You are on page 1of 10

Journal of Applied Science and Engineering, Vol. 21, No. 1, pp. 89-98 (2018) DOI: 10.6180/jase.201803_21(1).

0011

Effect of Inner Non-solvent Condition on Morphology


of PVDF Hollow Fiber Membranes Prepared by NIPS
Spinning Process
Hsu-Hsien Chang1,2, Sheng-Chang Chen1,2 and Liao-Ping Cheng1,3*
1
Department of Chemical and Materials Engineering, Tamkang University,
Tamsui, Taiwan 251, R.O.C.
2
Bright Sheland International Co., New Taipei City, Taiwan 221, R.O.C.
3
Energy and Opto-Electronic Materials Research Center, Tamkang University,
Tamsui, Taiwan 251, R.O.C.

Abstract
Hollow fiber membranes of PVDF were prepared by isothermal wet spinning process from the
water/DMF/PVDF system at 25 °C. The dope solution comprised 21 wt% PVDF in DMF, and pure
water was employed both as the inner and outer coagulants. SEM imaging of the membrane indicated
an unexpected interesting morphology: columnar macrovoids extended from the outer surface to the
central region, while cellular pores constituted the inner half of the membrane. As water was a harsh
nonsolvent, one tended to think that macrovoids shall form towards the inner surface of the hollow
fiber. Absence of macrovoids in this region was explained based on the fact that DMF would
accumulate rapidly in the inner extruding stream as it underwent mass-exchange with water in the
inner bath. Such activity rendered effectively the coagulant became effectively a soft one. As a result,
macrovoids was eliminated and the inner surface became porous.

Key Words: PVDF, Hollow Fiber, Membrane, Morphology, NIPS

1. Introduction the formed membranes are collected by means of a take-


up device [1-10]. By varying the spinning parameters,
Porous hollow fiber membranes are used over a wide e.g. the inner and outer diameters of the spinneret, the
range of applications, such as reverse osmosis, and ul- compositions of the polymer solution, non-solvent st-
tra-filtration, membrane distillation, etc. [1-5]. They are rength of the inner and outer baths, etc., porous hollow
generally prepared by the non-solvent induced phase se- fiber membranes can be produced encompassing a wide
paration (NIPS) spinning process. In such process, a range of morphologies.
polymer solution (dope) consisting of polymer, solvent, NIPS process typically results in asymmetric or sym-
and sometimes additives is pumped through the annulus, metric membranes (flat sheet or hollow fiber). Asymmet-
and a bore injection fluidis pumped through the inner ric membrane has a skin layer on the surface which ser-
needle tube of the spinneret. Together these two streams ves as a selective barrier to differentiate incoming feeds.
are extruded into a non-solvent bath where exchange of Underneath the skin is a porous bulk, which functions as
solvent and non-solvent across the dope/bath interfaces the mechanical support to sustain the transmembrane
induces instability and subsequent phase separation of pressure during filtration operations [11-17]. The po-
the dope solution. After appropriate washing and drying, rous bulk may be sponge-like or contain teardrop- or fin-
ger-like macrovoids, depending on the preparation con-
*Corresponding author. E-mail: lpcheng@mail.tku.edu.tw ditions. Asymmetric structure is typically formed by pre-
90 Hsu-Hsien Chang et al.

