You are on page 1of 13

Journal of Membrane Science 471 (2014) 155167

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Thin-lm composite tri-bore hollow ber (TFC TbHF) membranes


for isopropanol dehydration by pervaporation
Dan Hua a,b, Yee Kang Ong b, Peng Wang b, Tai-Shung Chung a,b,n
a
b

NUS Graduate School for Integrative Science and Engineering, National University of Singapore, Singapore 117456, Singapore
Department of Chemical & Biomolecular Engineering, National University of Singapore, Singapore 117585, Singapore

art ic l e i nf o

a b s t r a c t

Article history:
Received 21 May 2014
Received in revised form
17 July 2014
Accepted 26 July 2014
Available online 14 August 2014

For the rst time, this paper reports the fabrication of novel thin-lm composite tri-bore hollow ber
(TFC TbHF) membranes for pervaporation dehydration of isopropanol (IPA). The interfacial polymerization conditions were rst optimized using the conventional single-bore hollow ber (SbHF) substrate.
Then, the effects of spinning parameters such as bore uid composition, bore uid ow rate, air gap and
dope ow rate on TFC TbHF membranes' geometry, morphology and pervaporation performance were
systematically investigated. It was found that the bore uid composition is the most critical factor.
Moreover, ne TFC TbHF membranes with an outer diameter of approximately 1 mm exhibit better
separation performance for IPA dehydration than thick TFC TbHF membranes with an outer diameter of
approximately 2 mm. The optimal TFC TbHF membrane shows a ux of 2.65 kg m  2 h  1 with a
separation factor (water/IPA) of 261 for the dehydration of an IPA/water (85/15 wt%) mixture at 50 1C.
Most importantly, the newly designed TFC TbHF membrane not only shows enhanced separation
performance and mechanical strengths as compared with the conventional TFC SbHF spun from the
identical spinning conditions, but also exhibits satisfying long-term stability, suggesting its a great
potential for pervaporation applications.
& 2014 Elsevier B.V. All rights reserved.

Keywords:
Multi-bore hollow ber (MbHF)
Tri-bore hollow ber (TbHF)
Thin lm composite (TFC)
Isopropanol
Pervaporation

1. Introduction
Alcohols are of great importance in industry because they are
widely used as good solvents, cleaning agents, raw materials or
chemical intermediates for organic synthesis, and potentially clean
liquid fuels [17]. Accordingly, the dehydration of alcohol/water
mixtures is a critical issue in the production and recycle of
alcohols. However, the alcohol (ethanol, isopropanol, or butanol)
and water can form an azeotropic mixture, rendering the purication of alcohols by conventional methods such as distillation
inefcient and uneconomic. Considering the requirements for
energy-saving and environmentally friendly processes in the
chemical industry nowadays, the pervaporation process is of
considerable interest due to its high separation efciency, ability
to break azeotropes, lower energy consumption, as well as exible
process control and module fabrication [710].
The pervaporation membrane is the heart of pervaporation
processes. Based on its conguration, it can be categorized as at
sheet, hollow ber (HF) or tubular membrane. Among them, the
polymeric HF membrane has been widely applied in various
n
Corresponding author at: National University of Singapore, Department of
Chemical and Biomolecular Engineering, 10 Kent Ridge Crescent, Singapore
117585, Singapore. Tel.: 65 6516 6645; fax: 65 6779 1936.
E-mail address: chencts@nus.edu.sg (T.-S. Chung).

http://dx.doi.org/10.1016/j.memsci.2014.07.059
0376-7388/& 2014 Elsevier B.V. All rights reserved.

separation processes due to its superior advantages such as higher


surface area, higher packing density, excellent exibility and ease
of fabrication as well as scale-up [8,1113]. Polymeric pervaporation HF membranes including single-layer asymmetric HFs and
dual-layer composite HFs are usually fabricated by means of
spinning via a non-solvent induced phase inversion process
[1321]. Since heat treatment, silicone coating and chemical
crosslinking are often needed to improve the separation efciency
which incur extra costs and the dual-layer spinning technique is
complicated, the thin-lm composite hollow ber (TFC HF) membrane fabricated via interfacial polymerization has drawn attention in recent years for pervaporation due to its excellent
separation performance and ease of fabrication [2228].
The conventional HF membrane has single-bore geometry.
Extensive research works have been focused on the design and
fabrication of single-bore hollow ber (SbHF) membranes [11].
Fine polymeric SbHFs are found to suffer from the tendency of
ber breakage during continuous operations and thus lack longterm stability [29,30]. To enhance the mechanical properties,
mixed matrix SbHF membranes were fabricated by the incorporation of inorganic llers such as carbon nanotubes, graphene based
nanomaterials, and layered silicate into the polymer matrix
[3133]. Besides, a new generation of multi-bore hollow ber
(MbHF) membranes is emerging recently [3443]. The new
membranes also possess improved mechanical properties as

156

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

compared with traditional SbHFs owing to the presence of spokes


as the mechanical reinforcement, and thus reduce ber breakage
and increase operation reliability. Comparing with the mixed
matrix SbHFs, the polymeric MbHFs are free from issues such as
ller agglomeration and poor compatibility between the inorganic
ller and the polymer matrix.
Two MbHF membranes have been commercialized for ultraltration (UF) applications. They are (1) Multibores polyethersulfone (PES) HF membranes consisting of seven bores developed by
inge GmbH [35] and (2) tri-bore PES or polyvinylidene uoride
(PVDF) HF membranes launched by Hyux [36]. Both of them
claimed greater durability and mechanical strength while maintaining the same permeate production rate as compared with the
respective SbHF membranes [35,36]. Moreover, some showed
extra benets such as a  75% reduction in energy consumption
compared with the traditional UF membrane in the pretreatment
of reverse osmosis plants for seawater desalination [37,38].
Besides UF, the applications of MbHFs were also extended to
other membrane processes. Chung and his co-workers fabricated a
series of MbHF membranes with various dimensions, congurations
and pore sizes for UF, membrane distillation (MD) and forward
osmosis (FO) applications [3942]. The detailed MbHF spinneret
designs, membrane formation mechanisms and spinning parameters

were fully explored. Their MbHFs for MD not only showed high
permeation uxes and energy efciency but also possess superior
stability and robustness. In addition, TFC MbHFs for FO were
fabricated via a spinning process followed by interfacial polymerization. These bers showed comparable separation performance with
other FO membranes. Spruck et al. also fabricated TFC seven-bore HF
membranes which showed good mechanical strengths and acceptable performance in nanoltration for desalination and water softening applications [43].
In view of the aforementioned advantages of MbHF membranes
and TFC membranes, we aim to design and explore thin-lm
composite tri-bore hollow ber (TFC TbHF) membranes for the
application of IPA dehydration. To our best knowledge, this is a
pioneering work to fabricate polymeric outer-selective TFC TbHF
membranes for pervaporation dehydration. Firstly, the effects of
interfacial polymerization conditions on pervaporation performance of TFC HF membranes were studied by using conventional
SbHF substrates. A most suitable procedure for the subsequent
fabrication of TFC TbHF membranes was then identied based on
the aforementioned study. Afterwards, the effects of spinning
conditions on membrane morphology of the TbHF substrates and
their TFC TbHF performance for pervaporation dehydration of IPA
were systematically investigated. The outcome of this study may

Fig. 1. The scheme of fabricating TFC HF with different methods: (1) the central red solid line box: the interfacial polymerization process; (2) the green dashed line box: precoating HPEI and then interfacial polymerization; (3) the blue dotted line box: post-PDMS coating after the interfacial polymerization. (For interpretation of the references to
color in this gure legend, the reader is referred to the web version of this article.)

