You are on page 1of 11

Journal of Membrane Science 327 (2009) 125–135

Contents lists available at ScienceDirect

Journal of Membrane Science


journal homepage: www.elsevier.com/locate/memsci

Characteristics, performance and stability of polyethersulfone ultrafiltration


membranes prepared by phase separation method using different
macromolecular additives
Heru Susanto a,b,∗ , Mathias Ulbricht a
a
Lehrstuhl für Technische Chemie II, Universität Duisburg-Essen, 45117 Essen, Germany
b
Department of Chemical Engineering, Universitas Diponegoro, Semarang, Indonesia

a r t i c l e i n f o a b s t r a c t

Article history: Polyethersulfone (PES) ultrafiltration (UF) membranes were prepared by non-solvent-induced phase
Received 30 June 2008 separation (NIPS) method using different macromolecular additives, polyvinylpyrrolidone (PVP),
Received in revised form 5 October 2008 poly(ethylene glycol) (PEG) and poly(ethylene oxide)-b-poly(propylene oxide)-b-poly(ethylene oxide)
Accepted 9 November 2008
(Pluronic® , Plu). Their effects on membrane structure and their stability in the polymer membrane matrix
Available online 24 November 2008
as well as the resulting membrane performance were systematically compared in order to determine the
additive that should be preferred. The investigated membrane characteristics include surface hydrophilic-
Keywords:
ity (by contact angle), surface charge (by zeta potential), surface chemistry (by FTIR spectroscopy), water
Ultrafiltration membrane
Membrane preparation
flux and rejection of macromolecular test substances. Visualization of membrane surface and cross-
Polyethersulfone section morphology was also done by scanning electron microscopy. The membrane performance was
Polyvinylpyrrolidone examined by investigation of adsorptive fouling and ultrafiltration using solution of bovine serum albu-
Poly(ethylene glycol) min as the model system. The stability of additive was examined by incubating the membrane in water
Poly(ethylene glycol)-b-poly(propylene (40 ◦ C) and sodium hypochlorite solution. Modification effects on membrane characteristic as well as per-
glycol)-b-poly(ethylene glycol) formance via blending membrane polymer and macromolecular additive were clearly observed. Overall,
Pluronic® the results suggest that Pluronic showed the best behavior in all respects, and, consequently, it should be
considered in practical applications.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction (PES) UF membranes are broadly manufactured for industrial appli-


cations. Nevertheless, the hydrophobicity of those materials can
Ultrafiltration (UF) has become the main focus as promising cause severe fouling problems; therefore, membrane modification
separation tool in many industrial processes covering fraction- is usually done to increase the membrane resistance towards foul-
ation and concentration steps in the food, in pharmaceutical ing. Three different approaches including (i) membrane polymer
and biotechnological industries, in pure water production and modification (pre-modification), (ii) blending of the membrane
in water and wastewater treatments [1,2]. As consequence of polymer with a modifying agent (additive), and (iii) surface modifi-
this increasing demand, efforts to improve UF process perfor- cation after membrane preparation (post-modification) have been
mance are gaining more and more importance. In general, those proposed [3]. Even though blending technique can involve signif-
efforts include feed pretreatment, advanced membrane and mod- icant changes in composition of the casting solution, leading to
ule design, and process condition optimization. However, in many different membrane structure formed during the phase separa-
cases, the key for the performance of UF process is the mem- tion and, consequently, membrane properties can be quite different
brane itself. In this regard, three important characteristics for from the unmodified reference material, it is simple and no addi-
achieving high performance UF membrane are high flux as well tional step is needed during membrane manufacturing. In addition,
as selectivity, low fouling and performance stability for long-term although stability of the modifying agent in the membrane matrix
operation. can be a problem, the effect of hydrophilization of the polymer
Polymeric membranes prepared by non-solvent-induced phase membrane can clearly be observed. Therefore, this technique seems
separation (NIPS) are still dominating commercially available UF to be of highest relevance from practical point of view. Polymeric
membranes. Among them, polysulfone (PSf) and polyethersulfone additives (usually hydrophilic polymers) in a casting solution are
also used in order to increase both pore size and porosity (pore-
forming agent) and to suppress macrovoid formation. However,
∗ Corresponding author. Tel.: +49 201 1833066; fax: +49 201 1833147. depending on the polymer membrane, solvent and NIPS conditions,
E-mail address: heru.susanto@uni-due.de (H. Susanto). the opposite effect can also be observed.

0376-7388/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2008.11.025
126 H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135

