more dense as a function of the addition of
PEI in addition to PS, as has already been confirmed by the SEM
characterization. In addition, we find that there is a slight continuous decrease in the flux of resulting membranes during
the filtration time. This could be explained by the formation
of a concentration polarization layer resulting from the accumulation of dye aggregates on the surface of the membrane
due to their selectivity [75,79–82]. This layer, in the form of a
gel, plays an additional role of resistance to the permeation
of the dye molecules [74,82]. Also, it should be noted that the
reduction in flux and the increase in rejection have been quite
often reported by different authors [41,83–85].
In solution, many dyes self-combine to form molecular
aggregates whose size and structure vary according to the
concentration. Therefore, these results could also be due to
electrostatic interactions between membrane materials and
dye aggregates as well as between the materials themselves.
It has been recognized that electrostatic interaction is a
hydrophobic/hydrophilic interaction between the membrane
and the dye [1,34,67]. This interaction has a significant impact
on the fouling of the membrane by the adsorption of dyes on
the surface of the membrane as well as inside pores. This is
particularly true for the phenomenon of membrane clogging,
which causes the flow of the membrane to decrease [86,87].
The rejection rate of the two-colored solutions AO-74
and MO after the ultrafiltration process is illustrated in
Figs. 9 and 10. The resulting membranes retained more than
65% of the dye for a transmembrane pressure of 2 bar. The
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
ARTICLE IN PRESS
J MATER REs TECHNOL.
Fig. 10 – Rejection rate of the OM dye by M5, M10, M15 and
M20 as a function of time.
rejection rate of the AO-74 dye remained relatively constant
over time for the four membranes (Fig. 9). It reaches values
of 60.7%; 63.5%; 65.6% and 68.5% respectively for the membranes M5, M10, M15 and M20. These results, which do not
reach the maximum values of 90%, may be linked to the formation of a layer of dye aggregates retained on the surface
of the membranes [80]. Ultrafiltration of OM by organic membranes revealed two phases (Fig. 10): the retention rate of the
OM dye gradually increases during the 1st phase between 20
min and 100 min for the various membranes. Then during the
second phase, the rate remains constant. The maximum values of the retention rate obtained are 60.5%; 62.75%; 63.75%
and 64.7% respectively for the membranes M5, M10, M15 and
M20.
We observe that the percentage by weight of PEI (5%, 10%,
15% and 20%) plays in favor of improving the retention rate of
the membranes: RM5 < RM10 < RM15 < RM20. It results from our
experiences that the M20 membrane offers the highest retention power compared to the other membranes for the two dyes
with retention rate values:
RM20 (AO-74) = 68.5% and RM20 (OM) = 64.7% for a duration
of 2 h of filtration against those of MH5: RM5 (AO-74) = 60.7%
and RM5 (OM) = 60.5%. It can be noted that the addition of PEI
remarkably improves the selectivity of the PS/PEI, UF membrane. It could reach a maximum value of 68.5% and 64.7%
respectively for AO-74 and MO using the optimized membrane
of 20 wt.% of PEI. This can be explained by the large size of the
AO-74 molecule compared to that of MO [32].
This difference in selectivity could also be explained by
the spatial geometry of the dyes studied, which could suggest that the MO molecules pass easily through the membrane
in comparison with the AO-74 molecules. Moreover. It is confirmed that in the literature, the surface of the PSF membrane
is negative [61]. We also note a difference in the retention rate
between the two dyes for the same percentage by weight of PEI:
RM15 (AO-74) = 65.6% and RM15 (OM) = 63.75%. The dye rejection could reach a maximum value of R M20 (AO-74) = 68.5% and
RM20 (OM) = 64.7% respectively for AO-74 and OM. It should be
noted that Benkhaya et al. observed the same behavior [75].