Fig. 2. The parameters (right) for describing the conguration of the TbHF substrate (left).

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

provide useful insights towards the design of next generation HF


membranes for pervaporation applications.

2. Experimental methods

157

Table 1
Effects of amine solution composition on IPA dehydration performance of TFC SbHF a.
Condition Composition of the Total ux
Water
(g m  2 h  1) concentration at
aqueous amine
solution
the permeate (wt
%)

Separation
factor
(Water/IPA)

2.1. Materials
A commercial poly (ether imide) known as Ultems 1010 was
supplied by SABIC Innovative Plastics. N-methyl-2-pyrrolidone
(NMP, analytical grade, Merck) and ethanol (analytical grade,
Fisher) were acquired as the solvent and non-solvent respectively
to prepare the spinning solution, while NMP, 1-butanol (BuOH,
analytical grade, Fisher) as well as deionized water were used to
prepare the bore uids. A hyper-branched polyethyleneimine
(HPEI, 50 wt% in aqueous solutions, Sigma-Aldrich) with a molecular weight of 60 kg/mol was used to prepare the surface coating
solution. Sylgards184 silicone elastomer and its curing agent for
polydimethylsiloxane (PDMS) coating were purchased from Dow
Corning Singapore Pte. Ltd. The reaction monomers for the
interfacial polymerization, namely, m-phenylenediamine (MPD,
reagent grade, Tokyo Chemical Industry Co., Ltd.) and trimesoyl
chloride (TMC, reagent grade, Sigma-Aldrich) were dissolved in
aqueous and organic phases, respectively. In addition, sodium
dodecyl sulfate (SDS, reagent grade, Fluka), triethylamine (TEA,
reagent grade, Sigma-Aldrich), and cetyltrimethylammonium
bromide (CTAB, reagent grade, Sigma-Aldrich) were procured as
additives for the interfacial polymerization. Moreover, methanol
(reagent grade, Merck) and n-hexane (analytical grade, Fisher)
were employed as solvents in this study. Isopropanol (IPA, reagent
grade, Fisher) and deionized water were used to prepare the feed
mixture for pervaporation studies. All chemicals were used as
received.

Substrate Nil
1
MPD (2 wt%)
2
MPD(2 wt%) SDS
(0.1 wt%)b
3
MPD (2 wt%)
CTAB(0.1 wt%)b

61727 404
2682 7 37
25017 61

37.3 7 2.6
90.4 7 0.3
84.17 1.0

3.4 7 0.4
51.2 7 1.4
307 2.1

26377 38

83.0 7 1.0

287 2.0

a
Substrate spinning conditions: bore uid composition: NMP/H2O (95/5 wt%),
bore uid ow rate: 5 ml/min, air gap: 2.5 cm, dope ow rate: 5 ml/min.
b
TEA was used as the base source with a weight amount of 0.5 wt%.

end sealed by epoxy were immersed into the aqueous solution for
3 min at room temperature. They were then blotted with tissue
paper and dipped into the organic solution for 1 min for the
interfacial polymerization. Finally, the resultant bers were
annealed at 65 1C for 15 min to stabilize the selective TFC layer.
In the case of applying HPEI as the pre-treatment (referred to as
HPEI-TFC), the substrate was rst immersed into an aqueous 1 wt%
HPEI solution for 2 min, and then dried at 65 1C for 10 min.
Thereafter, the polyamide selective layer was synthesized as
described above. The resultant sample is referred to as HPEI-TFC.
In the case of PDMS post-treatment (referred to as TFC-PDMS), the
TFC-HF was immersed into a 3 wt% PDMS/hexane solution for
1 min and then the coated ber was cured for at least 24 h in air.
The scheme of all methods to fabricate TFC-HF membranes is
illustrated in Fig. 1.
2.4. The fabrication of HF modules

2.2. The fabrication of TbHF substrates


The TbHF substrates were prepared via a dry-jet wet spinning
process using a specically designed tri-bore spinneret with
blossom geometry [41]. The detailed spinning process has been
described elsewhere [15]. A homogeneous dope solution of
Ultems/ethanol/NMP with a ratio of 23/5/72 wt% was prepared,
poured into an ISCO syringe pump and degassed overnight before
the spinning process. This dope composition has been found to be
useful for producing SbHFs with suitable pore sizes for interfacial
polymerization [22]. Different bore uids were prepared by mixing
NMP with water or n-butanol as the non-solvent at certain ratios.
The dope solution and bore uid were concurrently extruded out
of the spinneret and the nascent bers entered into the water
coagulation tank with a certain air gap. The TbHF substrates were
then collected by a collection drum and rinsed in tap water for at
least 3 days to remove the residual solvent with water changed
daily. Then, the TbHF substrates were solvent exchanged with
three consecutive times of methanol followed by hexane. Finally,
they were air dried for further characterizations, postmodications and pervaporation tests.
2.3. The fabrication of TFC HF membranes
The TFC HF membranes were prepared by conducting interfacial polymerization on the HF substrates, where the monomer
MPD in the aqueous phase reacts with another monomer TMC in
the organic phase to form a polyamide selective layer on the outer
surface of the substrates [4447]. The aqueous phase consists of
2 wt% MPD in deionized water with or without 0.1 wt% SDS
(or CTAB) and 0.5 wt% TEA in certain conditions; while the organic
phase has 0.15 wt% TMC in hexane. The TbHF substrates with one

The HF modules were prepared by loading the HF membranes


into a module holder assembled from two Swagelok stainless steel
male run tees and a 3/8 in. peruoroalkoxy tube, and both ends of
the male run tee were sealed with epoxy. Each module contained
one HF with an effective length of 1415 cm. All modules were
cured for 48 h at ambient temperature before evaluation.
2.5. Characterizations of HF substrates and TFC HFs
The morphologies of TbHF and SbHF membranes were
observed by an optical microscope (microscope: Olympus,
SZX16; digital camera: Olympus, CMAD3) and a eld-emission
scanning electron microscope (FESEM, JEOL JSM-6700LV). The
dimensions of TbHF membranes could be determined from the
microscopic images, including the outer diameter, inner diameter,
wall thickness, as well as spoke thickness as shown in Fig. 2. The
shaded area represents the polymeric cross-section area.
The membrane surface topology was examined using a Nanoscope V atomic force microscope (AFM) from Bruker Dimension
ICON. A scanning size of 10 m  10 m was chosen and the mean
surface roughness (Ra) was used to quantify the surface roughness.
To study the surface hydrophilicity of original and HPEI coated HF
substrates, the water contact angle was measured by a Sigma 701
Tensiometer from KSV Instruments Limited. Each sample was
measured at least thrice. The membrane mechanical properties
of the fabricated HFs were characterized using an Instron tensiometer (Model 5542, Instron Corp.) at room temperature, where
the HFs were clamped at the both ends with an initial gauge
length of 50 mm and a testing rate of 10 mm/min. With the aid of
a slow beam positron apparatus, experiments were carried out
by both Doppler broadening energy spectroscopy (DBES) and

158

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

Fig. 3. FESEM images of SbHF outer surfaces made from different amine solutions, (a) substrate, (b) 1st amine solution, (c) 2nd amine solution, (d) 3rd amine solution.

