Você está na página 1de 6

Journal of Membrane Science 325 (2008) 92–97

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Correlation between macrovoid formation and the ternary phase diagram for
polyethersulfone membranes prepared from two nearly similar solvents
Jalal Barzin ∗ , Behrouz Sadatnia
Department of Biomaterials, Iran Polymer and Petrochemical Institute, P.O. Box 14965/115, Tehran, Iran

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

Article history: The morphological structure of membranes prepared from two nearly similar systems consisting
Received 5 March 2008 of water/N,N-dimethylacetamide (DMAc)/polyethersulfone (PES) and water/N-methyl-2-pyrrolidone
Received in revised form 17 June 2008 (NMP)/polyethersulfone (PES) has been studied. The morphology of the prepared membranes showed that
Accepted 2 July 2008
both systems exhibit an instantaneous liquid–liquid demixing that leads to the formation of macrovoids
Available online 10 July 2008
in the resulting structures. Nevertheless, the resulting macrovoid structures were contrary to the gen-
erally accepted concepts concerning macrovoid formation. The membranes morphologies showed that
Keywords:
in spite of better miscibility between water and DMAc, which must promote the formation of channel-
Macrovoid
Morphology
and finger-like structures, more sponge-like structures were observed in membranes prepared from the
Phase diagram water/DMAc/PES system compared to those prepared from the water/NMP/PES system. To find the source
Vitrification of this unexpected phenomenon, the complete ternary phase diagrams consisting of theoretical binodal
Polyethersulfone curves, vitrification boundaries, and gelation boundaries were constructed for both systems and it was
shown that gelation process occurs earlier in the water/DMAc/PES system compared to the other system,
which inhibits the growth of macrovoids in this system.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction Membranes that are prepared by the NIPS process usually con-
tain large elongated voids (macrovoids) in their sublayer structure
Today most of the industrially used polymeric membranes are [3]. The type and structure of these macrovoids has a large effect on
prepared via the phase separation (phase inversion) method. In this the performance and mechanical strength of the resulting mem-
process, a homogeneous polymer solution is separated into two branes and several studies have been performed to discover the
phases: a solid, polymer-rich phase that forms the matrix of the mechanism of macrovoid formation and to determine the effect of
membrane and a liquid, polymer-lean phase that forms the mem- various parameters on the structure of the resulting macrovoids
brane pores [1]. The phase separation process can be accomplished [3–32].
by a wide variety of techniques such as nonsolvent induced phase Usually, macrovoids occur under instantaneous demixing but, in
separation (NIPS), thermally induced phase separation (TIPS), vapor delayed demixing, the formation of macrovoids is suppressed and
induced phase separation (VIPS), and evaporation induced phase sponge-like structures are observed in the membrane structure [3].
separation (EIPS) [2]. Among these techniques, the NIPS process is Experimental results have shown that increasing the concentration
without doubt the most widely used technique for the preparation of polymer in the casting solution [4,5,12,16,19,28,30], increas-
of polymeric membranes. In the NIPS process, a homogenous poly- ing the evaporation time [5,18,19], choosing nonsolvent–solvent
mer solution is cast as a thin film or spun as a hollow fiber and then pairs with low mutual affinity [5,7,13], addition of solvent to the
immersed in a coagulation bath (typically water). The exchange of coagulation bath [6,7,9,21], and addition of nonsolvent to the poly-
nonsolvent and solvent induces the phase separation phenomena mer solution [3,5,12,13,16–18,21,23,24] suppress the formation of
in the polymer solution, which usually leads to the formation of macrovoids in membranes prepared by the NIPS process.
an anisotropic (asymmetric) structure with a relatively thin skin One of the most efficient ways to obtain different morpholog-
supported on a much thicker substrate in the resulting membrane. ical structures in membranes prepared by the NIPS process is to
use different solvents with varying mutual affinity with the non-
solvent present in the coagulation bath. Nevertheless, generally,
in these types of studies the morphological structures obtained
∗ Corresponding author. Tel.: +98 21 44580050; fax: +98 21 44580161. from a nonsolvent–solvent pair which has high mutual affin-
E-mail address: j.barzin@ippi.ac.ir (J. Barzin). ity and induces an instantaneous demixing is compared with