Polyvinylpyrrolidone (PVP) and poly(ethylene glycol) (PEG) and several cases have been reported where high performance
have been intensively used as additives during preparation of membranes could be obtained. However, considerable disagree-
PES UF membranes by phase separation methods. The mecha- ment was also observed among different reports. We believe that
nism of PES membrane formation with addition of PVP had well this is because many parameters including characteristics of mem-
been explained in many publications [4–8]. Addition of PVP into brane polymer (e.g., molar mass, concentration in casting solution),
PES–NMP (N-methyl-2-pyrrolidone) casting solution could sup- solvent (e.g., solubility parameter, viscosity), additive (e.g., molar
press the macrovoid formation via decreasing the effect of delayed mass, concentration), and preparation conditions (e.g., tempera-
demixing [7]. However, opposite effect, i.e., enlargement of the ture, humidity) are involved. Therefore, it is hard to determine the
macrovoid structure by addition of PVP into PES–DMF (dimethyl- best additive that should be used in preparation of PES UF mem-
formamide) was also reported [8]. Increasing water permeability brane based on results of previous studies. Torrestiana-Sanchez
as well as molecular weight cut off (MWCO) of PES membrane et al. [10] had briefly compared the use of PVP and PEG as non-
by the addition of PVP was another observation [9]. Addition of solvents during PES membrane preparation. However, the reported
PVP to PES–NMP led to higher water permeability than without data were very limited to water permeability and protein rejec-
PVP but the solute rejection (for PEG) was 15% lower than with- tion. Other important parameters such as hydrophilicity, fouling
out PVP. Moreover, it had been found that the characteristic and behavior and membrane stability have not yet been compared. In
performance of PES–PVP blend membranes were influenced by the present study, the effects of those macromolecular additives
the concentration and the molar mass of PVP [4,10,11]. The high- on characteristics, performance and stability of the resulting PES
est product permeation rate examined by using PEG solution was UF membranes are systematically compared. All preparation condi-
obtained at PVP/PES weight ratio of unity [4]. Observed retention tions, including composition of membrane casting solution as well
coefficient as well as surface roughness increased with increasing as molar mass of additive were maintained to be similar. Stability
molar mass of PVP [11]. The effect of PVP content on the resulting study was included because the immobilization of hydrophilic addi-
fouling behavior examined with protein was also investigated; the tives in the polymeric membrane matrix is one of the critical issues
presence of PVP could to some extent decrease degree of fouling of from a practical point of view. Thus, the best performing hydrophilic
the resulting membrane [12]. macromolecular additive for preparation PES UF membranes will be
PEG with various molar masses was added during preparation known based on the results of this work.
of PES UF membranes [13]. The performance of the resulting mem-
branes was influenced by both molar mass and concentration of
2. Experimental
PEG. Membranes prepared with higher molar mass of PEG had
higher pure water permeation and larger pores. The water perme-
2.1. Materials
ation increased as concentration of PEG (400 and 600 g/mol) was
increased, while solute separation decreased. Apparently, an opti-
Commercial PES (Ultrason E 6020 P) donated by BASF (Lud-
mum condition (high flux with acceptable solute rejection) was
wigshafen, Germany) was used and dried at 120 ◦ C for at
achieved at PEG concentration of ∼10 wt%. In addition, differences
least 4 h before use. N-methyl-2-pyrrolidone (NMP) was pur-
in surface morphology and roughness could also be detected. A sim-
chased from Merck (Hohenbrunn, Germany). Polyvinylpyrrolidone
ilar study was also performed by using PSf–NMP and PSf–DMAc
(MW ∼10,000 g/mol) was purchased from Serva Feinbiochemica
(dimethylacetamide) system [14]. The results showed that the
GmbH&Co. (Heidelberg, Germany). PEG (MW ∼10,000 g/mol) was
increase in PEG molar mass increased the membrane porosity lead-
purchased from Fluka Chemie AG (Buchs, Germany) and Pluronic
ing to increase in water permeability for both systems. Liu et al. [15]
F127 (Plu) (MW ∼12,600 g/mol) was purchased from BASF (Mount
found an optimum PEG content in PES–NMP system for increas-
Olive, NJ, USA). Bovine serum albumin (BSA) and sodium hypochlo-
ing water permeability, and further increase in PEG content would
rite were purchased from ICN Biomedicals, Inc. (California, USA)
decrease the resulting water flux. Moreover, the addition of PEG
and Sigma–Aldrich Chemie GmbH (Steinheim, Germany), respec-
alone could not suppress macrovoid formation even at high concen-
tively. Dextrans T-4, T-15, T-35, T-100 and T-200 (the numbers
tration. Only when relatively large amounts of water were added to
indicating molar mass in kg/mol) were from Serva Feinbiochemica
the dope solution, sponge-like membrane structure was obtained.
GmbH&Co. (Heidelberg, Germany). Potassium dihydrogen phos-
The effect of PEG/NMP ratio during preparation of PSf UF mem-
phate (KH2 PO4 ) and disodium hydrogen phosphate dihydrate
branes was investigated [16]. As the PEG content was increased
(Na2 HPO4 ·2H2 O) were purchased from Fluka Chemie AG (Buchs,
the pure water flux increased while the solute rejection would
Germany). Potassium chloride (KCl), potassium hydroxide (KOH),
decrease.
and hydrochloric acid (HCl), all of p.a. quality, were purchased from
Recently, the group of Jiang has intensively proposed the use of a
Bernd Kraft GmbH (Duisburg, Germany). Nitrogen gas purchased
triblock copolymer, poly(ethylene oxide)-b-poly(propylene oxide)-
from Messer Griesheim GmbH (Krefeld, Germany) was ultrahigh
b-poly(ethylene oxide) (Pluronic/Plu), as another attractive additive
purity. Water purified with a Milli-Q system from Millipore was
for preparation of PES UF membranes [17–20]. They observed that
used for all experiments.
Pluronic could improve not only the flux but also the resistance
towards fouling. The resulting fouling resistance was significantly
influenced by Pluronic content and PEO chain length [19]. It was 2.2. Membrane preparation
reported that the apparent protein adsorption amount decreased
significantly with increasing the Pluronic content and reached PES (15 wt%) was dissolved in NMP (75%) with stirring, and
∼0 ␮g/cm2 at concentration of 10.5% [17]. By contrast, the protein different additive with similar molar mass (PVP, PEG or amphiphilic
rejection and pure water flux were little influenced by the Pluronic triblock copolymer Plu; 10 wt%) was added to the polymer solu-
content. Very recently, the function of Pluronic as both hydrophilic tion. Attention should be given to PEG because it would not
modifier and pore-forming agent has been confirmed [20]. dissolve in the NMP at ambient temperature. Therefore, slight
In sum, three different macromolecular additives, i.e., PVP, PEG heating (∼45 ◦ C) was used during dissolution. This relative high
and Plu, have frequently been used in the preparation of PES UF concentration of additive was aimed to see significant effects on
membranes including a wide variety in preparation conditions the resulting membrane. Polymer solution without an additive
and polymer characteristics. Those additives determine the char- was also prepared for control experiments. The homogenous
acteristics as well as the performance of the resulting membrane, polymer solution was left without stirring until no bubbles were
H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135 127