Fig. 4. FESEM images of cross-section and outer surface of TFC SbHFs: (a) substrate, (b) only TFC, (c) TFC-PDMS, (d) HPEI-TFC. (The spinning conditions of the SbHF substrate:
bore uid composition: NMP/H2O (95/5 wt%), bore uid ow rate: 5 ml/min, air gap: 2.5 cm, dope ow rate: 5 ml/min.).

positron annihilation lifetime spectroscopy (PALS) to study the


depth prole of the selective layer as well as the free volume of
TFC TbHF membranes.
2.6. Pervaporation experiments
A laboratory scale pervaporation unit was employed to evaluate
the membrane performance and the details of the apparatus have
been depicted by Liu et al. [14]. A 2 L feed solution of IPA/water

mixture (85/15 wt%) was circulated through the shell side of the
module with a ow rate of 30 L/h at a feed temperature of 50 1C,
while the lumen side (permeate side) of the module was
vacuumed at a pressure below 1 mbar throughout the experiments. The system was stabilized for 2 h before the sample
collection. Thereafter, permeate samples were collected by a cold
trap immersed in liquid nitrogen. The samples were then weighted
and their compositions were analyzed by gas chromatography
(HewlettPackard GC 7890, HP-INNOWAX column, thermal

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

conductivity detector (TCD)). At least three permeate samples


were collected and their average was reported. The ux was
calculated by the following equation:
J

Q
At

wherein, Q, A and t are the weight of the permeate sample


collected, the effective membrane surface area, and the time
interval during the sample collection, respectively. The separation
factor is dened by the equation below:

ywi =ywj
xwi =xwj

wherein, yw and xw are the weight fractions of the component in


the permeate and feed, respectively; the subscripts i and j refer to
components of water and IPA, respectively.
The intrinsic properties of an asymmetric membrane could be
reected by the driving force normalized terms; namely, molebased permeance and selectivity [48]. Therefore, the ux and
separation factor were converted to the mole-based permeance
and selectivity based on the solution-diffusion model:
P
Ji
P~ i i
p
l
M i xi i psat
i  yi p

wherein, P~ i and Pi are the mole-based permeance and permeability of component i, respectively. l is the membrane thickness,
Ji, Mi, i, and psat
are the ux, molecular weight, activity coefcient,
i
and the saturated vapor pressure of component i, respectively,
xi and yi are the mole fractions of component i in the feed and
permeate, pp is the total pressure at the permeate side, which
could be assumed as zero due to the vacuum condition. Both of i
Table 2
The IPA dehydration performance of TFC SbHF membranes with different treatments a.
Coating
conditions

Total ux
Water concentration at
(g m  2 h  1) the
permeate (wt%)

Separation
factor
(Water/IPA)

Nil (substrate)
TFC (condition 1)
TFC-PDMS
HPEI-TFC

61727 404
2682 7 37
2480 7 57
23567 26

37.3 7 2.6
90.4 7 0.3
90.8 7 0.5
97.17 0.2

3.4 7 0.4
51.2 7 1.4
55.9 7 3.6
192.2 7 12.1

a
Substrate spinning conditions: bore uid composition: NMP/H2O (95/5 wt%),
bore uid ow rate: 5 ml/min, air gap: 2.5 cm, dope ow rate: 5 ml/min.

159

and psat
were determined by the Wilson equation and the Antoine
i
equation, respectively, with the aid of the AspenTech Process
Modeling software (version 7.2). Then, the mole-based selectivity
of the membrane () was dened as the ratio of the permeability
of components i and j:

P~ i =P~ j

3. Results and discussion


3.1. Effect of TFC selective layer
Since the polyamide selective layer plays an important role in
determining the overall separation performance, identifying a suitable
interfacial polymerization protocol for pervaporation dehydration of
IPA is necessary. To search for the proper protocol, SbHFs spun from
the same dope were used prior to applying to the novel TbHFs.

3.1.1. Interfacial polymerization conditions


Interfacial polymerizations with and without TEA additive (i.e., a
reaction accelerator) and surfactant (i.e., a wetting agent) have been
reported for pervaporation dehydration of alcohols [25,26]. Accordingly, three different amine solutions were conducted on SbHF
substrates to compare their effects on the polyamide selective layer
for pervaporation dehydration of IPA. As shown in Table 1, the TFCSbHF interfacially polymerized from the 1st amine solution and TMC
has greater dehydration performance in terms of ux and separation
factor than those TFC-SbHFs synthesized from TMC with the 2nd and
3rd amine solutions. Fig. 3 compares their surface morphology. The
membrane synthesized from the 1st amine solution and TMC has a
dense polyamide layer consisting of nodular structure. However, the
TFC HFs manufactured from TMC and amine solutions containing
surfactants have ake-like surface morphology. This is because
surfactants facilitate the diffusion of MPD to the organic phase, enlarge
the contact area and enhance the reaction. The resulting polyamide
layer with this morphology was reported to have a larger free volume
and loose structure [4951]. As a result, it has a lower separation factor
compared to the TFC HF synthesized from the condition 1. Therefore,
the condition 1 was applied to the subsequent sections for the
interfacial polymerization of TbHF substrates.

Fig. 5. 3-D AFM images and mean surface roughness of SbHF outer surfaces (10  10 m2) prepared under different coating methods (a) substrate (b) only TFC (c) TFCPDMS
(d) only HPEI; (e) HPEI-TFC. (The spinning conditions of the SbHF substrate: bore uid composition: NMP/H2O (95/5 wt%), bore uid ow rate: 5 ml/min, air gap: 2.5 cm,
dope ow rate: 5 ml/min.).

160

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

Fig. 6. The cross-section, outer surface as well as inner surface morphologies of TbHF substrates spun from the bore uid compositions of NMP/H2O (80/20 wt%), NMP/H2O
(95/5 wt%), NMP/BuOH (80/20 wt%), NMP/BuOH (95/5 wt%). Other spinning conditions: bore uid ow rate: 5 ml/min , air gap: 0.5 cm, dope ow rate: 5 ml/min.

Fig. 7. Schematic representation of the concept of phase inversion processes for the formation of thick and ne TbHF substrates.

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

3.1.2. Effect of HPEI or PDMS coating


To further improve the pervaporation performance of TFC HF
membranes, two treatments were explored to alleviate defects in
the polyamide selective layer. Fig. 4 shows the morphologies of four
types of bers; namely, (1) the hollow ber substrate, (2) TFC hollow
ber (i.e., interfacial polymerization under condition 1), (3) HPEI-TFC
(i.e., HPEI pre-treatment and then interfacial polymerization), and (4)
TFC-PDMS (i.e., interfacial polymerization and then PDMS coating),
while Table 2 summarizes their pervaporation performance. The HPEITFC hollow ber has the most impressive pervaporation performance.
Compared to the TFC hollow ber, it has about fourfold separation
factor without signicantly compromising the ux. This is likely due to
the fact that the HPEI coated substrate has a hydrophilic smooth
surface with smaller pores [24,25]. As a consequence, not only does
the modied substrate facilitate the absorbance of MPD for interfacial
polymerization but also promote the formation of a less defective
polyamide layer. The surface contact angle of the original HF substrate
reduced from 88.472.21 to 40.472.91 after the HPEI coating, which
proves the enhancement of surface hydrophilicity. The AFM images of
these membrane surfaces further conrm our hypotheses. As shown
in Fig. 5, the surface roughness decreases after the HPEI coating.
Moreover, the amine groups of HPEI can react with TMC, minimize the
interstitial space and thus further increase the separation factor
[24,25]. Similarly, the roughness of the TFC hollow ber decreases
after the PDMS coating. However, since there is no chemical reaction
between TFC and PDMS, the defects in the rough polyamide layer may
not be fully sealed by the PDMS coating. Thus, the TFC-PDMS hollow
ber does not show much improvement in separation factor. As a
result, the HPEI-TFC method was chosen to fabricate the TFC TbHF
membranes in the following study.