0376-7388/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2008.07.003
J. Barzin, B. Sadatnia / Journal of Membrane Science 325 (2008) 92–97 93

those obtained from a nonsolvent–solvent pair which has low Qian and Rudin [39]. The nonsolvent/polymer interaction parame-
mutual affinity and induces the delayed demixing [25,33,34]. ter, g13 , was determined by a swelling measurement according to
For example, Shieh and Chung [33] have studied the morphol- Flory–Rehner theory [38]. Details of the required procedures for
ogy of membranes prepared from water/N-methyl-2-pyrrolidone determination of interaction parameters and numerical calcula-
(NMP)/cellulose acetate (CA) and water/tetrahydrofuran (THF)/CA tions for construction of theoretical phase diagrams are presented
systems and by adjusting the ratio of NMP and THF in the elsewhere [40].
water/NMP-THF/CA system, they succeeded in suppressing the for-
mation of macrovoids in the resulting membranes. Li et al. [34] 2.3. Membrane preparation
have studied the morphological structure of water/NMP/CA and
water/␥-butyrolactone (GBL)/CA systems and reported that when Polymer solutions of 10, 12, 18, and 20 wt% of PES in DMAc and
ratio of NMP to GBL was increased to 80% in the water/NMP-GBL/CA NMP were cast at room temperature on glass plates with a uniform
systems, the macrovoids were eliminated and typical sponge-like thickness of 120 ␮m by a casting knife. Immediately after casting,
structures were observed in the resulting membranes. the glass plates were immersed in a coagulation bath containing
Kim et al. [35] have investigated the phase behavior of distilled water at 25 ◦ C. After polymer coagulation, the prepared
water/NMP/polyimide and water/DMSO/polyimide systems and membranes were transferred to a second bath containing fresh
reported that despite of higher mutual affinity between water water and then they were dried in air.
and DMSO compared to that of water and NMP, macrovoid-free,
sponge-like membrane was prepared from water/DMSO/polyimide
system, whereas a finger-like membrane was obtained from a 2.4. Scanning electron microscopy analysis
water/NMP/polyimide system.
In the present study, the morphological structure of two nearly The morphology of the membranes was studied by scanning
similar membrane-forming systems consisting of water/N,N- electron microscopy (Philips XL 30). Before analysis, the samples
dimethylacetamide (DMAc)/polyethersulfone (PES) and water/N- were fractured in liquid nitrogen, coated with gold in a sputtering
methyl-2-pyrrolidone (NMP)/polyethersulfone (PES) has been device (SCD 005, BAL-TEC), and then transferred into the micro-
evaluated. Both of these systems exhibit an instantaneous scope chamber.
liquid–liquid demixing that leads to the formation of macrovoids
in the resulting membranes. Nevertheless, some unexpected struc- 3. Results and discussion
tures were observed in the structure of the prepared membranes.
To find an explanation for this unusual phenomenon, the complete 3.1. Interaction parameters
ternary phase diagrams of these systems were constructed theo-
retically and according to the mechanism of gelation of amorphous In theoretical studies on ternary membrane-forming systems for
polymers the source of these observations has been explained. To the construction of theoretical ternary phase diagrams, the non-
the best of our knowledge, there is no report concerning evalua- solvent/solvent interaction parameter, g12 , is usually considered
tion of the effects of two nearly similar and strong solvents on the concentration dependent. There are several reports concerning g12
structures of membranes prepared from the NIPS process in the values of water/DMAc and water/NMP systems in the literature
literature. but, these reports are not in accordance with each other and some
inconsistencies are observed in the reported values and relations
[41–46].
2. Experimental
In Fig. 1 the g12 values which have been used in our numerical
calculations are presented for water/DMAc and water/NMP sys-
2.1. Materials
tems. According to the results of numerical calculations presented
in the following sections, it has been shown that these two set
Polyethersulfone (PES) Ultrason E 6020 P (BASF, Germany)
of g12 values result in theoretical binodal curves that are in good
was used as the polymer. This polymer has Mw = 58,000 g/mol
agreement with the experimental cloud point data.
and Mn = 16,000 g/mol as reported by the manufacturer. N,N-
dimethylacetamide (DMAc) (Merck) and N-methyl-2-pyrrolidone
(NMP) (Riedel-de-Haën) were used as solvent and double-distilled
water was used as the nonsolvent.

2.2. Construction of ternary phase diagrams

The theoretical ternary phase diagrams for water/DMAc/PES and


water/NMP/PES systems were constructed based on an extended
form of Flory–Huggins theory of polymer solutions [36,37]. Accord-
ing to this theory, the binodal and spinodal curves and tie lines of
these systems were determined by numerical calculations based
on the Newton–Raphson method.
Binary interaction parameters are the main input parameters
of the Flory–Huggins theory. The nonsolvent/solvent interac-
tion parameters, g12 , were determined from excess Gibbs free
energy, GE , data. GE can be determined from activity coefficients,
vapor–liquid equilibrium data, or it can be obtained directly from
data that are available in the literature [38]. The solvent/polymer
interaction parameters, g23 , were determined through measure- Fig. 1. Concentration-dependent nonsolvent/solvent interaction parameter of
ment of the intrinsic viscosity based on a model proposed by water/DMAc (䊉) and water/NMP () systems.
94 J. Barzin, B. Sadatnia / Journal of Membrane Science 325 (2008) 92–97