2 0 2 0;x x x(x x):xxx–xxx
7
The retention rate of the membranes reaches a limit
beyond which the passage of water becomes difficult because
the aggregates of the dye penetrate inside the pores and adsorb
on the surface of the membranes, which causes the clogging
of the pores at the filtration test courses. This phenomenon
contributes to the reduction in the size of the pores of the
membrane as well as to the formation of a polarization phenomenon [35]. It should be mentioned that Bouazizi et al.
reported the same phenomenon [80]. In addition, this phenomenon participates in the modification of the coefficient
of permeability of pure water and of the retention of the dye
[88,89].
The surface of the membrane is negatively charged and
therefore promotes repulsive interactions between the membrane surface and the anionic dye [75]. Consequently, it could
be concluded that the surface charge could play an essential
role in improving the selectivity of the organic membrane. It
should be noted that Saja et al. observed the same behavior
[67]. The observed decline in flux could be explained primarily
by the increase in osmotic pressure which is directly linked
to the concentrations of dyes [67,90]. In addition, the absorption of pores from the membrane causes the phenomenon
of clogging, which also promotes a decrease in flow [78]. It is
observed for all membranes that the rejection increases continuously over the filtration time, this could also be explained
by the formation of a thin layer of dyes next to the membrane.
This layer plays a role of a second permanent membrane layer
[1,79]. The decrease in the retention flux can also be explained
by the difference in the molecular size of the dyes used relative to the sizes of the pores of the membrane, which causes
high dye retention. This result can also be linked to the formation of a polarization layer due to the electrostatic interactions
between the aromatic nuclei of the membrane and the adjacent dye molecules [35,91]. These interactions result in very
high retention, which cannot be explained by a pure sieving
mechanism [80].
4.
Conclusion
PS/PEI blends were effectively used for the preparation of lowcost organic UF membrane via the phase inversion method.
This preparation revealed the positive influence of the wt.% of
PEI from 5 to 20 wt.% on the properties of the membrane. The
effect of PE content on membrane characteristics was studied
in the range of 5–20 wt.%, and it is found that optimized membrane (M20) was obtained using bentonite content of 20 wt.%.
Therefore, PEI is a new type of additive that can be extended
to the manufacture of other distinguished membranes for the
ultrafiltration of soluble dyes. According to SEM examination,
the morphology of membranes is uniform and does not exhibit
any defects. Further characterizations showed that prepared
membrane has water permeability of 30 L/h m2 bar and high
with dye rejection as a function of the decrease in wt.% of PEI.
Rejection rate of RM20 (AO-74) = 68.5% and RM20 (OM) = 64.7%
respectively.
Finally, filtration results are important and indicate that
prepared PS/PEI UF membrane might be useful for treatment
of textile industry effluents and other harmful particles in
wastewater, etc.
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
8
ARTICLE IN PRESS
J MATER REs TECHNOL.
2 0 2 0;x x x(x x):xxx–xxx
Conflicts of interest
The author declares no conflicts of interest.
[15]
Appendix A. Supplementary data
Supplementary material related to this article can be
found, in the online version, at doi:https://doi.org/10.1016/
j.jmrt.2020.02.102.
[16]
[17]
REFERENCEs
[1] Benkhaya S, Achiou B, Ouammou M, Bennazha J, Younssi SA,
M’rabet S, et al. Preparation of low-cost composite
membrane made of polysulfone/polyetherimide
ultrafiltration layer and ceramic pozzolan support for dyes
removal. Mater Today Commun 2019;19:212–9.
[2] Minas F, Chandravanshi BS, Leta S. Chemical precipitation
method for chromium removal and its recovery from
tannery wastewater in Ethiopia. Chem Int 2017;3:392–405.
[3] Hassen EB, Asmare AM. Predictive performance modeling of
Habesha brewery wastewater treatment plant using artificial
neural network; 2019.