3.2. The effects of spinning parameters on TbHF substrates and


pervaporation performance of TFC TbHFs
In order to fabricate TbHF substrates with a desirable pore-size
distribution and open-cell substructure to maximize the separation
Table 3
Effects of bore uid composition on IPA dehydration performance of TFC TbHF
membranes a.
Bore uid
composition

NMP/H2O
(80/20 wt%)
NMP/H2O
(95/5 wt%)
NMP/BuOH
(80/20 wt%)
NMP/BuOH
(95/5 wt%)

Total ux
(g m  2 h  1)

Water
concentration
at the permeate
(wt%)

Separation factor
(Water/IPA)

943 7 11

77.8 7 0.5

20.17 0.6

24757 80

97.1 7 0.2

197.0 7 14.6

27307 47

95.1 7 0.3

113.0 7 6.5

3080 7 112

98.0 7 0.3

274.0 7 40.5

a
Other spinning conditions: bore uid ow rate: 5 ml/min, air gap: 0.5 cm,
dope ow rate: 5 ml/min.

161

performance of TFC TbHFs, it is necessary to investigate the key


spinning factors for the formation of TbHFs [11]. Therefore, the
effects of bore uid composition, bore uid ow rate, air-gap
distance as well as dope ow rate on TbHF morphology and its
pervaporation performance after interfacial polymerization were
systematically investigated.

3.2.1. Bore uid composition


Fig. 6 depicts the morphologies of TbHF substrates spun from
bore uids with different compositions. The bore uid composition has signicant impact toward the ber dimension and
morphology. The outer diameter, wall thickness, and spoke thickness of TbHFs spun from a strong bore uid (namely, NMP/H2O
(80/20 wt%) which contains 20 wt% H2O (strong non-solvent)) are
much larger than those TbHFs spun from three other weaker bore
uids consisting of more NMP (solvent) or BuOH (weak nonsolvent). As a result, the former is a large and thick TbHF, while the
latter three are ne TbHFs. All the ne TbHFs show partial
deformed spokes. However, they have a denser outer surface and
a much more porous inner surface compared with the thick TbHF.
The mechanism of forming ne and thick TbHFs can be
elucidated from Fig. 7. During the dry-jet wet-spinning process,
the phase inversion begins at the lumen side of the nascent ber
once the bore uid contacts with the polymer dope solution at the
exit of the spinneret. Meanwhile, a partial phase inversion occurs at
the outer surface of the nascent ber due to the moisture induced
phase inversion in the air gap region. The nascent ber would have
a complete phase inversion at its outer surface once it is fully
immersed into the external coagulation bath of water [5254].
Since NMP/H2O (95/5 wt%), NMP/BuOH (80/20 wt%) and NMP/
BuOH (95/5 wt%)) are weak bore uids, while H2O is a strong
coagulant, these result in a much slower phase inversion at the
lumen side than at the shell side of the ber [5557]. The resultant
TbHF substrate has a porous inner surface and a dense outer
surface. The slow phase inversion in the air-gap region also makes
the nascent ber stretchable by gravity. As a consequence, ne
TbHFs are produced. It is worth noting that the coagulation
strength of BuOH is weaker than water [57]; therefore, the inner
surface of TbHFs spun from NMP/BuOH has a more porous
structure than those spun from NMP/water at the same
composition.
On the contrary, a faster phase inversion occurs at the lumen
side of the ber when a stronger bore uid (i.e. NMP/H2O (80/
20 wt%)) is applied. It not only quickly solidies the inner dimension of the nascent ber but also tightens its inner skin structure.
As a result, the resultant dry-jet wet-spun TbHF has a large
diameter and a dense inner skin. Moreover, since the solidifying
rate of the lumen side is faster than that at the shell side, the net
mass transfer direction of the solvent has the tendency to move
towards the shell side of the ber, yielding a TbHF substrate with a
denser inner surface and a more porous outer surface.
As shown in Table 3, the thick TFC TbHF membrane has a much
lower ux compared with the ne ones due to its less porous
structure (typically at inner surface) and thicker walls as

Table 4
R parameters, TFC layer thicknesses and positron lifetime results of the ne and thick TFC TbHF membranes a.
Sample

R1

L1 (nm)

3 (ns)

I3 (%)

R ()

FFV (%)

Thick TbHF
Fine TbHF

0.4194 70.0004
0.4190 70.0003

199 712
101 74

2.234 70.025
1.595 70.037

12.85 70.12
10.53 70.27

3.067 0.02
2.45 7 0.04

2.78 7 0.07
1.177 0.08

a
The TbHF substrates for the thick and ne TFC TbHFs were spun from bore uids of NMP/H2O (80/20 wt%) and NMP/H2O (95/5 wt%), respectively. R1 is the R parameter
obtained from VEPFIT; L1 is the thickness of the top selective layer obtained from VEPFIT; 3 is the o-Ps lifetime obtained from PALS; I3 is the intensity of 3 obtained from
PALS; R is the mean free volume radius obtained from PALS; FFV is the fractional free volume obtained from PALS.

162

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

Fig. 8. (a) The cross-section, outer surface as well as inner surface morphologies of TbHF substrates spun from the bore uid ow rates of 3, 4 and 5 ml/min when using
NMP/H2O 95/5 wt% as the bore uid. Other spinning conditions: air gap: 0.5 cm, dope ow rate: 5 ml/min. (b) The cross-section, outer surface as well as inner surface
morphologies of TbHF substrates spun from the bore uid ow rates of 3, 4 and 5 ml/min when using NMP/BuOH 95/5 wt% as the bore uid. Other spinning conditions: air
gap: 0.5 cm, dope ow rate: 5 ml/min.

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

163

Table 5
Effects of bore uid rate on IPA dehydration performance of TFC TbHF membranesa.
Bore uid composition

Bore rate (ml/min)

Total ux (g m  2 h  1)

Water concentration at the permeate (wt%)

Separation factor (Water/IPA)

NMP/H2O (95/5 wt%)

3
4
5
3
4
5

20757 34
2348 7 29
24757 80
2099 7 54
25157 74
3080 7 112

88.9 7 0.5
95.17 0.2
97.17 0.2
95.4 7 0.3
97.6 7 0.4
98.0 7 0.3

45.4 72.2
110.7 73.8
197.0 714.6
120.7 78.2
233.6739.1
274.0 740.5

NMP/BuOH (95/5 wt%)

Other spinning conditions: air gap: 0.5 cm, dope ow rate: 5 ml/min.

Fig. 9. The cross-section, outer surface as well as inner surface morphologies of TbHF substrates spun at the air gap distances of 0.5, 2.5 and 5.0 cm. Other spinning
conditions: bore uid composition: NMP/H2O (95/5 wt%), bore uid ow rate: 5 ml/min, dope ow rate: 5 ml/min.