cipitation of the dope solution in a so-called harsh non- 2.2 Phase Diagram Determination
solvent bath, e.g. water, such that polymer can accumu- The gelation phase boundary at 25 °C for the ternary
late rapidly near the surface region (contacting the bath) system, water/DMF/PVDF, was determined by the widely
to give a continuous gel layer that eventually develops used cloud point method [20,23]. Briefly, a specific
into a skin. The pores in the bulk are derived from poly- amount of PVDF (dried in an oven at 60 °C) was mixed
mer-lean phase formed during phase separation induced with DMF and sealed in a glass bottle with Teflon-lined
by the influx of non-solvent. In contrast to asymmetric cap. The mixture was blended at 60 °C until the polymer
membrane, symmetric membrane is skinless and has a was completely dissolved. A known quantity of water
relatively uniform pore distribution across the mem- was then added to the polymer solution. Local precipita-
brane [18-24]. The membrane may be formed by carry- tion may be observed at the contact of water with the so-
ing out polymer precipitation in a soft nonsolvent bath lution. The mixture (termed dope) was further blended at
(e.g., a bath containing significant amount of solvent) 60 °C until it became clear and homogeneous again.
such that exchange of solvent and nonsolvent is slow and Then it was put in a thermostat maintained at 25 °C over
polymer concentration would not build-up sharply dur- the period of 30 days during which gelation might occur
ing the phase separation process [18-24]. and the dope became a gel. The gelation points in the
In the present research, porous PVDF hollow fiber phase diagram were identified as the compositions at
membranes were prepared by isothermal NIPS process which gelation first occurred in a series of samples with
from the water/DMF/PVDF system. Morphologies of increasing non-solvent/solvent ratios.
the membranes were observed by means of SEM. It is in- The terpolymer (Kynar 9301) is largely amorphous,
teresting to find that finger-like macrovoids on the lumen and hence can be used as a model to access the liquid-liq-
(inner) side can be eliminated simply by using spinneret uid (L-L) demixing boundary (often termed binodal in
with a sufficiently small inner tube. Such is the case even the literature). Such is based on the fact that Kynar 9301
when a harsh nonsolvent, e.g., water, is employed as the is chemical structurally similar to PVDF (Kynar 740)
inner coagulant. The reason for this unusual new finding and that these two polymers have close solubility para-
is explained based on mass transfer and phase separation meters, 7.2 and 7.74 (cal/cm3)1/2 respectively for the ter-
considerations. With a macrovoid-free lumen side, the polymer and PVDF. To experimentally access the bino-
membrane is expected to be stronger than ordinary ultra- dal, a series of solutions composed of water/DMF/Kynar
or nano-filtration membranes that have macrovoids on 9301were prepared and then put in a thermostat main-
both inner and outer sides of the membrane. tained at 25 °C, as in the procedure described above. Un-
stable solutions underwent L-L demixing into two clear
2. Experimental homogeneous liquid layers. The binodal points were de-
fined as the compositions of the separated liquid phases.
2.1 Material It is also noted that the dope would not gel (still flows)
Poly(vinylidene fluoride) (PVDF) polymer (Kynar even at the concentration of 50 wt.%, which may be as-
740, Elf Atochem; intrinsic viscosity = 0.881 dL/g, Mn = sociated with the low molecular weight of the polymer.
254,000 g/mole) and a terpolymer (Kynar 9301) of vin-
ylidene fluoride (VDF), tetrafluoro ethylene (TFE), and 2.3 Membrane Preparation and SEM Observation
hexafluoropropylene (HFP) with VDF/HFP/TFE = 60/ A schematic representation of the employed NIPS
20/20, intrinsic viscosity = 0.4 dl/g, and Mn = 79200 g/ spinning process is illustrated in Figure 1. The dope so-
mol was supplied by Elf Atochem Inc. N,N-dimethyl- lution for spinning was prepared by dissolving 21 wt%
formamide (DMF, Baker Analyzed, reagent grade) was of PVDF in DMF at 60 °C. It was cooled to 25 °C and
used as the solvent for preparation of dopes for hollow fi- then pumped through the annulus of the spinneret. On
ber spinning. Distilled-deionized water was used as the the other hand, water was pumped through the central
non-solvent for precipitation of PVDF from the dope. tube of the (ID/OD = 0.16/0.4 mm). Together these two
All materials were used as received. streams entered the coagulation bath where phase sepa-
Effect of Inner Non-solvent Condition on Morphology of PVDF Hollow Fiber Membranes Prepared by NIPS Spinning Process 91

Figure 1. Schematic representation of the hollow fiber spinning process. 1: inner non-solvent fluid; 2: hollow fiber spinning solu-
tion; 3: outer coagulation bath; P: pumps; (a): fluids in spinneret and inner coagulation bath; (b): cross section of the in-
ner tube and the spinneret.