[4] Fu Y, Xu P, Huang D, Zeng G, Lai C, Qin L, et al. Au
nanoparticles decorated on activated coke via a facile
preparation for efficient catalytic reduction of nitrophenols
and azo dyes. Appl Surf Sci 2019;473:578–88.
[5] Fu Y, Qin L, Huang D, Zeng G, Lai C, Li B, et al. Chitosan
functionalized activated coke for Au nanoparticles
anchoring: green synthesis and catalytic activities in
hydrogenation of nitrophenols and azo dyes. Appl Catal B
Environ 2019;255:117740.
[6] Khan R, Patel V, Khan Z. Chapter 5 — bioremediation of dyes
from textile and dye manufacturing industry effluent. In:
Singh P, Kumar A, Borthakur A, editors. Abat. Environmental
Pollution. Elsevier; 2020. p. 107–25,
http://dx.doi.org/10.1016/B978-0-12-818095-2.00005-9.
[7] Khouni I, Louhichi G, Ghrabi A. Assessing the performances
of an aerobic membrane bioreactor for textile wastewater
treatment: influence of dye mass loading rate and biomass
concentration. Process Saf Environ Prot 2020;135:364–82,
http://dx.doi.org/10.1016/j.psep.2020.01.011.
[8] Shindy HA. Problems and solutions in colors, dyes and
pigments chemistry: a review. Chem Int 2017;3:97–105.
[9] Iqbal M, Abbas M, Nisar J, Nazir A, Qamar A. Bioassays based
on higher plants as excellent dosimeters for ecotoxicity
monitoring: a review. Chem Int 2019;5:1–80.
[10] Shindy HA. Basics in colors, dyes and pigments chemistry: a
review. Chem Int 2016;2:29–36.
[11] Lellis B, Fávaro-Polonio CZ, Pamphile JA, Polonio JC. Effects of
textile dyes on health and the environment and
bioremediation potential of living organisms. Biotechnol Res
Innov 2019;3:275–90,
http://dx.doi.org/10.1016/j.biori.2019.09.001.
[12] Berradi M, Hsissou R, Khudhair M, Assouag M, Cherkaoui O,
El Bachiri A, et al. Textile finishing dyes and their impact on
aquatic environs. Heliyon 2019;5:e02711.
[13] Iqbal M. Vicia faba bioassay for environmental toxicity
monitoring: a review. Chemosphere 2016;144:785–802,
http://dx.doi.org/10.1016/j.chemosphere.2015.09.048.
[14] Baig U, Uddin MK, Gondal MA. Removal of hazardous azo dye
from water using synthetic nano adsorbent: facile synthesis,
characterization, adsorption, regeneration and design of
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
experiments. Colloids Surf Physicochem Eng Asp
2020;584:124031,
http://dx.doi.org/10.1016/j.colsurfa.2019.124031.
Rai HS, Bhattacharyya MS, Singh J, Bansal TK, Vats P,
Banerjee UC. Removal of dyes from the effluent of textile and
dyestuff manufacturing industry: a review of emerging
techniques with reference to biological treatment. Crit Rev
Environ Sci Technol 2005;35:219–38.
Szyguła A, Guibal E, Palacín MA, Ruiz M, Sastre AM. Removal
of an anionic dye (Acid Blue 92) by coagulation–flocculation
using chitosan. J Environ Manage 2009;90:2979–86.
Torres NH, Souza BS, Ferreira LFR, Lima ÁS, dos Santos GN,
Cavalcanti EB. Real textile effluents treatment using
coagulation/flocculation followed by electrochemical
oxidation process and ecotoxicological assessment.
Chemosphere 2019;236:124309,
http://dx.doi.org/10.1016/j.chemosphere.2019.07.040.
Oktem YA, Yuzer B, Aydin MI, Okten HE, Meric S, Selcuk H.
Chloride or sulfate? Consequences for ozonation of textile
wastewater. J Environ Manage 2019;247:749–55,
http://dx.doi.org/10.1016/j.jenvman.2019.06.114.