Table 6
Effects of air gap on IPA dehydration performance of TFC TbHF membranesa.
Air gap
(cm)

Total ux
(g m  2 h  1)

Water concentration at the


permeate (wt%)

Separation factor
(Water/IPA)

0.5
2.5
5

2475 780
2647 749
2877 7110

97.1 70.2
97.8 70.2
96.9 70.3

197.0 714.6
260.7 726.9
180 719.4

a
Other spinning conditions: bore uid composition: NMP/H2O (95/5 wt%),
bore uid ow rate: 5 ml/min, dope ow rate: 5 ml/min.

mentioned above. Moreover, the lower separation factor may be


ascribed to the relatively large pores at the ber outer surface,
which is unfavorable for forming a uniform and defect-free TFC
layer [24]. The difference in TFC layer between thick and ne
TbHFs has been further veried by PAS DBES and PALS. Firstly, the
R parameter curves of TFC TbHF membranes as a function of
positron incident energy were obtained from PAS DBES. By tting
these curves using the VEPFIT program with a three-layer model,
R1 and thickness L1 of the selective layer of TFC membranes could
be obtained [22]. As presented in Table 4, the TFC selective layer of
the ne TFC TbHF is much thinner than that of the thick TFC TbHF

(104 74 vs. 199 712 nm), thus the former has a higher ux than
the latter.
Since the o-Ps lifetime 3 and its intensity I3 obtained from
PALS analyses correspond to the free volume size and concentration, respectively, the mean free volume radius (R) and fractional
free volume (FFV) can be calculated according to established semiempirical correlation equations from the pick-off annihilation [58
60]. Table 4 shows the calculated results and indicates that the ne
TFC TbHF have a smaller R and FFV than the thick TFC TbHF. The
smaller R implies that the ne TFC TbHF has a higher rejection
against IPA than the thick TFC TbHF. In addition, the smaller FFV
does not affect the ne TFC TbHF because it has a thinner selective
layer. As a result, the ne TbHF substrates spun from weak bore
uids were more preferred for fabricating TFC membranes with
good overall pervaporation performance.

3.2.2. Bore uid ow rate


The inuences of bore uid ow rate on TbHF morphology
were investigated and shown in Fig. 8(a and b). An increase in bore
uid ow rate results in an enlargement in outer diameter and
inner diameter, as well as a reduction in wall thickness. In
addition, the large surface pores on the inner skin become small

164

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

Fig. 10. The cross-section, outer surface as well as inner surface morphologies of TbHF substrates spun from dope ow rate of 5, 6 and 7 ml/min. Other spinning conditions:
bore uid composition: NMP/H2O (95/5 wt%), bore uid ow rate: 5 ml/min , air gap: 2.5 cm.

Table 7
Effects of dope ow rate on IPA dehydration performance of TFC TbHF membranesa.
Dope ow rate Total ux
(ml/min)
(g m  2 h  1)

Water concentration at the Separation factor


permeate (wt%)
(Water/IPA)

5
6
7

97.8 7 0.2
93.6 7 0.5
90.8 7 0.3

26477 49
23347 81
2081 7 75

260.7 7 26.9
81.8 7 7.5
55.5 7 2.1

a
Other spinning conditions: bore uid composition: NMP/H2O (95/5 wt%),
bore uid ow rate: 5 ml/min, air gap: 2.5 cm.

and scattered at a higher bore uid ow rate, which may be due to


the fact that these large inner surface pores are formed because of
non-solvent intrusion from the outer surface [54,56]. This is
proven by Fig. 8 where the enlarged outer cross-section and inner
surface morphologies change as a function of bore uid ow rate.
The outward-pointed macrovoids (i.e., the tails of macrovoids face
to the outer surface [54]) across the ber walls become smaller at
a higher bore uid ow rate. As a result, a high bore-uid ow rate
would retard the non-solvent intrusion and reduce the pore sizes.
Moreover, a higher bore uid ow rate also causes more severe
deformation of the spokes because of slow phase inversion and
stress unbalance [61].
The ux and separation factors of ne TFC TbHFs prepared
under various bore uid ow rates are shown in Table 5. A higher
bore uid ow rate leads to an increase in ux, which is consistent
with aforementioned observations because of thinner wall and
less transport resistance. A higher bore uid ow rate also
enhances the separation factor by radially expanding the nascent
ber, which may result in an increase in hoop elongation and
induce molecular orientation along the circumference [62]. In
addition, the effects of bore uid ow rate on pervaporation
performance are valid for both TFC TbHFs when using NMP/H2O
(95/5 wt%) and NMP/BuOH (95/5 wt%) as bore uids. Even though
the bore uid of NMP/BuOH (95/5 wt%) produces TFC TbHFs with
better separation performance than the NMP/H2O (95/5 wt%), the

former has weaker or more bended spokes than the latter because
H2O is a much stronger coagulant than BuOH. As a result, the TbHF
substrate spun from the bore uid of NMP/H2O (95/5 wt%) at a
ow rate of 5 ml/min is preferred.
3.2.3. Air gap
Fig. 9 shows the effects of air gap on cross-section, outer and
inner skin morphologies of the TbHF substrates. The increase in air
gap results in TbHFs with a smaller outer diameter and a thinner
wall due to the elongational stretch induced by the gravitational
force [14,52]. The spokes of TbHF become severely deformed at the
air gap of 5 cm possibly due to unbalanced ow stresses and slow
phase inversion. Nevertheless, a dense outer layer and a porous
inner layer are observed for all the TbHFs.
As shown in Table 6, the ux of the TFC TbHF continuously
increases at a higher air gap due to the decrease of wall thickness
and consequently reduced transport resistance. The up and down
trend of separation factor with air gap distance is caused by two
competing factors: the elongational stress induced by the gravity
may bring about molecular orientation at the skin (i.e., selective)
layer and enhance the separation factor, while a large air gap may
overstretch the skin and create defects that favor ux but impair
the separation factor [14,52,63]. Therefore, the TbHF substrate
spun from an air gap distance of 2.5 cm is selected for the
subsequent studies considering its separation performance and
morphology in maintaining the mechanical integrity of the TbHFs.
3.2.4. Dope ow rate
The dope ow rate also inuences the TbHF morphology as
demonstrated in Fig. 10. The thicknesses of both wall and spokes
increase with an increase in dope ow rate. At the optimum air
gap of 2.5 cm, the spokes of all TbHFs are able to maintain their
geometry. However, a higher dope ow rate results in a less
porous inner surface because a thick wall may retard non-solvent
intrusion from the outer surface. As a consequence, the ux of the

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

165

Table 8
Comparisons of mechanical properties and pervaporation performance of TFC SbHF and TFC TbHF membranesa.
Comparisons
Geometries
Mechanical properties

PV performance

polymeric cross-section area (mm2)


Outer diameter (mm)
Elongation at break (%)
Maximum tensile Stress(MPa)
Young's Modulus (MPa)
Total ux (g m  2 h  1)
Separation factor

SbHF

thick TbHF

ne TbHF

0.33
1.00
42.8 72.8
11.8 70.5
350.5 7 9.0
2356 726
192.2 7 12.1

1.53
2.18
42.7 73.2
14.9 7 0.3
411.4 710.8
10087 56
22 70.6

0.30
1.02
50.0 7 2.9
15.0 70.7
426.2 7 9.1
26477 49
260.7 7 26.9

a
Spinning conditions: dope ow rate: 5 mL min  1; bore uid ow rate: 5 mL min  1; air gap: 2.5 cm; coagulant bath: water; bore uid composition NMP/H2O (95/5 wt
%) for the SbHF and ne TbHF and NMP/H2O (80/20 wt%) for the thick TbHF; take up speed: free fall.

Table 9
Performance benchmark of the ne TFC TbHF with other polymeric membranes for IPA pervaporative dehydration.
Membrane

T/oC Feed
(IPA
wt%)

Flux/
Separation factor
g m  2 h  1 (water/IPA)

mole based water permeance


(mol m  2 h  1k Pa  1)

mole-based selectivity
(water/IPA)

Ref.