ration occurred and polymer precipitated to form a po- 3. Results and Discussion
rous hollow fiber membrane. The membrane was wa-
shed repeatedly first with distilled water, then with iso- 3.1 Phase Diagram of the Water/DMF/PVDF System
propanol, followed by hexane. Afterwards, it was dried In Figure 2, the experimental phase equilibrium cur-
at 50 °C in a convective oven. Flat sheet membranes ves at 25 °C are presented both for the amorphous ter-
were also prepared for comparison of porous structures. polymer and the crystalline PVDF in water-DMF sys-
The membrane was prepared by immersion the 21 wt% tems. The solid circles (·) represent the cloud point com-
dope cast on a glass plate into a water bath at 25 °C. The positions measured for the terpolymer in water-DMF so-
nascent membrane, as being peeled off from the glass lutions. These points mark the L-L demixing phase
plate, was washed and dried with the same procedure for boundary; inside which an originally homogeneous solu-
processing hollow fiber membranes. tion will separate into two liquid phases in equilibrium.
Morphologies of the porous membranes (both flat The filled triangle points (s) in Figure 2 depict the crys-
sheet and hollow fiber) were observed using a field emis- tallization-induced gelation equilibrium points observed
sion scanning electron microscope (FESEM, Leo 1530, for crystalline PVDF in water-DMF solutions. These
Carl Zeiss, Oberkochen, Germany). A piece of mem- data define a gelation phase boundary. Above the gela-
brane sample was vacuum-dried and then attached to a tion line is a single-phase region, in which a homoge-
sample holder by means of conductive copper tapes. The neous dope can be prepared with a long-term stability.
cross section of the membrane was obtained by fractur- Below this line is a metastable zone, in which an origi-
ing in liquid nitrogen. Silver paste was applied at the nally homogeneous dope solution will gel upon standing
edges of the sample to enhance electronic conductivity. at 25 °C over an extended period of time. The fact that
Then, the sample was sputtered with a thin layer (~2 nm) gelation was induced by crystallization has been proven
of Pt-Pd alloy and observed under a low acceleration by various techniques, such as differential scanning cal-
voltage, 2 kV, by means of an in-lens detector. orimetry (DSC), polarized optical microscopy (POM),
92 Hsu-Hsien Chang et al.

SEM, etc. [11,25]. As the gelation region encloses the served. On the pore walls, spherulitic entities (e.g., cen-
binodal, a dope inside the binodal can undergo both L-L tral region, pointed out by an arrow) are in evidence,
demixing and crystallization types of phase separation. which were derived from spherulitic growth of PVDF
crystallites. Top surface of the membrane, as shown in
3.2 Morphology of PVDF Hollow Fiber and Flat Figure 3(d), is dense without appreciable pores under the
Membranes resolution of ca. 40 nm. Such is typical of immersing a
The casting dope for preparing the PVDF flat mem- good dope solution in a harsh non-solvent bath, such as
brane consisted of 21 wt% PVDF dissolved in DMF (cf., water. Upon contact of dope and bath, a stiff gellayer
the solid square point in Figure 2). It was a clear uniform forms due to a sharp upsurge of polymer concentration at
viscous solution. When immersed in pure water (a typi- the bath-dope interface, which subsequently develops
cal harsh nonsolvent) it underwent a very rapid precipi- into the dense skin of the membrane. The bottom surface,
tation process to form a membrane with morphology that as shown in Figure 3(e), consists of truncated PVDF glo-
is characteristic of L-L demixing dominant precipitation bular objects along with cellular pores. Figure 3(f) shows
process. The total cross section of the membrane, as the high magnification of Figure 3(e), nano-granular and
shown in Figure 3(a), comprises finger-like macrovoids dendritic features are clearly demonstrated, and because
and cellular pores resembling those seen in common am- the globules grow pressing-against the glass plate during
orphous membranes. That is, L-L demixing took place precipitation, they appear flattened [23-25].
first and set-in the overall porous morphology before Asymmetric polymer membranes consisting of a thin
crystallization occurred in the polymer-rich phase that
surrounded the polymer-poor liquid droplets. Figure 3(b)
shows the high magnification image of the upper part of
Figure 3(a). A thin skin followed by parallel columnar
macrovoids extending into the central regionis seen.
Roughened and dendritic features on the pore walls sug-
gest occurrence of crystallization during the late stage of
the phase separation process. DSC and XRD analyses of
similar PVDF membranes can be found in the literature
[25]. Figure 3(c) shows the enlarged view of the cross
section near the bottom surface of the membrane. Cellu-
lar pores of about 3-5 mm in diameter are clearly ob-