observed. The membranes were prepared by using Coatmaster 2.5. Rejection measurement
509 MC, Erichsen Testing Equipment. The polymer solution was
cast with a thickness of 200 m using a steel casting knife on a Rejection tests were conducted with a five component mixture
glass substrate (casting speed 25 mm/s) and subjected to humid of dextrans with molar mass ranging from 4 to 200 kg/mol at a
air (RH = 50–60%) for 1 min. Thereafter, the proto-membrane was total concentration of 1 g/L. Experiments were performed using a
solidified in a coagulation bath containing water (20 ± 1 ◦ C) for 1 h. dead-end stirred filtration system (Amicon cell model 8050, Mil-
The resulting membranes were washed and soaked in the water lipore; cf. Section 2.4) at a pressure of 100 kPa and a stirring rate
for 24 h before drying. Drying was sequentially done by immersing of 300 rpm. Around 10 mL of permeate was collected. The com-
in water, water/ethanol, ethanol and hexane. positions of dextran mixtures in the permeate (Cdownstream ) and
feed/retentate (Cupstream ) sides of membrane were analyzed using
2.3. Shrinkage measurements gel permeation chromatography (GPC). The apparent rejection for
each molar mass was calculated using the following equation:
Membrane shrinkage was calculated by measuring the area
Cdownstream
of the initial film before solidification (“proto-membrane”) and R=1− (4)
Cupstream
the membrane after solidification as well as after drying. These
areas were then compared to calculate degree of shrinkage (“plane
2.6. Membrane morphology
shrinkage”). In addition, membrane shrinkage was also calculated
based on the resulting membrane thickness and initial film thick-
The top surface and cross-section morphology of the mem-
ness (“vertical shrinkage”).
  branes were observed by using a Quanta 400 FEG (FEI)
A1 environmental scanning electron microscope (ESEM) at standard
degree of shrinkage (plane) (%) = 1− × 100 (1)
A0 high-vacuum conditions. A K 550 sputter coater (Emitech, U.K.) was
 l1
 used to coat the outer surface of the sample with gold/palladium.
degree of shrinkage (vertical) (%) = 1− × 100 (2) For cross-section analysis, the membranes were broken in liquid
l0
nitrogen and sputtered for 1.5 min, while for analysis of outer mem-
where A0 is the area of proto-membrane, A1 is the area of membrane brane surface, sputtering was done for 0.5 min.
after solidification or after drying and l0 and l1 are the thickness of
membrane before solidification and after drying, respectively. The
2.7. Contact angle (CA)
membrane thickness was measured by Coolant Proof micrometer
IP 65, Mutico Co., Japan.
Sessile drops static CA was measured using an optical con-
tact angle measurement system (OCA 15 Plus; Dataphysics GmbH,
2.4. Compaction, hydraulic permeability measurement,
Filderstadt, Germany). Five microlitres of water was dropped on the
adsorptive fouling and ultrafiltration procedures
membrane surface from a microsyringe with a stainless steel nee-
dle in room temperature (21 ± 1 ◦ C). At least seven measurements
All experiments were carried out by using a dead-end stirred cell
of drops at different locations were averaged to obtain CA for one
filtration system (Amicon cell model 8010 for compaction, hydraulic
membrane sample.
permeability measurement and adsorptive fouling, model 8050 for
UF experiments) connected to a reservoir (∼450 mL) and pres-
2.8. Zeta potential (ZP)
surized by nitrogen from a gas tank. Membrane compaction was
performed by filtration of pure water at 450 kPa for 2 h. The flux
The membrane surface charge was investigated by an outer
profile over time was monitored online gravimetrically. In all fur-
surface streaming potential measurement. Experiments were car-
ther experiments, the membrane was firstly compacted for at
ried out in a flat-sheet tangential flow module described in detail
least 1 h (this time was enough to achieve steady flux, cf. Section
in previous study [21]. Before measurement, the membrane was
3.2). Hydraulic membrane permeability was measured at differ-
equilibrated by soaking in 0.001 mol/L KCl solution. The stream-
ent transmembrane pressures within the range 100–400 kPa and
ing potentials of membranes were measured using 0.001 mol/L KCl
at least five measurements from different membrane samples were
solution within the pH range 3–10 and at a temperature of 25 + 1 ◦ C.
averaged.
The ZP was calculated using the Helmholtz–Smoluchowski equa-
For static adsorption experiments (adsorptive fouling), a solu-
tion.
tion of BSA (1 g/L, pH 7 in phosphate buffer) was added to the cell
and the outer membrane surface was exposed for 3 h without any
flux at a stirring rate of 3.1416 × 10−1 rad/s (300 rpm). Afterwards, 2.9. Surface chemistry
the solution was removed, and the membrane surface was rinsed
two times by filling the cell with pure water (5 mL) and shaking it The membrane surface chemistry was analyzed by using the
for 30 s. Water fluxes before and after exposure were measured at instrument Varian 3100 Fourier transform infrared spectroscopy
the same pressure (300 kPa). The evaluation of membrane perfor- (FTIR) Excalibur series. A total of 64 scans were performed at a
mance was expressed in terms of the relative flux reduction, RFR resolution of 4 cm−1 and the temperature of 21 + 1 ◦ C. The Varian’s
(Eq. (3)). Resolution Pro 4.0 was used to record the membrane spectra versus
the corresponding background spectra.
J0 − Ja
RFR (%) = × 100 (3)
J0 2.10. Stability test
Ultrafiltration experiments at a constant transmembrane pres-
sure (300 kPa) were conducted using a BSA solution (0.1 g/L, pH 7 in The stability of macromolecular additive in polymer membrane
phosphate buffer) as the feed. The balance was connected to the PC, matrix was examined by incubating membrane samples in water
the weight of permeate was recorded online, and the flux was cal- (20 and 40 ◦ C) and sodium hypochlorite solution (active chlorine
culated. The permeate flux profile over time was investigated. BSA concentration 400 mg/L) for up to 10 days. Water is usually used
concentrations were determined by measuring its UV absorbance for membrane washing before chemical cleaning will be done.
at 280 nm. The apparent BSA rejection was calculated using Eq. (4). Hypochlorite solution had been known as one of the most popular
128 H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135

Fig. 1. “Plane” membrane shrinkage after coagulation and drying for different mem- Fig. 2. Water flux profile over time during compaction at pressure of 450 kPa.
branes. PES(only), PES–PVP, PES–PEG and PES–Plu are membranes prepared without
an additive, with PVP, with PEG and with Pluronic additives, respectively. The error
bars represent standard deviation. solidification may be able to explain [26,27]. It is clearly seen that
different macromolecular additive would result in different poly-
mer characteristics and consequently, polymer contraction would
chemical cleaning agents to remove irreversible fouling [22–24].
also be different. It should also be noted that the total polymer
Surface chemistry (by FTIR), surface hydrophilicity (by CA) and
concentration for PES(only) was significantly smaller than the oth-
water flux measurements were used to investigate changes in mem-
ers and consequently, lower viscosity and faster solvent exchange
brane characteristics and transport property. The experiments were
should be obtained. Difference in miscibility between PES and addi-
done by using a different membrane. In water flux stability study,
tive for different additives used might also contribute. We have
the pure water flux was firstly measured (before soaking) and
tried to investigate the miscibility by using differential scanning
membranes were then soaked in hypochlorite solution in closed
calorimetry (DSC). However, clear explanation could not be drawn
containers for a certain number of days. Thereafter, the membrane
from the resulting data. Residual solvent in the sample might have
was rinsed and the water flux was again measured.
influence on the resulting DSC thermogram. Such phenomena seem
to be found also in previously reported literature [18]. Further anal-
3. Results and discussions ysis of this miscibility is now currently under investigation. Overall,
PES–Plu and PES–PEG membranes showed smaller shrinkage than
3.1. Shrinkage and membrane thickness both PES(only) and PES–PVP membranes.