Wang W-L, Hu H-Y, Liu X, Shi H-X, Zhou T-H, Wang C, et al.
Combination of catalytic ozonation by regenerated granular
activated carbon (rGAC) and biological activated carbon in
the advanced treatment of textile wastewater for
reclamation. Chemosphere 2019;231:369–77,
http://dx.doi.org/10.1016/j.chemosphere.2019.05.175.
Wang C, Yediler A, Lienert D, Wang Z, Kettrup A. Ozonation
of an azo dye CI Remazol Black 5 and toxicological
assessment of its oxidation products. Chemosphere
2003;52:1225–32.
Asgari G, Faradmal J, Nasab HZ, Ehsani H. Catalytic ozonation
of industrial textile wastewater using modified C-doped MgO
eggshell membrane powder. Adv Powder Technol
2019;30:1297–311, http://dx.doi.org/10.1016/j.apt.2019.04.003.
Arslan I, Balcioǧlu IA, Bahnemann DW. Advanced chemical
oxidation of reactive dyes in simulated dyehouse effluents
by ferrioxalate-Fenton/UV-A and TiO2/UV-A processes. Dyes
Pigm 2000;47:207–18.
Daij KB, Bellebia S, Bengharez Z. Comparative experimental
study on the COD removal in aqueous solution of pesticides
by the electrocoagulation process using monopolar iron
electrodes. Chem Int 2017;3:420–7.
Jamal MA, Muneer M, Iqbal M. Photo-degradation of
monoazo dye blue 13 using advanced oxidation process.
Chem Int 2015;1:2–6.
Yi H, Jiang M, Huang D, Zeng G, Lai C, Qin L, et al. Advanced
photocatalytic Fenton-like process over biomimetic
hemin-Bi2WO6 with enhanced pH. J Taiwan Inst Chem Eng
2018;93:184–92.
Int C. The use of insoluble matter of Moroccan oil shale for
removal of dyes from aqueous solution; 2019.
Chham A, Khouya E, Oumam M, Abourriche AK, Gmouh S,
Mansouri S, et al. The use of insoluble mater of Moroccan oil
shale for removal of dyes from aqueous solution. Chem Int
2018;4:67–77.
Laissaoui M, Elbatal Y, Vioque I, Manjon G. Adsorption of
methylene blue on bituminous schists from
Tarfaya-Boujdour. Chem Int 2017;3:442–51.
Djehaf K, Bouyakoub AZ, Ouhib R, Benmansour H, Bentouaf
A, Mahdad A, et al. Textile wastewater in Tlemcen (Western
Algeria): impact, treatment by combined process. Chem Int
2017;3:414–9.
Qin L, Zeng Z, Zeng G, Lai C, Duan A, Xiao R, et al.
Cooperative catalytic performance of bimetallic Ni-Au
nanocatalyst for highly efficient hydrogenation of
nitroaromatics and corresponding mechanism insight. Appl
Catal B Environ 2019;259:118035.
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
ARTICLE IN PRESS
J MATER REs TECHNOL.
[31] Abbas M, Adil M, Ehtisham-ul-Haque S, Munir B, Yameen M,
Ghaffar A, et al. Vibrio fischeri bioluminescence inhibition
assay for ecotoxicity assessment: a review. Sci Total Environ
2018;626:1295–309,
http://dx.doi.org/10.1016/j.scitotenv.2018.01.066.
[32] Fradj AB, Boubakri A, Hafiane A, Hamouda SB. Removal of
azoic dyes from aqueous solutions by chitosan enhanced
ultrafiltration. Results Chem 2020;2:100017.
[33] Benkhaya S, M’rabet S, El Harfi A. Classifications, properties,
recent synthesis and applications of azo dyes. Heliyon
2020;6:e03271,
http://dx.doi.org/10.1016/j.heliyon.2020.e03271.