Figure 3. Morphologies of the cross section, top, and bottom


surfaces of the flat sheet membrane prepared by im-
mersing a 21 wt% PVDF/DMF dope in pure water.
Figure 2. Phase diagram of the water/DMF/PVDF ternary (a) total cross section; (b) high magnification of the
system at 25 °C. Binodal was determined for the upper part of (a); (c) high magnification of the
water/DMF/Kynar 9301 system. Gelation line was lower part of (a); (d) top surface; (e) bottom sur-
determined for the water/DMF/Kynar 740 system. face; (f) high magnification of (e).
Effect of Inner Non-solvent Condition on Morphology of PVDF Hollow Fiber Membranes Prepared by NIPS Spinning Process 93

top layer supported by a porous sublayer composed of outer non-solvent bath and an inner stream of non-sol-
macrovoids and cellular pores are often observed for vent coagulant that flows through the center region of the
membranes prepared by NIPS process in a harsh non- system. That is, there exist two membrane/bath inter-
solvent bath [1,11-17,26-31]. It is well accepted that faces, across which solvent and non-solvent exchange
cellular pores are resulted from radial growth of poly- occurs. As harsh non-solvent was employed for both the
mer-poor phase during L-L demixing. As to the macro- inner and outer baths in the present work, one would
voids, the causes of their formation have been exten- speculate that the morphology of the membrane shall be
sively studied in the literature [26-31]. A number of similar to the combination of two flat-sheet membranes
mechanisms have been proposed, e.g., surface breakage arranged in the Top/Bottom/Bottom/Top order, as shown
due to shrinkage stress, liquid-liquid demixing and dif- in Figure 6. That is, finger-like macrovoids dominate the
fusional growth of cellular pores, osmosis effect during regimes near the inner and outer surfaces, while cellular
growth of cellular pores, instability caused by perturba- pores are inserted in-between the macrovoids. The real
tion of the interface in steep concentration profile, etc. situation is quite different, however, as shown below.
Based on Mulder’s proposition [1], formation of macro- SEM photo micrographs of the PVDF hollow fiber
voids is initiated by L-L demixing. The liquid micro- membrane prepared by spinning of 21 wt% PVDF/DMF
drops (micelles) of the polymer-poor phase are the em- dope solution in pure water are shown in Figure 7. Pre-
bryos for macrovoids developments. A schematic repre-
sentation is illustrated in Figure 4 for such case. L-L
demixing occurs rapidly after the binodal is crossed due
to exchange of nonsolvent and solvent at the mem-
brane-bath interface. This causes formation of first mi-
crodrops at t = t1 near the top surface. Continuing in-
fluxes of solvent and non-solvent into the microdrops, as
driven by osmosis forces, lead to the downward-growth
of macrovoids (upward growth is unlikely due to higher
polymer concentration there is) cf., t = t2. This activity
proceeds until the polymer concentration of the poly-
mer-rich phase enclosing the macrovoids becomes so
high that gelation occurs, and further macrovoids growth
becomes inhibited, cf., t = t3. Finally, asymmetric mem-
braneis formed with macrovoids extending from the top
surface to the central or even bottom region of the mem-
brane.
A schematic representation of the mass transfer ac-
ross the polymer solution/bath interface during immer-
sion of a casting dope in the form of flat sheet is shown in
Figure 5(a). The solvent diffuses into the non-solvent
bath with flux J2, whereas the non-solvent diffuse into
the casting dope with flux J1. As is well recognized that
such mass transfer gives rise to phase separations (the
dope crosses the binodal/crystallization boundaries) and
lead to porous structure formation in the membrane. For Figure 4. Schematic representation of the growth of macro-
the spinning of hollow fibers, a somewhat different cir- voids at different times (t1, t2, and t3) during precipi-
tation. t1: initial formation of microdrops near the
cumstance is encountered. As shown in Figure 5(b), the top surface. t2: downward-growth of macrovoids.
polymer solution is sandwiched between a reservoir of t3: completeness of macrovoids formation.
94 Hsu-Hsien Chang et al.