Understanding of membrane shrinkage will give information


3.2. Membrane compaction, water permeability and membrane
for predicting the area of a non-supported asymmetric membrane
rejection
product but not for a supported asymmetric membrane. High
shrinking tendency will result in tensions within the membrane,
Compaction – a compression of the membrane structure under
and this can have influence on structure of the resulting mem-
a transmembrane pressure difference causing a decrease in mem-
branes. It was reported that shrinkage had effect on membrane
brane permeability due to mechanical deformation of the solid
porosity [25]. Fig. 1 and Table 1 show the plane and the vertical
polymer – is a common phenomenon during application of
shrinkages, respectively.
polymeric membranes [28,29]. In this compaction study, the mem-
As clearly seen in Fig. 1 and Table 1, decreases in membrane
branes were pressurized at high pressure (450 kPa) for 2 h.
area and thickness indicate that the PES film shrank not only in the
Fig. 2 shows an example of the water flux profile over time
plane view but also in the vertical direction. In plane view, PES–PVP
during compaction for all membranes. Indeed, gradual decrease
membrane showed the highest degree of shrinkage (25%) followed
in flux over the duration of compaction time was observed for all
by PES (only). The highest shrinkage of PES–PVP might contribute
membranes, and flux reached a steady value after ∼60 min of com-
to smaller surface porosity as noticed by its smallest water perme-
paction. The steady fluxes were ∼80%, ∼35%, ∼40% and ∼60% of the
ability among the PES membranes prepared with an additive (cf.
initial flux for PES(only), PES–PVP, PES–PEG and PES–Plu, respec-
Section 3.2). Although the membranes have been dried by exchange
tively. Interestingly, the membrane prepared by using Pluronic
with organic non-solvent (for PES) method, shrinkage after drying
as the additive had initially lower flux than the membrane pre-
could still be observed. Different with plane view, all membranes
pared using PEG as the additive, but beyond 30 min duration of
prepared using an additive showed similar degree of shrinkage in
compaction it showed higher flux. Because all membranes had
vertical view. Even though the reason behind this shrinkage phe-
asymmetric structure (cf. SEM images), the compaction at high
nomenon is still not clearly understood, polymer contraction upon
pressure would cause densification of the more porous support
layer leading to a thickening of the skin layer (selective barrier).
Table 1 Consequently, thicker membrane would result in lower flux. Such
Membrane thickness and the resulting “vertical” shrinkage after drying (polymer
phenomenon had been observed in the previous reported literature
solution has been cast with 200 ␮m thickness).
[28,30].
No. Membrane Thickness (␮m) Degree of shrinkage (%) Table 2 shows the hydraulic membrane permeability measured
1 PES(only) 98 ± 4 51 after 1 h of compaction. The hydraulic permeability of the wet
2 PES–PVP 130 ± 4 35 membranes (never dried) is also included. The membrane prepared
3 PES–PEG 135 ± 5 32 without an additive showed the highest hydraulic permeability
4 PES–Plu 143 ± 6 29
for both conditions. It is seen that drying process significantly
H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135 129

Table 2 chosen) showed that most of the results showed similar trends
Hydraulic membrane permeability measured after 1 h compaction.
with Fig. 3 and the standard deviations of MWCO were within the
No. Membrane Lp (L/m2 h kPa) after drying Lp (L/m2 h kPa) before drying range 15–30%. The possible reason for this observation would be
1 PES 0.251 ± 0.16 1.05 ± 0.09
that even though the PES(only) membrane had smaller MWCO it
2 PES–PVP 0.066 ± 0.02 1.320 ± 0.06 had a higher pore density than the membranes prepared with an
3 PRS–PEG 0.101 ± 0.02 1.421 ± 0.03 additive.
4 PES–Plu 0.131 ± 0.04 1.178 ± 0.05 Comparing the membranes prepared with an additive, it is
clearly shown that all membranes showed similar rejection curve
decreased the hydraulic permeability of all membranes. The flux trend. The hydraulic membrane permeability, rejection and perme-
reductions after solvent exchange drying were ∼76%, ∼95%,∼93% ate flux data suggest that the PES–Plu membrane had either the
and ∼88% for PES(only), PES–PVP, PES–PEG and PES–Plu mem- highest pore density or largest pore size leading to the highest sur-
branes, respectively (drying without solvent exchange resulted in face porosity among the membranes prepared with an additive.
much greater flux reductions). The absence of impregnation may Comparing PES–PVP and PES–PEG, it is shown that PES–PEG had
contribute to this very large decrease in flux after drying but it higher hydraulic permeability. This observation agrees with result
should not be the only reason. Nevertheless, the effect of drying reported by Torrestiana-Samschez et al. [10].
on membrane performance was beyond the scope of this paper.
Because all relative flux reductions showed similar trend; all fur- 3.3. Membrane morphology
ther experiments were performed and discussed by using dried
membranes. Membrane surface morphology as well as cross-section struc-
In general, addition of a hydrophilic polymer can facilitate liquid ture was visualized by using SEM. It is important to mention that
demixing because the system will be closer to phase separation on the SEM observation for membrane surface was repeated for every
the one hand, and slow down phase separation by hindering non- membrane and similar image was obtained for most of samples. As
solvent inflow to the polymer–solvent mixture (delayed demixing) presented in Fig. 4, all membranes had asymmetric structure con-
due to higher viscosity on the other hand. By contrast, the poly- sisting of a thin fine porous selective barrier and a much thicker
mer solution without an additive would facilitate faster solvent porous sub-structure. The phenomena behind the formation of this
exchange due to much lower viscosity. It was reported that addition typical structure had been explained in many previous publications
of 10% of PVP or PEG into 20% of PES in NMP solution increased the [6,7]. Cross-section images support the results of membrane thick-
casting solution viscosity up to more 200% (from 0.70 to 2.75 Pa s) ness measurement (cf. Section 3.1), except for PES–PVP membrane.
[10]. In this work, the hydraulic permeability data of wet mem- It seemed that part of PES–PVP membrane disappeared during sam-
branes showed that addition of a hydrophilic polymer increased the ple preparation. Visualization of surface morphology showed that
resulting water flux suggesting that the effect of delayed demixing the membrane surface had fine pore structure with dimensions in
was less significant compared to the effect of instantaneous liquid the nanometer range (<10 nm). It is clearly seen that PES–PVP had
demixing. the roughest surface morphology (it should be noted that the drying
In order to quantify the pore size and its distribution of selective process has been done by non-solvent exchange method, thereby
barrier, the rejection of polydisperse macromolecular test sub- collapse of the pore structure should be minimized). Quantification
stances (dextran) was then measured. Fig. 3 shows that the addition of surface roughness performed by Ochoa et al. [11] showed that
of a macromolecular modifier into polymer solution changed the the addition of PVP in preparation of PES UF membranes increased
rejection curve of the membrane. As shown in Fig. 3, the mem- the surface roughness. Comparing surface images of PES(only) and
brane prepared without an additive had a steeper separation curve PES–Plu membranes, it appears that the PES–Plu membrane had
implying that the pore size distribution is narrower than for the larger pore size than PES(only). Nevertheless, it seems that the pore
membranes prepared with an additive. In addition, it had smaller density of PES(only) membrane is higher than that of the PES–Plu
MWCO but showed the greatest permeate flux during rejection membrane. Thus, smaller pore size but higher pore density for
experiment (data not shown) and water flux (cf. above). It is impor- PES(only) compared to PES–Plu supports the previous discussion
tant to mention that repeating experiments (with selection criteria of hydraulic permeability and rejection curve results.
for the membrane samples, i.e., only membranes that have water
permeability in the range of 15% relative to the average values were 3.4. Membrane surface hydrophilicity