[34] Breida M, Younssi SA, Bouazizi A, Achiou B, Ouammou M, El
Rhazi M. Nitrate removal from aqueous solutions by μ-Al2O3
ultrafiltration membranes. Heliyon 2018;4:e00498.
[35] Karim A, Achiou B, Bouazizi A, Aaddane A, Ouammou M,
Bouziane M, et al. Development of reduced graphene oxide
membrane on flat Moroccan ceramic pozzolan support.
Application for soluble dyes removal. J Environ Chem Eng
2018.
[36] Fonseca Couto C, Lange LC, Santos Amaral MC. A critical
review on membrane separation processes applied to
remove pharmaceutically active compounds from water and
wastewater. J Water Process Eng 2018;26:156–75,
http://dx.doi.org/10.1016/j.jwpe.2018.10.010.
[37] Castro-Muñoz R, Boczkaj G, Gontarek E, Cassano A, Fíla V.
Membrane technologies assisting plant-based and agro-food
by-products processing: a comprehensive review. Trends
Food Sci Technol 2020;95:219–32,
http://dx.doi.org/10.1016/j.tifs.2019.12.003.
[38] Conidi C, Drioli E, Cassano A. Membrane-based agro-food
production processes for polyphenol separation, purification
and concentration. Curr Opin Food Sci 2018;23:149–64,
http://dx.doi.org/10.1016/j.cofs.2017.10.009.
[39] Li X, Shan H, Zhang W, Li B. 3D printed robust
superhydrophilic and underwater superoleophobic
composite membrane for high efficient oil/water separation.
Sep Purif Technol 2020;237:116324,
http://dx.doi.org/10.1016/j.seppur.2019.116324.
[40] Achiou B, Elomari H, Bouazizi A, Karim A, Ouammou M,
Albizane A, et al. Manufacturing of tubular ceramic
microfiltration membrane based on natural pozzolan for
pretreatment of seawater desalination. Desalination
2017;419:181–7.
[41] Bouazizi A, Breida M, Karim A, Achiou B, Ouammou M, Calvo
JI, et al. Development of a new TiO2 ultrafiltration membrane
on flat ceramic support made from natural bentonite and
micronized phosphate and applied for dye removal. Ceram
Int 2017;43:1479–87.
[42] Ang EYM, Toh W, Yeo J, Lin R, Liu Z, Geethalakshmi KR, et al.
A review on low dimensional carbon desalination and gas
separation membrane designs. J Membr Sci 2020;598:117785,
http://dx.doi.org/10.1016/j.memsci.2019.117785.
[43] Saleem H, Trabzon L, Kilic A, Zaidi SJ. Recent advances in
nanofibrous membranes: production and applications in
water treatment and desalination. Desalination
2020;478:114178,
http://dx.doi.org/10.1016/j.desal.2019.114178.
[44] Scholes CA. Hydrogen cyanide recovery by membrane gas
separation. Chem Eng J 2020;386:124049,
http://dx.doi.org/10.1016/j.cej.2020.124049.
[45] Guo Y-S, Ji Y-L, Wu B, Wang N-X, Yin M-J, An Q-F, et al.
High-flux zwitterionic nanofiltration membrane constructed
by in-situ introduction method for monovalent
salt/antibiotics separation. J Membr Sci 2020;593:117441,
http://dx.doi.org/10.1016/j.memsci.2019.117441.
[46] Oussama N, Bouabdesselam H, Ghaffour N, Abdelkader L.
Characterization of seawater reverse osmosis fouled
2 0 2 0;x x x(x x):xxx–xxx
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
9
membranes from large scale commercial desalination plant.
Chem Int 2019;5:158–67.
Benkhaya S, El Harfi A. Synthèse d’une nouvelle membrane
asymétrique et semi préalable à base d’un mélange du
polysulfone (PSU) et polyétherimide (PEI): application à
l’élimination des solutions colorées. [Synthesis of a new
asymmetric membrane and semi prior based on a mixture of
polysulfone (PSU) and polyether imide (PEI): application to
the removal of coloring solution]. Int J Innov Appl Stud
2015;10:285.