Heat treated polyacrylonitrile (PAN) HF


P84 co-polyimide hollow ber
Heat treated P84/PES dual layer HF
Chitosan/polyacrylonitrile (PAN) composite HF
Torlons/P84 blended HF
Matrimids HF
Torlon 4000 T-MV/Ultem dual layer HF
6FDA-ODA-NDA/Ultems HF
Cellulose/polysulfone dual-layer HF
MPD-TMC polyamide TFC/torlon HF
HPEI-TMC polyamide TFC/ torlon HF
HPEI-HGOTMS polyamide TFC/ Ultems HF
EDA-TMC polyamide/modied PTFE at sheet
membrane
TAEA- TMC polyamide TFC/mPAN
EDA-TMC polyamide TFC/ mPAN TFC at sheet
membrane
MPD-TMC polyamide TFC /CPA-5 at sheet
membrane with heat treatment
HPEI/MPD-TMC polyamide TFC/Ultems1010 TbHF

25
60
60
25
60
80
60
60
25
50
50
50
70

90
85
85
90
85
84
85
85
95
85
85
85
70

186
883
570
145
1000
1800
765
480
4.4
1374
1980
3519
1720

1116
10585
125
2991
185
132
1944
2332
94981
53
349
278
177

5.5
3.4
2.1
4.3
3.8
2.7
3.0
1.8
0.2
7.8
12.3
21.8
3.5

1022
13227
145
2739
216
155
2275
2750
66731
62
407
324
364

[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[24]
[24]
[23]
[70]

70
25

70
90

340
213

1150
105

0.7
5.8

2394
96

[71]
[72]

16

90

140

40

11.0

59

[73]

50

85

2647

261

15.8

294

This
work

overstretch of polymer chains when the dope ow rate exceeds


a critical value [6567].
Based on the above studies on TbHFs, the ne TbHF spun from
the bore uid composition of NMP/H2O (95/5 wt%), air gap of
2.5 cm, the dope ow rate of 5 ml/min is the most preferred
membrane for IPA dehydration, owing to its excellent separation
performance and desirable geometry.

3.3. A comparison of TFC TbHF and TFC SbHF membranes

Fig. 11. Long-term pervaporation performance of the ne TFC TbHF for the
dehydration of aqueous IPA (IPA/water 85/15 wt%) at 50 1C.

TFC TbHF decreases at a higher dope ow rate due to the increased


wall thickness and less porous inner wall (shown in Table 7). The
separation factor also decreases possibly because of higher substructure resistance [64] and slight defects created by the

Table 8 compares the ne and thick TFC TbHF membranes with


the conventional TFC SbHF membranes in three aspects including
geometries, mechanical properties as well as the pervaporation
performance. In terms of geometry, the outer diameters of the ne
TbHF and SbHF are similar, which are smaller than that of the thick
TbHF. However, the ne TFC TbHF membrane outperforms the
other two in all aspects in terms of mechanical strength and IPA
dehydration performance. It has the highest tensile strength,
Young's Modulus, total ux and separation factor. In contrast, the
thick TFC TbHF shows the worst pervaporation performance.
Clearly, designing the TbHF substrate to have a proper morphology
is essential to ensure the resultant TFC TbHF membranes with
superior performance.

166

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

3.4. Benchmarking and long-term pervaporation performance


of the ne TFC TbHF
Table 9 compares the pervaporation performance of the ne
TFC TbHF with other pervaporation membranes for IPA dehydration available in literatures. The fabricated ne TFC TbHF shows a
high permeance with a reasonable selectivity among the reported
data. Moreover, it is worth mentioning that the performance of a
membrane with a low permeance but a very high selectivity is
reported to easily fall into the pressure-ratio-limited region and
thus increases the capital investment due to the requirement of a
larger membrane area [68,69]. As a result, the newly fabricated
ne TFC TbHF membrane with a high permeance and a moderate
selectivity is desirable for industrial applications in comparison to
those membranes with low permeance and very high separation
factors.
Moreover, the long-term performance stability is also an
important parameter to evaluate the membrane. Therefore, the
dehydration of aqueous IPA using the ne TFC TbHF membrane
was carried out and the permeate compositions were continuously
monitored for more than 200 h at 50 1C. From the results shown in
Fig. 11, a slight decline in ux but constant product purity are
observed, evidencing a relatively stable performance of the ne
TFC TbHF throughout the monitored period. The slight decline in
ux is possibly due to the stabilization of the TFC layer or the TbHF
substrates. It will be studied in the future.

4. Conclusions
We have molecularly designed TFC TbHF membranes with
excellent pervaporation performance for IPA dehydration and
superior mechanical strength to traditional SbHF membranes.
The optimal TFC TbHF membrane shows a ux of 2.65 kg m  2 h  1
with a separation factor of 246 for water/IPA separation at 50 1C
using 85/15 wt% IPA/water as the feed. The following conclusions
can be drawn from this study:
(1) The interfacial polymerization conditions play important roles
in determining the morphology of the TFC layer and its
performance for pervaporation dehydration. The pretreatment of HPEI on substrates not only smoothens the
substrate surface but also produces the TCF layer with fewer
defects.
(2) The effects of spinning conditions on the geometry and
morphology of TbHFs have been determined. Bore uid
composition plays the most important role. To prepare highperformance TFC TbHFs for pervaporation dehydration of IPA,
the preferred conditions to prepare TbHF substrates are as
follows: spinning from a dope made of Ultems/ethanol/NMP
(23/5/72 wt%) with a bore uid composition of NMP/H2O (95/
5 wt%), a bore uid ow rate of 5 ml/min, an air gap of 2.5 cm,
and a dope ow rate of 5 ml/min.
(3) In comparison with the TFC SbHF with a similar outer
diameter, the ne TFC TbHF shows superior pervaporation
performance and mechanical strength.
(4) The newly developed ne TFC TbHF shows both satisfactory
separation performance and long-term stability. It may have
potential for the industrial IPA dehydration application.

Acknowledgments
The authors are grateful to National Research Foundation,
Prime Minister's Ofce, Singapore for funding this research under