As clearly seen in Table 3, PES membrane without an additive


had lower CA (∼65◦ ) than typically measured for non-porous PES
film (∼76◦ ) [31]. Porous structure in the outer membrane surface
would be the reason for this difference. Therefore, care should be
taken to interpret the CA results because the value is influenced not
only by membrane material but also by surface porosity. Indeed,
such effect is observed by comparing the CA data of PES(only),
PES–PEG and PES–Plu membranes. The membrane prepared with-
out an additive showed similar CA with the membranes prepared

Table 3
Static water contact angle, measured by sessile drop method, of membranes pre-
pared using different additives.

No. Membrane Contact angle (◦ )

1 PES(only) 64.5 ± 3.3


2 PES–PVP 64.1 ± 2.9
3 PES–PEG 54.1 ± 1.9
Fig. 3. Rejection curve of membranes determined by using a dextran mixture solu-
4 PES–Plu 61.7 ± 2.8
tion with total concentration of 1 g/L (as the feed) and at a pressure of 100 kPa.
130 H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135

Fig. 4. SEM images of the cross-section and membrane surface morphology: from the top to the bottom panel: PES(only), PES–PVP, PES–PEG and PES–Plu, respectively.

with a hydrophilic additive (PES–PVP and PES–PEG). This could observed (note that these membranes had similar rejection curve,
be explained because the PES(only) membrane had higher surface cf. Fig. 3). It is interesting to note that PES–PEG membrane had
porosity as noticed by its higher permeability. Comparing the CA significantly lower CA compared to other membranes indicating
data of PES–PEG and PES–Plu membranes suggests that the contri- that this was the most hydrophilic membrane, while PES–PVP and
bution of hydrophobic part of block copolymer in Pluronic could be PES–Plu membranes showed similar value.
H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135 131

Fig. 7. FTIR spectra of membranes prepared with different macromolecular addi-


Fig. 5. Zeta potentials as a function of pH calculated from the tangential streaming tives.
potential of the outer surface of different membranes (at 0.001 mol/L KCl).

mary amide stretch was observed for the membrane prepared with
3.5. Membrane surface charge
addition of PVP. By contrast, no additional peak was observed for
the membranes prepared with addition of PEG or Plu. The reason
Zeta potential was measured in order to get information on sur-
for this result would be overlapping bands of the strongest bands
face charge of the membranes. As presented in Fig. 5, all membrane
for PEG and Plu (ether) with bands for PES (ether; cf. Figs. 6 and 7).
surfaces had a negative charge over the entire pH range studied, and
Indeed, a significant increase in transmittance at ∼1105 cm−1 , due
the absolute values decreased to acidic pH range (cf. Ref. [21] for the
to additional intensity of C–O bond stretch (from PEG and Plu) was
reasons behind this phenomenon). The addition of a macromolecu-
observed and this confirms the presence of the additives in the
lar modifier into polymer solution obviously changed the effective
membrane polymer matrix. Overall, the IR spectra data indicate
surface charge of the PES membrane. Nevertheless, a linear decrease
that changes in surface chemistry were detected after addition of
in ZP with increasing pH solution for all membranes reflects typ-
the macromolecular modifiers to the membrane polymer solution.
ical behavior of a surface that has a negative charge due to anion
adsorption from the solution. Both PEG and Plu decreased the sur-
face charge toward smaller absolute value. By contrast, addition of 3.7. Membrane performance based on adsorptive fouling and
PVP increased the negative charge of membrane. These indicated ultrafiltration
that adsorption of anions (hydroxyl, chloride) from the solution was
more significant for PES–PVP than for the other membranes. The effect of additive on membrane performance was inves-
tigated with respect to adsorptive fouling and ultrafiltration.
Adsorptive fouling was studied by exposing the outer membrane
3.6. Membranes surface chemistry
surface (selective barrier side) to protein (BSA) feed solution. The
relative water flux reduction (RFR) was used to identify the extent
Figs. 6 and 7 show the IR spectra of macromolecular additives
of adsorptive fouling. As clearly seen in Fig. 8, the membranes pre-
used for membrane preparation and of the resulting membranes,
pared with an addition of hydrophilic modifier showed significantly
respectively. As expected, all membranes showed typical spectra
higher resistance towards adsorptive fouling than the membrane
of PES, i.e., aromatic bands at 1578 and 1485 cm−1 from the ben-
prepared without an additive as noticed by their much lower RFR. It
zene ring and C C bond stretch and aromatic ether band around
1240 cm−1 . A new significant peak at 1678 cm−1 assigned to a pri-