Benkhaya S, Cherkaoui O, Assouag CM, Mrabetd S, Rafik M,
Harfi A. Synthesis of a new asymmetric composite
membrane with Bi-component collodion: application in the
ultra filtration of baths of reagent dyes of fabric
rinsing/padding. n.d.
Zanoune K. Synthèse, caractérisation et mise en ø euvre de
membranes composites à base d’oxyde de titane. PhD
Thesis. Abdallah: OUAGUED; 2011.
Benkhaya S, El Harfi A. Design of a new layered composite
membrane based on a physical copolymer (PSU/PEI/PPc)
supported by a Tri-component support (PA6/Fiberglass/PA6).
Application of ultrafiltration baths based on azoic and
antraquinonique dyes. Moroc J Chem 2017;5:1–15.
Qabaqous O, Tijani N, Bennani MN, El Krouk A. Elaboration et
caractérisation des supports plans à base d’argile (Rhassoul)
pour membranes minérales (Preparation and
characterization of supports plans from the (Rhassoul) Clay
for mineral membranes) n.d.
Iqbal DN, Tariq M, Khan SM, Gull N, Iqbal SS, Aziz A, et al.
Synthesis and characterization of chitosan and guar gum
based ternary blends with polyvinyl alcohol. Int J Biol
Macromol 2020;143:546–54.
Swier S, Shaw MT, Weiss RA. Morphology control of
sulfonated poly (ether ketone ketone) poly (ether imide)
blends and their use in proton-exchange membranes. J
Membr Sci 2006;270:22–31.
Min S, Kim D. SAXS cluster structure and properties of
sPEEK/PEI composite membranes for DMFC applications.
Solid State Ion 2010;180:1690–3.
Mikhailenko SD, Zaidi SMJ, Kaliaguine S. Electrical properties
of sulfonated polyether ether ketone/polyetherimide blend
membranes doped with inorganic acids. J Polym Sci B
2000;38:1386–95.
Liu S, Wang L, Ding Y, Liu B, Han X, Song Y. Novel sulfonated
poly (ether ether keton)/polyetherimide acid-base blend
membranes for vanadium redox flow battery applications.
Electrochim Acta 2014;130:90–6.
Kaleekkal NJ, Thanigaivelan A, Rana D, Mohan D. Studies on
carboxylated graphene oxide incorporated polyetherimide
mixed matrix ultrafiltration membranes. Mater Chem Phys
2017;186:146–58.
Li X-M, Ji Y, Yin Y, Zhang Y-Y, Wang Y, He T. Origin of
delamination/adhesion in polyetherimide/polysulfone
co-cast membranes. J Membr Sci 2010;352:173–9.
Govardhan B, Chandrasekhar SS, Sridhar S. Purification of
surface water using novel hollow fiber membranes prepared
from polyetherimide/polyethersulfone blends. J Environ
Chem Eng 2017;5:1068–78.
Pagidi A, Saranya R, Arthanareeswaran G, Ismail AF,
Matsuura T. Enhanced oil–water separation using
polysulfone membranes modified with polymeric additives.
Desalination 2014;344:280–8.
Ng HM, Leo CP, Abdullah AZ. Selective removal of dyes by
molecular imprinted TiO2 nanoparticles in polysulfone
ultrafiltration membrane. J Environ Chem Eng 2017;5:3991–8.
Dos Santos AB, Cervantes FJ, van Lier JB. Review paper on
current technologies for decolourisation of textile
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
JMRTEC-1467; No. of Pages 10
10
ARTICLE IN PRESS
J MATER REs TECHNOL.
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
wastewaters: perspectives for anaerobic biotechnology.
Bioresour Technol 2007;98:2369–85.