cipitation is observed to occur very rapidly to yield a


morphology dictated by L-L demixing. The cross section
of the hollow fiber membrane, cf., Figure 7(a), exhibits
finger-like macrovoids from outer surface to the central
part of the membrane followed by cellular pores all the
way to the inner surface, which is similar to those shown
in Figure 3(a). The outer surface, cf., Figure 7(b), forms a
relatively dense skin with sporadic nm-scale dints evenly
distributed on the surface. The magnified view of the in-
ner surface can be seen from Figure 7(c). It is corrugate
with identifiable pores of ~2 mm; namely, the surface is
not a skin. Surprisingly, finger-like macrovoids are ab-
sent towards the lumen side of the hollow fiber, even
with water being used as the inner coagulant, as opposed
to what is illustrated in Figure 6. The reason can be ex-
plained based on mass transfer events that occur across
the inner-interface during immersion. As DMF in the
dope out flows into the inner tube, water and DMF mixed
quickly to make the inner bath a soft coagulant (i.e., con-
Figure 5. Schematic representation of a polymer solution/ taining significant amount of solvent) for the polymer.
bath interface. (a) formation of a flat membrane by Such is possible as long as the inner tube is thin enough
NIPS method; (b) formation of a hollow fiber mem-
brane by NIPS spinning process. J1: flux of non- to allow a quick build-up of DMF concentration in the
solvent diffusing into the dope. J2: flux of solvent inner bath. In this case, the inner gel layer would be soft
diffusing into the bath.

Figure 6. Schematic representation of the formation of PVDF follow fiber membrane by NIPS spinning process.
Effect of Inner Non-solvent Condition on Morphology of PVDF Hollow Fiber Membranes Prepared by NIPS Spinning Process 95

(as oppose to the rigid one on the outer surface) and eas-
ily broken into pores during subsequent precipitation
process [20-23]. Shieh et al. has demonstrated a similar
immersion situation for the water/formic acid/Nylon-66
system. The calculated diffusion trajectory indicates clearly
accumulation of solvent in the inner bath, and hence the
polymer concentration at the inner-interface is much
lower than that at the outer-interface. By contrast, if a
large inner tube is used, the solvent-concentration built-
up in the inner bath may become ineffective. In such
case, the inner bath remains harsh and the formed mem-
brane has macrovoids from both inner and outer sides of
the hollow fiber, as is illustrated in Figure 8. The mem-
brane was prepared by spinning 21% PVDF/DMF dope
in water by means of a spinneret with larger inner tube
(ID/OD = 0.33/0.6 mm/mm, comparing with the small
tube, 0.16/0.4, shown in Figure 1). Obviously, the cross
section has a macrovoids/cellular pore/macrovoids sand-
wich structure, being consistent with the immersion situ-
ation shown in Figure 6 where dope solution contacts
harsh bath on both sides.