its Competitive Research Program for the project entitled, New


Biotechnology for Processing Metropolitan Organic Wastes into
Value-Added Products (CRP Award No. NRF-CRP 52009-5 (NUS
Grant number R-279-000-311-281)). Special thanks to Dr. Guimin
Shi, Dr. Jian Zuo, Dr. Zhengzhong Zhou and Mr. Yupan Tang for
their valuable suggestions.
References
[1] L.M. Vane, A review of pervaporation for product recovery from biomass
fermentation processes, J. Chem. Technol. Biotechnol. 80 (2005) 603629.
[2] M.L. Gimenes, L. Liu, X. Feng, Sericin/poly(vinyl alcohol) blend membranes for
pervaporation separation of ethanol/water mixtures, J. Membr. Sci. 295 (2007)
7179.
[3] Y. Huang, R.W. Baker, J.G. Wijmans, Peruoro-coated hydrophilic membranes
with improved selectivity, Ind. Eng. Chem. Res. 52 (2012) 11411149.
[4] P.S. Rachipudi, M.Y. Kariduraganavar, A.A. Kittur, A.M. Sajjan, Synthesis and
characterization of sulfonated-poly(vinyl alcohol) membranes for the pervaporation dehydration of isopropanol, J. Membr. Sci. 383 (2011) 224234.
[5] G. Liu, D. Yang, Y. Zhu, J. Ma, M. Nie, Z. Jiang, Titanate nanotubes-embedded
chitosan nanocomposite membranes with high isopropanol dehydration
performance, Chem. Eng. Sci. 66 (2011) 42214228.
[6] S. Atsumi, J.C. Liao, Metabolic engineering for advanced biofuels production
from Escherichia coli, Curr. Opin. Biotechnol. 19 (2008) 414419.
[7] G. Liu, W. Wei, W. Jin, Pervaporation membranes for biobutanol production,
ACS Sustain. Chem. Eng. 2 (2014) 546560.
[8] X. Feng, R.Y.M. Huang, Liquid separation by membrane pervaporation: a
Review, Ind. Eng. Chem. Res. 36 (1997) 10481066.
[9] L.Y. Jiang, Y. Wang, T.S. Chung, X.Y. Qiao, J.Y. Lai, Polyimides membranes for
pervaporation and biofuels separation, Prog. Polym. Sci. 34 (2009) 11351160.
[10] M.Y. Teng, K.R. Lee, S.C. Fan, D.J. Liaw, J. Huang, J.Y. Lai, Development of
aromatic polyamide membranes for pervaporation and vapor permeation, J.
Membr. Sci. 164 (2000) 241249.
[11] N. Peng, N. Widjojo, P. Sukitpaneenit, M.M. Teoh, G.G. Lipscomb, T.S. Chung, J.
Y. Lai, Evolution of polymeric hollow bers as sustainable technologies: past,
present, and future, Prog. Polym. Sci. 37 (2012) 14011424.
[12] Y. Su, G.G. Lipscomb, H. Balasubramanian, D.R. Lloyd, Observations of recirculation in the bore uid during hollow ber spinning, AIChE J. 52 (2006)
20722078.
[13] H.A. Tsai, Y.S. Ciou, C.C. Hu, K.R. Lee, D.G. Yu, J.Y. Lai, Heat-treatment effect on
the morphology and pervaporation performances of asymmetric PAN hollow
ber membranes, J. Membr. Sci. 255 (2005) 3347.
[14] R.X. Liu, X.Y. Qiao, T.S. Chung, The development of high performance P84 copolyimide hollow bers for pervaporation dehydration of isopropanol, Chem.
Eng. Sci. 60 (2005) 66746686.
[15] R.X. Liu, X.Y. Qiao, T.S. Chung, Dual-layer P84/polyethersulfone hollow bers
for pervaporation dehydration of isopropanol, J. Membr. Sci. 294 (2007)
103114.
[16] H.A. Tsai, W.H. Chen, C.Y. Kuo, K.R. Lee, J.Y. Lai, Study on the pervaporation
performance and long-term stability of aqueous iso-propanol solution
through chitosan/polyacrylonitrile hollow ber membrane, J. Membr. Sci.
309 (2008) 146155.
[17] M.M. Teoh, T.S. Chung, K.Y. Wang, M.D. Guiver, Exploring Torlon/P84 copolyamide-imide blended hollow bers and their chemical cross-linking
modications for pervaporation dehydration of isopropanol, Sep. Purif. Technol. 61 (2008) 404413.
[18] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Dehydration of alcohols by pervaporation
through polyimide matrimids asymmetric hollow bers with various modications, Chem. Eng. Sci. 63 (2008) 204216.
[19] Y. Wang, S.H. Goh, T.S. Chung, P. Na, Polyamide-imide/polyetherimide duallayer hollow ber membranes for pervaporation dehydration of C1C4
alcohols, J. Membr. Sci. 326 (2009) 222233.
[20] N. Widjojo, T.S. Chung, Pervaporation dehydration of C2C4 alcohols by 6FDAODA-NDA/Ultems dual-layer hollow ber membranes with enhanced separation performance and swelling resistance, Chem. Eng. J. 155 (2009) 736743.
[21] Z. Mao, X. Jie, Y. Cao, L. Wang, M. Li, Q. Yuan, Preparation of dual-layer
cellulose/polysulfone hollow ber membrane and its performance for isopropanol dehydration and CO2 separation, Sep. Purif. Technol. 77 (2011)
179184.
[22] J. Zuo, T.S. Chung, Design and synthesis of a uoro-silane amine monomer for
novel thin lm composite membranes to dehydrate ethanol via pervaporation,
J. Mater. Chem. A 1 (2013) 9814.
[23] J. Zuo, Y. Wang, T.S. Chung, Novel organicinorganic thin lm composite
membranes with separation performance surpassing ceramic membranes for
isopropanol dehydration, J. Membr. Sci. 433 (2013) 6071.
[24] J. Zuo, Y. Wang, S.P. Sun, T.S. Chung, Molecular design of thin lm composite
(TFC) hollow ber membranes for isopropanol dehydration via pervaporation,
J. Membr. Sci. 405406 (2012) 123133.
[25] G.M. Shi, T.S. Chung, Thin lm composite membranes on ceramic for
pervaporation dehydration of isopropanol, J. Membr. Sci. 448 (2013) 3443.
[26] P. Sukitpaneenit, T.S. Chung, Fabrication and use of hollow ber thin lm
composite membranes for ethanol dehydration, J. Membr. Sci. 450 (2014)
124137.

D. Hua et al. / Journal of Membrane Science 471 (2014) 155167

[27] S.H. Huang, C.H. Wu, K.R. Lee, J.Y. Lai, Polysulfonamide thin-lm composite
membranes applied to dehydrate isopropanol by pervaporation, Appl. Mech.
Mater. 377 (2013) 222226.
[28] H.A. Tsai, L.H. Chung, K.R. Lee, J.Y. Lai, The preparation of polyamide/
polyacrylonitrile composite hollow ber membranes for pervaporation, Appl.
Mech. Mater. 377 (2013) 246249.
[29] S. Elmore, G. Glenn Lipscomb, Analytical approximations of the effect of a ber
size distribution on the performance of hollow ber membrane separation
devices, J. Membr. Sci. 98 (1995) 4956.
[30] P. Le-Clech, A. Fane, G. Leslie, A. Childress, MBR focus: the operators'
perspective, Filtr. Sep. 42 (2005) 2023.
[31] P.S. Goh, A.F. Ismail, S.M. Sanip, B.C. Ng, M. Aziz, Recent advances of inorganic
llers in mixed matrix membrane for gas separation, Sep. Purif. Technol. 81
(2011) 243264.
[32] X.-Y. Song, Y.-P. Shi, J. Chen, A novel extraction technique based on carbon
nanotubes reinforced hollow ber solid/liquid microextraction for the measurement of piroxicam and diclofenac combined with high performance liquid
chromatography, Talanta 100 (2012) 153161.
[33] V.M. Magueijo, L.G. Anderson, A.J. Fletcher, S.J. Shilton, Polysulfone mixed
matrix gas separation hollow bre membranes lled with polymer and carbon
xerogels, Chem. Eng. Sci. 92 (2013) 1320.
[34] P. Apetel, J.-M. Espenan, Progress for extruding semi-permeable membrane
having separated hollow passgeways, US Patent 5171493; 1992.
[35] http://www.inge.ag/index_en.php?
section=multibore_membrane&pic=produkte.
[36] http://www.hyuxmembranes.com/kristal_membrane.html.
[37] D. Gille, W. Czolkoss, Ultraltration with multi-bore membranes as seawater
pre-treatment, Desalination 182 (2005) 301307.
[38] K.A. Bu-Rashid, W. Czolkoss, Pilot tests of multibore UF membrane at Addur
SWRO desalination plant, Bahrain, Desalination 203 (2007) 229242.
[39] N. Peng, M.M. Teoh, T.S. Chung, L.L. Koo, Novel rectangular membranes with
multiple hollow holes for ultraltration, J. Membr. Sci. 372 (2011) 2028.
[40] P. Wang, T.S. Chung, Design and fabrication of lotus-root-like multi-bore
hollow ber membrane for direct contact membrane distillation, J. Membr.
Sci. 421422 (2012) 361374.
[41] P. Wang, L. Luo, T.S. Chung, Tri-bore ultra-ltration hollow ber membranes
with a novel triangle-shape outer geometry, J. Membr. Sci. 452 (2014)
212218.
[42] L. Luo, P. Wang, S. Zhang, G. Han, T.S. Chung, Novel thin-lm composite tribore hollow ber membrane fabrication for forward osmosis, J. Membr. Sci.
461 (2014) 2838.
[43] M. Spruck, G. Hoefer, G. Fili, D. Gleinser, A. Ruech, M. Schmidt-Baldassari,
M. Rupprich, Preparation and characterization of composite multichannel
capillary membranes on the way to nanoltration, Desalination 314 (2013)
2833.
[44] J.E. Cadotte, Interfacially synthesized reverse osmosis membrane, US Patent
4277344; 1981.
[45] R.B. Hodgdon, Using all aliphatic polyamides from polyamino compounds and
compounds having two or more acid halide groups, US Patent 5616249; 1997.
[46] P.S. Singh, S.V. Joshi, J.J. Trivedi, C.V. Devmurari, A.P. Rao, P.K. Ghosh, Probing
the structural variations of thin lm composite RO membranes obtained by
coating polyamide over polysulfone membranes of different pore dimensions,
J. Membr. Sci. 278 (2006) 1925.
[47] A.K. Ghosh, E.M.V. Hoek, Impacts of support membrane structure and
chemistry on polyamide-polysulfone interfacial composite membranes, J.
Membr. Sci. 336 (2009) 140148.
[48] R.W. Baker, J.G. Wijmans, Y. Huang, Permeability, permeance and selectivity: a
preferred way of reporting pervaporation performance data, J. Membr. Sci. 348
(2010) 346352.
[49] M. Duan, Z. Wang, J. Xu, J. Wang, S. Wang, Inuence of hexamethyl
phosphoramide on polyamide composite reverse osmosis membrane performance, Sep. Purif. Technol. 75 (2010) 145155.
[50] Y. Mansourpanah, K. Alizadeh, S.S. Madaeni, A. Rahimpour, H. Soltani Afarani,
Using different surfactants for changing the properties of poly(piperazineamide) TFC nanoltration membranes, Desalination 271 (2011) 169177.
[51] Y. Cui, X.Y. Liu, T.S. Chung, Enhanced osmotic energy generation from salinity
gradients by modifying thin lm composite membranes, Chem. Eng. J. 242
(2014) 195203.