Fig. 8. Relative flux reduction after static adsorption using BSA (1 g/L in phosphate
Fig. 6. FTIR spectra of the macromolecular additives used. buffer 0.05 M, pH 7, 3 h exposure). The error bars represent standard deviation.
132 H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135

Fig. 9. Normalized flux during ultrafiltration of BSA solutions (0.1 g/L in phosphate buffer 0.05 M, pH 7) at a transmembrane pressure of 300 kPa. Water fluxes after external
cleaning with water, relative to initial water flux are also included.

should be kept in mind that the effects of adsorptive fouling depend Table 4
Apparent protein rejection.
also on the barrier pore size, and the highest flux reductions were
found for matching pores and solute sizes (cf. [31]); considering No. Membrane Rejection (%)
that the pore size distributions were different but still in the same 1 PES(only) 87
range (cf. Fig. 3), the hydrophobicity of PES seemed to have an addi- 2 PES–PVP 71
tional impact on RFR. This suggests that blending of hydrophilic 3 PES–PEG 75
4 PES–Plu 72
macromolecular additive with membrane polymer could indeed
significantly increase the hydrophilicity of the resulting membrane.
PES–PEG membrane showed the lowest RFR among the membranes
prepared with an additive. This result can be explained by the high- Indeed, the presence of hydrophilic macromolecular additive
est hydrophilicity of this membrane (cf. Table 3). It should be noted increased the normalized flux indicating that higher resistance
that the protein in solution had negative charge and if electro- towards fouling has been obtained. The membrane prepared with-
static interactions would play a dominant role, PES–PVP membrane out an additive had permeate flux of only ∼30% relative to the
should have the lowest RFR (cf. Fig. 5). initial water flux, whereas the PES–Plu membrane had the high-
Lower adsorptive fouling for PES membranes prepared with an est permeate flux (more than 70%). Of course, the highest initial
additive compared to PES(only) membrane was also observed from flux of the membrane without an additive also contributed to the
ZP measurements after adsorptive fouling (cf. Fig. 5). The change in lowest normalized flux but the effect of hydrophilic modifier was
ZP after adsorptive fouling indicates the presence of protein on the quite clear. Interestingly, at the beginning of filtration PES–PEG
membrane surface. The much larger change for PES(only) than for membrane had higher normalized flux than PES–Plu membrane
PES–Plu suggests that the degree of adsorptive fouling was higher but further decrease with filtration time was more significant. The
for PES(only) membrane. The positive value of ZP for pH values <4.9 possible reason for this phenomenon would be the stability of the
after adsorptive fouling can be explained by considering the IEP of additive in the matrix polymer membrane (cf. Section 3.8).
the BSA (pH ∼4.8), i.e. an excess of positive surface charge has been Rejection data presented in Table 4 show that the PES membrane
introduced via the protein. prepared without an additive had the highest protein rejection
To investigate ultrafiltration performance, dead-end stirred while all membranes prepared with an additive showed similar
ultrafiltration was performed with constant transmembrane pres- protein rejection. This result is in good agreement with rejection
sure (300 kPa). The results are presented in terms of permeate curve if the molar mass of BSA is fitted. In general, performance
flux relative to initial water flux (Fig. 9). It was observed that per- test showed that the membrane prepared with addition of Pluronic
meate flux dropped rapidly in the beginning of filtration for all as modifier agent showed the best performance, i.e., the lowest flux
membranes. On the one hand, this phenomenon indicates that decline and similar rejection could be obtained.
concentration polarization has taken place. On the other hand,
difference in flux profile for the membranes having similar rejec- 3.8. Stability test study
tion curve (PES–PVP, PES–PEG and PES–Plu, cf. Fig. 3) suggests
that fouling also contributed to the permeate flux decline. Water Because cleaning cannot be avoided in practical application, the
flux measurements after external cleaning using water confirmed performance of membrane prepared with a hydrophilic additive
that both reversible and irreversible fouling have occurred. Fur- will strongly depend on the stability of the additive in the mem-
thermore, higher increase in water flux after external cleaning brane polymer matrix during membrane cleaning. In this work, the
compared to permeate flux for membranes prepared with an addi- stability of hydrophilic additive in water (20 and 40 ◦ C) and sodium
tive implies that reversible fouling was more significant for those hypochlorite solution was investigated. Surface chemistry by FTIR,
membranes than for the membrane prepared without an addi- contact angle and transport property by water flux were used as
tive. It is interesting to note that the PES(only) membrane had the indicators to evaluate the membrane stability.
highest permeability and a distinctly different pore size distribu- Fig. 10 shows that no change in CA was observed for the
tion (significantly larger sieving for dextran molar masses below PES(only) membrane after soaking in water (40 ◦ C), while slight
about 50 kDa—which corresponds to the size of BSA, cf. Fig. 3). decrease was observed after incubating in NaOCl solution. Increas-
The higher observed BSA rejection could then be due to fouling ing porosity as noticed by increase in water flux (cf. Fig. 12)
(cf. below). would be the possible reason for this change. For the mem-
H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135 133