4. Conclusions

PVDF membranes in the forms of hollow fibers and


flat sheets were prepared by nonsolvent induced phase
separation at 25 °C from the water/DMF/PVDF system.
SEM micrographs of the flat sheet membrane indicate
an asymmetric structure featuring finger-like macro-

Figure 7. SEM micrographs of the hollow fiber membrane


prepared by spinning of a 21 wt% solution in water Figure 8. SEM micrograph of the PVDF hollow fiber mem-
bath. (a) outer surface/cross section; (b) outer sur- brane prepared by using a spinneret with larger in-
face (c) inner surface/cross section. ner tube (ID/OD = 0.33/0.6 mm/mm).
96 Hsu-Hsien Chang et al.

voids and cellular pores, typical of L-L demixing domi- [6] Kang, G. D. and Cao, Y. M., “Application and Modifi-
nated phase separation process during precipitation. cation of Poly(vinylidene fluoride) (PVDF) Mem-
For the PVDF hollow fiber membrane spun by means of branes – A Review,” Journal of Membrane Science,
a spinneret with small inner tube, a similar asymmetric Vol. 463, No. 1, pp. 145-165 (2014). doi: 10.1016/
structure was demonstrated, in which finger-like ma- j.memsci.2014.03.055
crovoids were absent from the inner (lumen) side con- [7] Drioli, E., Ali, A., Simone, S., Macedonio, F., Al-Jiil,
tacting the harsh water bath. Such unexpected result S. A., Shabonah, F. S., Al-Romaih, H. S., Al-Harbi, O.,
was explained based on the fact that the out-fluxed Figoli, A. and Criscuoli, A., “Novel PVDF Hollow Fi-
DMF can quickly soften the small inner bath and delay ber Membranes for Vacuum and Direct Contact Mem-
the demixing process. In case that a larger inner tube brane Distillation Applications,” Separation and Puri-
was employed in the spinneret, macrovoids formed on fication Technology, Vol. 115, No. 30, pp. 27-38
the lumen side as usual simply because the amount of (2013). doi: 10.1016/j.seppur.2013.04.040
out-fluxed DMF was too small to effectively soften the [8] Bottino, A., Camera, G. R., Capannelli, G. and Munari,
inner bath. S., “The Formation of Microporous Polyvinylidene
Difluoride Membranes by Phase Separation,” Journal
Acknowledgement of Membrane Science, Vol. 577, No. 1, pp. 1-20
(1991). doi: 10.1016/S0376-7388(00)81159-X
The authors would like to thank the Ministry of Sci- [9] Lin, D. J., Lin, C. L. and Guo, S. Y., “Network
ence and Technology of Taiwan for financial support. Nano-porous Poly(vinylidene fluoride-co-hexafluoro-
(NSC 99-2221-E-032-002-MY3). propene) Membranes by Nano-gelation Assisted Phase
Separation of Poly(vinylidene fluoride-co-hexafluoro-
References propene)/Poly(methyl methacrylate) Blend Precursor
in Toluene,” Macromolecules, Vol. 45, No. 21, pp.
[1] Strathmann, H., Introduction to Membrane Science 8824-8832 (2012). doi: 10.1021/ma301075e
and Technology, Wiley-VCH Verlag & Co., Germany [10] Xu, J. and Xu, Z. L., “Poly(vinyl chloride) (PVC) Hol-
(2011). low Fiber Ultrafiltration Membranes Prepared from
[2] Mulder, M., Basic Principles of Membrane Technol- PVC/additives/solvent,” Journal of Membrane Sci-
ogy, 2nd ed., Kluwer Academic Publisher, Dordrecht ence, Vol. 208, No. 1-2, pp. 203-212 (2002). doi:
(1996). 10.1016/S0376-7388(02)00261-2
[3] Culfaz, P. Z., Rolevink, E., Rijn, C. V., Lammertink, R. [11] Cheng, L. P., “Effect of Temperature on the Formation
G. H. and Wessling, M., “Microstructured Hollow Fi- of Microporous PVDF Membranes by Precipitation
bers for Ultrafiltration,” Journal of Membrane Sci- from 1-Octanol/DMF/PVDF and Water/DMF/PVDF
ence, Vol. 347, No. 1-2, pp. 32–41 (2010). doi: 10. Systems,” Macromolecules, Vol. 32, No. 20, pp.
1016/j.memsci.2009.10.003 6668-6674 (1999). doi: 10.1021/ma990418l
[4] Peng, N., Widjojo, N., Sukitpaneenit, M. M., Teoh, P., [12] Sukitpaneenit, P. and Chung, T. S., “Molecular Eluci-
Lipscomb, G. G., Chung, T. S. and Lai, J. Y., “Evolu- dation of Morphology and Mechanical Properties of
tion of Polymeric Hollow Fibers as Sustainable Tech- PVDF Hollow Fiber Membranes from Aspects of Phase
nologies: Past, Present, and Future,” Progress in Poly- Inversion, Crystallization and Rheology,” Journal of
mer Science, Vol. 37, No. 10, pp. 1401–1424 (2012). Membrane Science, Vol. 340, No. 1-2, pp. 192-205
doi: 10.1016/j.progpolymsci.2012.01.001 (2009). doi: 10.1016/j.memsci.2009.05.029
[5] Liu, F., Hashim, N. A., Liu, Y., Abed, M. R. M. and Li, [13] Ansorge, W., Schuster, O., Wechs, F. and Dombrowski,
K., “Review-progress in the Production and Modifica- K., U.S. Patent 8,727,136 B2 (2014).
tion of PVDF Membranes,” Journal of Membrane [14] Setiawan, L., Wang, R., Li, K. and Fane, A. G., “Fabri-
Science, Vol. 375, No. 1-2, pp. 1-27 (2011). doi: cation and Characterization of Forward Osmosis Hol-
10.1016/j.memsci.2011.03.014 low Fiber Membranes with Antifouling NF-like Selec-
Effect of Inner Non-solvent Condition on Morphology of PVDF Hollow Fiber Membranes Prepared by NIPS Spinning Process 97