167

[52] T.S. Chung, X. Hu, Effect of air-gap distance on the morphology and thermal
properties of polyethersulfone hollow bers, J. Appl. Polym. Sci. 66 (1997)
10671077.
[53] H.A. Tsai, C.Y. Kuo, J.H. Lin, D.M. Wang, A. Deratani, C. Pochat-Bohatier, K.
R. Lee, J.Y. Lai, Morphology control of polysulfone hollow ber membranes via
water vapor induced phase separation, J. Membr. Sci. 278 (2006) 390400.
[54] N. Widjojo, T.S. Chung, Thickness and air gap dependence of macrovoid
evolution in phase-inversion asymmetric hollow ber membranes, Ind. Eng.
Chem. Res. 45 (2006) 76187626.
[55] J.G. Wijmans, J.P.B. Baaij, C.A. Smolders, The mechanism of formation of
microporous or skinned membranes produced by immersion precipitation, J.
Membr. Sci. 14 (1983) 263274.
[56] J.J. Qin, T.S. Chung, Effects of orientation relaxation and bore uid chemistry
on morphology and performance of polyethersulfone hollow bers for gas
separation, J. Membr. Sci. 229 (2004) 19.
[57] F. Tasselli, E. Drioli, Tuning of hollow ber membrane properties using
different bore uids, J. Membr. Sci. 301 (2007) 1118.
[58] S.J. Tao, Positronium annihilation in molecular substances, J. Chem. Phys. 56
(1972) 54995510.
[59] M. Eldrup, D. Lightbody, J.N. Sherwood, The temperature dependence of
positron lifetimes in solid pivalic acid, Chem. Phys. 63 (1981) 5158.
[60] H. Chen, W.S. Hung, C.H. Lo, S.H. Huang, M.L. Cheng, G. Liu, K.R. Lee, J.Y. Lai, Y.
M. Sun, C.C. Hu, R. Suzuki, T. Ohdaira, N. Oshima, Y.C. Jean, Free-volume depth
prole of polymeric membranes studied by positron annihilation spectroscopy: layer structure from interfacial polymerization, Macromolecules 40
(2007) 75427557.
[61] S. Bonyadi, T.S. Chung, W.B. Krantz, Investigation of corrugation phenomenon
in the inner contour of hollow bers during the non-solvent induced phaseseparation process, J. Membr. Sci. 299 (2007) 200210.
[62] Y. Wang, M. Gruender, T.S. Chung, Pervaporation dehydration of ethylene
glycol through polybenzimidazole (PBI)-based membranes. 1. Membrane
fabrication, J. Membr. Sci. 363 (2010) 149159.
[63] T.S. Chung, The limitations of using Flory-Huggins equation for the states of
solutions during asymmetric hollow-ber formation, J. Membr. Sci. 126 (1997)
1934.
[64] R.Y.M. Huang, X. Feng, Resistance model approach to asymmetric polyetherimide membranes for pervaporation of isopropanol/water mixtures, J. Membr.
Sci. 84 (1993) 1527.
[65] T.S. Chung, J.J. Qin, J. Gu, Effect of shear rate within the spinneret on
morphology, separation performance and mechanical properties of ultraltration polyethersulfone hollow ber membranes, Chem. Eng. Sci. 55 (2000)
10771091.
[66] I.D. Sharpe, A.F. Ismail, S.J. Shilton, A study of extrusion shear and forced
convection residence time in the spinning of polysulfone hollow ber
membranes for gas separation, Sep. Purif. Technol. 17 (1999) 101109.
[67] A.F. Ismail, M.I. Mustaffar, R.M. Illias, M.S. Abdullah, Effect of dope extrusion
rate on morphology and performance of hollow bers membrane for ultraltration, Sep. Purif. Technol. 49 (2006) 1019.
[68] R.W. Baker, Membrane Technology and Applications, England, John Wiley and
Sons, 2004.
[69] Y. Huang, T.C. Merkel, R.W. Baker, Pressure ratio and its impact on membrane
gas separation processes, J. Membr. Sci. 463 (2014) 3340.
[70] Y.L. Liu, C.H. Yu, J.Y. Lai, Poly(tetrauoroethylene)/polyamide thin-lm composite membranes via interfacial polymerization for pervaporation dehydration on an isopropanol aqueous solution, J. Membr. Sci. 315 (2008) 106115.
[71] C.L. Li, S.H. Huang, D.J. Liaw, K.R. Lee, J.Y. Lai, Interfacial polymerized thin-lm
composite membranes for pervaporation separation of aqueous isopropanol
solution, Sep. Purif. Technol. 62 (2008) 694701.
[72] S.H. Huang, G.J. Jiang, D.J. Liaw, C.L. Li, C.C. Hu, K.R. Lee, J.Y. Lai, Effects of the
polymerization and pervaporation operating conditions on the dehydration
performance of interfacially polymerized thin-lm composite membranes, J.
Appl. Polym. Sci. 114 (2009) 15111522.
[73] J. Albo, J. Wang, T. Tsuru, Application of interfacially polymerized polyamide
composite membranes to isopropanol dehydration: effect of membrane pretreatment and temperature, J. Membr. Sci. 453 (2014) 384393.

You might also like