branes prepared with an additive, soaking in the water 40 ◦ C By contrast, after immersion in sodium hypochlorite solution,
did not change the CA of PES–PVP and PES–Plu membranes significant increase in flux was observed for all membranes
while soaking in NaOCl slightly decreased. Decrease in CA via including the membrane prepared without an additive. In this
increasing porosity seemed to be dominant compared to increase case, PES degradation by NaOCl leading to higher porosity was
in CA via extracting hydrophilic additive during immersing in more dominant rather than extracting of hydrophilic additive.
NaOCl solution. Indeed, significant increase in CA was observed ZP measurement after soaking in NaOCl solution (the same con-
for PES–PEG after incubating in water and NaOCl solution even centration as in stability test) supports this observation. The
only after 2 days of incubating. This indicates that the stabil- negative charge of both PES(only) and PES–Plu membranes sig-
ity of PEG in the membrane polymer matrix was quite low. nificantly increased after soaking. The ZP at pH 9.5 decreased
Consequently, hydrophilic character of the resulting membrane from −11.5 to −22 mV and from −6.1 to −11.5 for PES(only) and
would easily be reduced even when only water is used for clean- PES–Plu membranes, respectively. Chain scission of the PES C–S
ing. Similar results with respect to stability in water at 40 ◦ C bond by attack of NaOCl yielding finally sulfonic acid groups
were found during stability test in water at 20 ◦ C (data not has been evoked [24], while extraction or reduced swelling
shown). of hydrophilic macromolecular additive has also been reported
As presented in Fig. 11, slight decrease in IR absorbance for the [23,32,33].
introduced functional additive (cf. Section 3.6) was observed for
all additives used after soaking in both water and NaOCl solution.
This indicates that none of the additives was completely stable in
the matrix polymer membrane. However, irrespective this loss, sig-
nificant surface modification could still be observed as noticed by
either the appearance of new peak (for PES–PVP) or increase in
transmittance (for PES–PEG and PES–Plu) even after soaking for 10
days in all studied potential cleaning agents.
Fig. 12 shows the flux stability of the membranes dur-
ing incubating. It is clearly seen that flux of all membranes
did not change after incubating in water (data for 20 ◦ C are
not shown). Slight decreases in flux for PES–PVP and PES–PEG
membranes were still within the range of experimental error.

Fig. 10. Stability test investigated by measuring the contact angle as a function of
incubating time. Soaking the membranes in water (20 ◦ C) resulted in similar effect Fig. 11. Stability test investigated by measuring the IR absorbance of functional
with soaking in water 40 ◦ C. additive as a function of incubating time.
134 H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135

prepared with macromolecular additives, whereas the rejection


curves examined using mixture of dextrans were similar for all
three modified membranes. All membranes had negative surface
charge within the entire pH range studied. Surface hydrophilic-
ity measured with CA indicates that the PES–PEG membrane was
the most hydrophilic membrane; nevertheless, the stability of that
macromolecular additive within the membrane polymer matrix
was the most crucial problem. Performance evaluation via inves-
tigation of adsorptive fouling and ultrafiltration using BSA suggests
that PES–PEG membrane showed the lowest RFR after static adsorp-
tion followed by PES–Plu. Ultrafiltration experiments demonstrated
that the antifouling effects of PES–Plu were the most efficient
at similar protein rejection: permeate flux during ultrafiltration
using the PES–Plu was much higher than using the PES–PEG and
the PES–PVP membranes, and more than 70% of the initial water
flux could be recovered after UF just by external cleaning with
water. Overall, performance test and stability study suggest that
the membrane prepared with the addition of Pluronic as modifier
agent showed the best performance as well as the best stability;
therefore, it should be considered as additive in practical applica-
tions.

Acknowledgements

The authors thank Dieter Jacobi (Technische Chemie, Universität


Duisburg Essen) for the GPC analysis and Smail Boukercha (Anor-
ganische Chemie, Universität Duisburg Essen) for his contribution
on SEM observation. We also thank BASF (Germany) for supplying
the PES polymer.

References

Fig. 12. Stability test investigated by measuring the water flux as a function of incu- [1] M. Cheryan, Ultrafiltration and Microfiltration Handbook, Technomic Publish-
bating time. Soaking the membranes in water (20 ◦ C) yielded similar effect with ing Company Inc., Pennsylvania, 1998.
soaking in water 40 ◦ C. [2] R.W. Baker, Membrane Technology and Applications, 2nd ed., John Wiley & Sons,
Ltd., Chichester, 2004.
[3] H. Susanto, M. Ulbricht, Polymeric membranes for molecular separations, in:
In general, in this stability study, PES–Plu membrane showed E. Drioli, L. Giorno (Eds.), Membrane Operations. Innovative Separations and
Transformations, Wiley–VCH, Weinheim, in press.
the best stability followed by PES–PVP among the membranes pre-
[4] L.Y. Lafreniere, F.D.F. Talbot, T. Matsuura, S. Sourirajan, Effect of polyvinylpyrroli-
pared with an additive event though slight removal observed from done additive on the performance of polyethersulfone ultrafiltration
surface chemistry by FTIR could still be detected. The reason for membranes, Ind. Eng. Chem. Res. 26 (1987) 2385.
[5] R.M. Boom, H.W. Reinders, H.H.W. Rolevink, Th. van den Boomgaard, C.A.
this phenomenon would be that the hydrophobic part of Pluronic
Smolders, Equilibrium thermodynamics of a quaternary membrane-forming
enhances the PES–additive interaction. Not only entanglement but system with two polymers. I. Experiments, Macromolecules 27 (1994)
also hydrophobic–hydrophobic interactions may occur. The possi- 2041.
ble interaction of Pluronic in PES polymer membrane matrix was [6] M. Wienk, R.M. Boom, M.A.M. Beerlage, A.M.W. Bulte, C.A. Smolders, Recent
advances in the formation of phase inversion membranes made from amor-
described by Jiang et al. [17,20]. The interaction between PES and phous or semi-crystalline polymers, J. Membr. Sci. 113 (1996) 361.
PVP has been identified through investigation of viscoelastic behav- [7] R.M. Boom, I.M. Wienk, Th. Van den Boomgaard, C.A. Smolders, Microstruc-
ior of their polymer mixtures [4]. The authors proposed that the tures in phase inversion membranes. Part 2. The role of a polymeric additive, J.
Membr. Sci. 73 (1992) 277.
interaction between PES and PVP could be either between NH–C O [8] H.T. Yeo, S.T. Lee, M.J. Han, Role of polymer additive in casting solution in prepa-
groups in PVP and O S O groups in PES or between side cyclic ration of phase inversion polysulfone membranes, J. Chem. Eng. Jpn. 33 (2000)
groups of PVP and aromatic ring of PES. 180.
[9] D.B. Mosqueda-Jimenez, R.M. Narbaitz, T. Matsuura, G. Chowdhury, G. Pleizier,
J.P. Santerre, Influence of processing conditions on the properties of ultrafiltra-
4. Conclusions tion membranes, J. Membr. Sci. 231 (2004) 209.
[10] B. Torrestiana-Sanchez, R.I. Ortiz-Basurto, E.B. Fuente, Effect of nonsolvents on
properties of spinning solutions and polyethersulfone hollow fiber ultrafiltra-
The characteristics, performance and stability of PES membranes tion membranes, J. Membr. Sci. 152 (1999) 19.
prepared by NIPS with three different macromolecular additives, [11] N.A. Ochoa, P. Pradanos, L. Palacio, C. Pagliero, J. Marchese, A. Hernandez, Pore
i.e., PVP, PEG and Plu, have been investigated. The characteristic size distributions based on AFM imaging and retention of multidisperse poly-
mer solutes: characterisation of polyethersulfone UF membranes with dopes
of additive and the preparation conditions were maintained to be
containing different PVP, J. Membr. Sci. 187 (2001) 227.
similar so that the performance of those additives can be fairly com- [12] J. Marchese, M. Ponce, N.A. Ochoa, P. Pradanos, L. Palacio, A. Hernandez, Foul-
pared. Indeed, the presence of those macromolecular additives in ing behaviour of polyethersulfone UF membranes made with different PVP, J.
Membr. Sci. 211 (2003) 1.
membrane polymer matrix as noticed by FTIR data significantly
[13] A. Idris, N.M. Zain, M.Y. Noordin, Synthesis, characterization and performance of
determined the characteristic as well as performance of the result- asymmetric polyetehrsulfone (PES) ultrafiltration membranes with polyethy-
ing membranes. This observation was confirmed by investigation lene glycol of different molecular weights as additives, Desalination 207 (2007)
of PES membrane prepared without an additive as comparison. 324.
[14] B. Chakrabarty, A.K. Ghoshal, M.K. Purkait, Effect of molecular weight of PEG
The PES membrane prepared with addition of Pluronic (PES–Plu) on membrane morphology and transport properties, J. Membr. Sci. 309 (2008)
showed the highest hydraulic permeability among the membranes 209.
H. Susanto, M. Ulbricht / Journal of Membrane Science 327 (2009) 125–135 135