Mutamim NSA, Noor ZZ, Hassan MAA, Olsson G. Application
of membrane bioreactor technology in treating high strength
industrial wastewater: a performance review. Desalination
2012;305:1–11.
Madaeni SS, Samieirad S. Chemical cleaning of reverse
osmosis membrane fouled by wastewater. Desalination
2010;257:80–6.
Reemtsma T, Zywicki B, Stueber M, Kloepfer A, Jekel M.
Removal of sulfur-organic polar micropollutants in a
membrane bioreactor treating industrial wastewater.
Environ Sci Technol 2002;36:1102–6.
Xu Z, Liao J, Tang H, Li N. Antifouling polysulfone
ultrafiltration membranes with pendent sulfonamide
groups. J Membr Sci 2018;548:481–9.
Saja S, Bouazizi A, Achiou B, Ouaddari H, Karim A, Ouammou
M, et al. Fabrication of low-cost ceramic ultrafiltration
membrane made from bentonite clay and its application for
soluble dyes removal. J Eur Ceram Soc 2020;40:2453–62,
http://dx.doi.org/10.1016/j.jeurceramsoc.2020.01.057.
Tabernero A, Cardea S, Masa R, Baldino L, del Valle EMM,
Reverchon E. Preparation and characterization of cellulose
acetate-Laponite® composite membranes produced by
supercritical phase inversion. J Supercrit Fluids
2020;155:104651,
http://dx.doi.org/10.1016/j.supflu.2019.104651.
Yu L, Kanezashi M, Nagasawa H, Tsuru T. Phase
inversion/sintering-induced porous ceramic microsheet
membranes for high-quality separation of oily wastewater. J
Membr Sci 2020;595:117477,
http://dx.doi.org/10.1016/j.memsci.2019.117477.
Durmaz EN, Zeynep Çulfaz-Emecen P. Cellulose-based
membranes via phase inversion using [EMIM]OAc-DMSO
mixtures as solvent. Chem Eng Sci 2018;178:93–103,
http://dx.doi.org/10.1016/j.ces.2017.12.020.
Omonov TS, Harrats C, Moldenaers P, Groeninckx G. Phase
continuity detection and phase inversion phenomena in
immiscible polypropylene/polystyrene blends with different
viscosity ratios. Polymer 2007;48:5917–27.
Harrats C, Groeninckx G, Thomas S. Micro-and
nanostructured multiphase polymer blend systems: phase
morphology and interfaces. CRC Press; 2005.
Tiron LG, Pintilie ŞC, Vlad M, Birsan IG, Baltă Ş.
Characterization of polysulfone membranes prepared with
thermally induced phase separation technique. IOP
Conference Series: Materials Science and Engineering, 209.
IOP Publishing; 2017. p. 12013.
Bouazizi A, Breida M, Achiou B, Ouammou M, Calvo JI,
Aaddane A, et al. Removal of dyes by a new nano–TiO2
ultrafiltration membrane deposited on low-cost support
prepared from natural Moroccan bentonite. Appl Clay Sci
2017;149:127–35.
Benkhaya S, Achiou B, Ouammou M, Bennazha J, Alami
Younssi S, M’rabet S. Preparation of low-cost composite
membrane made of polysulfone/polyetherimide
ultrafiltration layer and ceramic pozzolan support for dyes
removal. Mater Today Commun 2019;19:212–9,
http://dx.doi.org/10.1016/j.mtcomm.2019.02.002.
Purkait MK, Sinha MK, Mondal P, Singh R. Chapter 1 —
introduction to membranes. In: Purkait MK, Sinha MK,
Mondal P, Singh R, editors. Interface Science and Technology,
2 0 2 0;x x x(x x):xxx–xxx
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
[90]
[91]
25. Elsevier; 2018. p. 1–37,
http://dx.doi.org/10.1016/B978-0-12-813961-5.00001-2.