tive Layer,” Journal of Membrane Science, Vol. 394- [22] Young, T. H., Lin, C. W., Cheng, L. P. and Hsieh, C. C.,
395, No. 15, pp. 80-88 (2012). doi: 10.1016/j.memsci. “Preparation of EVAL Membranes with Smooth and
2011.12.026 Particulate Morphologies for Neuronal Culture,” Bio-
[15] Gryta, M. and Barancewicz, M., “Influence of Mor- materials, Vol. 22, No. 13, pp. 1771-1777 (2001). doi:
phology of PVDF Capillary Membranes on the Per- 10.1016/S0142-9612(00)00337-9
formance of Direct Contact Membrane Distillation,” [23] Chang, H. H., Chang, L. K., Yang, C. D., Lin, D. J.
Journal of Membrane Science, Vol. 358, No. 1-2, and Cheng, L. P., “Effect of Polar Rotation on the
pp. 158-167 (2010). doi: 10.1016/j.memsci.2010. Formation of Porous Poly(vinylidene fluoride) Mem-
04.044 branes by Immersion Precipitation in an Alcohol
[16] Shih, C. H., Gryte, C. C. and Cheng, L. P., “Precipita- Bath,” Journal of Membrane Science, Vol. 513, No.
tion Dynamics for the Formation of Nylon-6 Poly- 1, pp. 186-196 (2016). doi: 10.1016/j.memsci.2016.
amide Membranes by Isothermal Precipitation in Wa- 04.052
ter/Formic Acid Solutions,” Journal of Applied Sci- [24] Chang, H. H., Chen, S. C., Lin, D. J. and Cheng, L. P.,
ence and Engineering, Vol, 15, No. 4, pp. 357-366 “The Effect of Tween-20 Additive on the Morphology
(2012). doi: 10.6180/jase.2012.15.4.06 and Performance of PVDF Membranes,” Journal of
[17] Lin, D. J., Chang, C. L., Chen, T. C. and Cheng, L. P., Membrane Science, Vol. 466, No. 15, pp. 302-312
“On the Structure of Porous Poly(vinylidene fluoride) (2014). doi: 10.1016/j.memsci.2014.05.011
Membrane Prepared by Phase Inversion from Water- [25] Lin, D. J., Chang, C. L., Huang, F. M. and Cheng, L. P.,
NMP-PVDF System,” Tamkang Journal of Science “Effect of Salt Additive on the Formation of Micro-
and Engineering, Vol. 5, No. 2, pp. 95-98 (2002). doi: porous Poly(vinylidene fluoride) Membranes by Phase
10.6180/jase.2002.5.2.04 Inversion From LiClO4/Water/DMF/PVDF System,”
[18] Lin, D. J., Beltsios, K., Chang, C. L. and Cheng, L. P., Polymer, Vol. 44, No. 2, pp. 413-422 (2003). doi:
“Fine Structure and Formation Mechanism of Particu- 10.1016/S0032-3861(02)00731-0
late Phase-inversion Poly(vinylidene fluoride) Mem- [26] Smolders, C. A., Reuvers, A. J., Boom, R. M. and