[15] Y. Liu, G.H. Koops, H. Strathmann, Characterization of morphology controlled [25] L. Wu, J. Sun, Q. Wang, Poly(vinylidine fluoride)/polyethersulfone blend mem-
polyethersulfone hollow fiber membrane by the addition of polyethylene branes: effect of solvent sort, polyethersulfone and polyvinylpyrrolidone
glycol to the dope and bore liquid solution, J. Membr. Sci. 223 (2003) concentration on their properties and morphology, J. Membr. Sci. 285 (2006)
187. 290.
[16] J.H. Kim, K.H. Lee, Effect of PEG additive on membrane formation by phase [26] P. Menut, Y.S. Su, W. Chinpa, C. Pochat-Bohatier, A. Deratani, D.M. Wang, P.
separation, J. Membr. Sci. 138 (1998) 153. Huguet, C.Y. Kuo, J.Y. Lai, C. Dupuy, A top surface liquid layer during mem-
[17] Y. Wang, T. Wang, Y. Su, F. Peng, H. Wu, Z. Jiang, Remarkable reduction of irre- brane formation using vapour-induced phase separation (VIPS)—evidence and
versible fouling and improvement of permeation properties of polyethersulfone mechanism of formation, J. Membr. Sci. 310 (2008) 278.
ultrafiltration membrane by blending with Pluronic F127, Langmuir 21 (2005) [27] T.A. Tweddle, O. Kutowy, W.L. Thayer, S. Sourirajan, Polysulfone ultrafiltration
11856. membranes, Ind. Eng. Chem. Prod. Res. Dev. 22 (1983) 320.
[18] Y. Wang, Y. Su, Q. Sun, X. Ma, X. Ma, Z. Jiang, Improved permeation performance [28] V.E. Reinsch, A.R. Greenberg, S.S. Kelley, R. Peterson, L.J. Bond, A new technique
of Pluronic F127-polyethersulfone blend ultrafiltration membranes, J. Membr. for the simultaneous, real-time measurement of membrane compaction and
Sci. 282 (2006) 44. performance during exposure to high-pressure gas, J. Membr. Sci. 171 (2000)
[19] Y.Q. Wang, Y.L. Su, X.L. Ma, Q. Sun, Z.Y. Jiang, Pluronic polymers and 217.
polyethersulfone blend membranes with improved fouling resistant ability and [29] L. Brinkert, N. Abidine, P. Aptel, On the relation between compaction and
ultrafiltration performance, J. Membr. Sci. 283 (2006) 440. mechanical properties for ultrafiltration hollow fiber, J. Membr. Sci. 77 (1993)
[20] W. Zhao, Y. Su, C. Li, Q. Shi, X. Ning, Z. Jiang, Fabrication of antifouling polyether- 123.
sulfone ultrafiltration membranes using Pluronic F127 as both surface modifier [30] K.W. Lawson, M.S. Hall, D.R. Lloyd, Compaction of microporous membranes
and pore-forming agent, J. Membr. Sci. 318 (2008) 405. used in membrane distillation. I. Effect on gas permeability, J. Membr. Sci. 101
[21] H. Susanto, M. Ulbricht, Influence of ultrafiltration membrane characteristics (1995) 99.
on adsorptive fouling with dextrans, J. Membr. Sci. 266 (2005) 132. [31] H. Susanto, S. Franzka, M. Ulbricht, Dextran fouling of polyethersulfone ultrafil-
[22] S. Rouaix, C. Causeserand, P. Aimar, Experimental study of the effects of tration membranes—causes, extent and consequences, J. Membr. Sci. 296 (2007)
hypochlorite on polysulfone membrane properties, J. Membr. Sci. 277 (2006) 147.
137. [32] S.H. Wolff, A.L. Zydney, Effect of bleach on the transport characteristics of poly-
[23] I.M. Wienk, E.E.B. Meuleman, Z. Borneman, A. van den Boomgaard, C.A. Smol- sulphone hemodialyzers, J. Membr. Sci. 243 (2004) 389.
ders, Chemical treatment of membranes of a polymer blend: mechanism of the [33] J.J. Qin, M.H. Oo, Y. Li, Development of high flux polyethersulfone hollow fiber
reaction of hypochlorite with poly(vinylpyrrolidone), J. Polym. Sci. A: Polym. ultrafiltration membranes from a low a critical solution temperature dope via
Chem. 33 (1995) 49. hypochlorite treatment, J. Membr. Sci. 247 (2005) 137.
[24] E. Arkhangelsky, D. Kuzmenko, V. Gitis, Impact of chemical cleaning on proper-
ties and functioning of polyethersulfone membranes, J. Membr. Sci. 305 (2007)
176.

You might also like