Arthanareeswaran G, Mohan D, Raajenthiren M. Preparation
and performance of polysulfone-sulfonated poly (ether ether
ketone) blend ultrafiltration membranes. Part I. Appl Surf Sci
2007;253:8705–12.
Ouaddari H, Karim A, Achiou B, Saja S, Aaddane A, Bennazha
J, et al. New low-cost ultrafiltration membrane made from
purified natural clays for direct Red 80 dye removal. J Environ
Chem Eng 2019;7:103268.
Achiou B, Beqqour D, Elomari H, Bouazizi A, Ouammou M,
Bouhria M, et al. Preparation of inexpensive NaA zeolite
membrane on pozzolan support at low temperature for
dehydration of alcohol solutions. J Environ Chem Eng
2018;6:4429–37.
Bouazizi A, Breida M, Karim A, Achiou B, Ouammou M, Calvo
JI, et al. Development of a new TiO2 ultrafiltration membrane
on flat ceramic support made from natural bentonite and
micronized phosphate and applied for dye removal. Ceram
Int 2017;43:1479–87.
Petrinić I, Andersen NPR, Šostar-Turk S, Le Marechal AM. The
removal of reactive dye printing compounds using
nanofiltration. Dyes Pigm 2007;74:512–8.
Yu S, Chen Z, Cheng Q, Lü Z, Liu M, Gao C. Application of
thin-film composite hollow fiber membrane to submerged
nanofiltration of anionic dye aqueous solutions. Sep Purif
Technol 2012;88:121–9.
Saffaj N, Loukili H, Younssi SA, Albizane A, Bouhria M, Persin
M, et al. Filtration of solution containing heavy metals and
dyes by means of ultrafiltration membranes deposited on
support made of Moroccan clay. Desalination 2004;168:301–6.
Saffaj N, Persin M, Younssi SA, Albizane A, Bouhria M, Loukili
H, et al. Removal of salts and dyes by low ZnAl2O4–TiO2
ultrafiltration membrane deposited on support made from
raw clay. Sep Purif Technol 2005;47:36–42.
Akbari A, Remigy JC, Aptel P. Treatment of textile dye
effluent using a polyamide-based nanofiltration membrane.
Chem Eng Process Process Intensif 2002;41:601–9.
Chang I-S, Le Clech P, Jefferson B, Judd S. Membrane fouling
in membrane bioreactors for wastewater treatment. J
Environ Eng 2002;128:1018–29.
Fan L, Harris JL, Roddick FA, Booker NA. Influence of the
characteristics of natural organic matter on the fouling of
microfiltration membranes. Water Res 2001;35:4455–63.
Saffaj N, Persin M, Younsi SA, Albizane A, Cretin M, Larbot A.
Elaboration and characterization of microfiltration and
ultrafiltration membranes deposited on raw support
prepared from natural Moroccan clay: application to
filtration of solution containing dyes and salts. Appl Clay Sci
2006;31:110–9.
Lin J, Ye W, Baltaru M-C, Tang YP, Bernstein NJ, Gao P, et al.
Tight ultrafiltration membranes for enhanced separation of
dyes and Na2SO4 during textile wastewater treatment. J
Membr Sci 2016;514:217–28.
Banerjee P, De S. Steady state modeling of concentration
polarization including adsorption during nanofiltration of
dye solution. Sep Purif Technol 2010;71:128–35.
Jiang M, Ye K, Lin J, Zhang X, Ye W, Zhao S, et al. Effective dye
purification using tight ceramic ultrafiltration membrane. J
Membr Sci 2018;566:151–60.
Please cite this article in press as: Said B, et al. Synthesis of new low-cost organic ultrafiltration membrane made from Polysulfone/Polyetherimide
blends and its application for soluble azoic dyes removal. J Mater Res Technol. 2020. https://doi.org/10.1016/j.jmrt.2020.02.102
Download