branes,” Journal of Polymer Science Part B Polymer Wienk, I. M., “Microstructures in Phase-inversion
Physics, Vol. 41, No. 13, pp. 1578-1588 (2003). doi: Membranes. Part 1: Formation of Macrovoids,” Jour-
10.1002/polb.10513 nal of Membrane Science, Vol. 73, No. 2-3, pp. 259-
[19] Kuo, C. Y., Lin, H. N., Tsai, H. A., Wang, D. M. and 275 (1992). doi: 10.1016/0376-7388(92)80134-6
Lai, J. Y., “Fabrication of a High Hydrophobic PVDF [27] Strathmann, H., in Lloyd DR (editor), “Material Sci-
Membrane via Nonsolvent Induced Phase Separa- ence of Synthetic Membrane,” ACS Symposium Se-
tion,” Desalination, Vol. 233, No. 1-3, pp. 40-47 ries, Washington, DC: American Chemical Society,
(2008). doi: 10.1016/j.desal.2007.09.025 pp. 165-195 (1985).
[20] Chang, H. H., Chen, S. C., Lin, D. J. and Cheng, L. P., [28] McKelvey, S. A. and Koros, W. J., “Phase Separation,
“Preparation of Bi-continuous Nylon-66 Porous Vitrification, and the Manifestation of Macrovoids in
Membranes by Coagulation of Incipient Dopes in Polymeric Asymmetric Membranes,” Journal of Mem-
Soft Non-solvent Baths,” Desalination, Vol. 313, No. brane Science, Vol. 112, No. 1, pp. 29-39 (1996). doi:
15, pp. 77-86 (2013). doi: 10.1016/j.desal.2012.12. 10.1016/0376-7388(95)00197-2
009 [29] Ray, R. J., Krantz, W. B. and Sani, R. L., “Linear Sta-
[21] Buonomenna, M. G., Macchi, P., Davoli, M. and bility Theory Model for Finger Formation in Asym-
Drioli, E., “Poly(vinylidene fluoride) Membranes by metric Membranes,” Journal of Membrane Science,
Phase Inversion: the Role the Casting and Coagulation Vol. 23, No. 2, pp. 155-182 (1985). doi: 10.1016/
Conditions Play in Their Morphology, Crystalline St- S0376-7388(00)82216-4
ructure and Properties,” European Polymer Journal, [30] Termonia, Y., “Fundamentals of Polymer Coagula-
Vol. 43, No. 4, pp. 1557-1572 (2007). doi: 10.1016/ tion,” Journal of Polymer Science Part B Polymer
j.eurpolymj.2006.12.033 Physics, Vol. 33, No. 2, pp. 279-288 (1995). doi:
98 Hsu-Hsien Chang et al.

10.1002/polb.1995.090330213 31-43 (1999). doi: 10.1016/S0376-7388(98)00292-0


[31] Lai, J. Y., Lin, F. C., Wu, T. T. and Wang, D. M., “On
the Formation of Macrovoids in PMMA Membranes,” Manuscript Received: Mar. 9, 2017
Journal of Membrane Science, Vol. 155, No. 1, pp. Accepted: Jul. 30, 2017

You might also like