desired polymer solutions based on Qian and Rudin model [39]. For
NMP/PES and DMAc/PES pairs this parameter was determined to be
0.37 and 0.39, respectively. These values show that the interaction
parameter of NMP/PES pair is lower than that of DMAc/PES system
which confirms that NMP is a better solvent for PES compared to
DMAc.
The nonsolvent/polymer interaction parameter, 13 , was deter-
mined by the equilibrium swelling measurement. This method is
perhaps the most widely used technique for determination of this
parameter. These measurements result in a value of 2.83 for 13
of water/PES system. In spite of its widespread popularity, the
equilibrium swelling measurement results in 13 values that are
somewhat larger than the real values for this parameter [40]. To
find a reasonable value for this parameter, the theoretical bin-
odal curves were fitted on the experimental cloud point data for
both water/DMAc/PES and water/NMP/PES systems. This process
resulted in a value of 1.6 for this parameter and therefore this value
of 13 were used for the construction of theoretical ternary phase
Fig. 2. Theoretical binodal (—) and spinodal (– –) curves for water/DMAc/PES and diagrams in the following sections.
water/NMP/PES systems. To verify the theoretical calculations, the experimental
cloud point data (䊉) are presented as well.
3.2. Phase diagram of water/DMAc/PES and water/NMP/PES
systems
For the water/NMP system, the g12 values were determined
through GE data reported by Zeman and Tkacik [44] and for the In Fig. 2 the theoretical binodal and spinodal curves for the
water/DMAc system, these values were determined through GE water/DMAc/PES and water/NMP/PES systems are presented. These
data which have been determined by activity coefficients calculated curves were constructed by numerical calculations based on an
from the van Laar equation [43]. The required relation for determi- extended form of Flory–Huggins theory of polymer solutions
nation of g12 values through GE data is available in the literature [36,37] by the determined binary interaction parameters as input
[38]. data. To verify the theoretical calculations, the results of experimen-
It can be seen from Fig. 1 that in all compositions the g12 values tal cloud point data, which have been determined by the titration
of water/DMAc system are smaller than those of the water/NMP method [40], are also presented in Fig. 2.
system, which indicates that there is a better miscibility between This figure shows that the theoretical binodal curve of
water and DMAc compared to water and NMP system. water/DMAc/PES system is closer to the polymer–solvent axis com-
For the solvent/polymer interaction parameters, g23 , and non- pared to that of water/NMP/PES system; therefore less water is
solvent/polymer interaction parameter, g13 , no concentration needed for the precipitation of PES in the former system com-
dependencies were considered; therefore, these parameters have pared to the latter one. It is widely accepted that higher values
been shown with the symbol of  in the preceding sections. of g12 and lower values of g23 shift the binodal curve toward the
The solvent/polymer interaction parameters, 23 , were deter- nonsolvent–polymer axis and produce larger homogenous regions
mined through measurement of the intrinsic viscosity of the in the phase diagram of a ternary system [40,47].

Fig. 3. Micrographs of the cross-sections of membranes prepared from NMP/PES solutions coagulated in water. The compositions of polymer solutions were (a) 10, (b) 12,
(c) 18, and (d) 20 wt% PES in NMP.
J. Barzin, B. Sadatnia / Journal of Membrane Science 325 (2008) 92–97 95

Fig. 4. Micrographs of the cross-sections of membranes prepared from DMAc/PES solutions coagulated in water. The compositions of polymer solutions were (a) 10, (b) 12,
(c) 18, and (d) 20 wt% PES in DMAc.

The interaction parameter values which have been determined solution viscosity, suppresses the formation of macrovoids and
for these two systems show that water/NMP/PES system has higher causes long, open channel-like macrovoids turn to finger- and
g12 and lower 23 values compared to water/DMAc/PES system. tear-like structures with more sponge-like areas in the resulting
These two factors cause the binodal curve of the latter system be membranes [4,5,12,16,19,28,30].
closer to the polymer–solvent axis. When the morphological structures of membranes prepared
from these two nearly similar solvents were compared with each
3.3. Membrane morphology other, the most important finding of this study was revealed. A
comparison between the morphologies in Figs. 3 and 4 shows that
In Figs. 3 and 4 the morphology of the membranes prepared from in all polymer concentrations, especially in high polymer concen-
water/NMP/PES and water/DMAc/PES systems are shown. These trations, more sponge-like structures are observed in membranes
figures show that macrovoids are observed in the sublayer of all prepared from DMAc compared to those prepared from NMP. This
the prepared membranes. The presence of macrovoids proves that observation is contrary to the generally accepted concepts regard-
there is a high mutual affinity between water and the solvents in ing macrovoid formation and suppression. It is widely accepted that
these systems as it has been shown in the nonsolvent/solvent inter- a low tendency of mixing between nonsolvent–solvent pair sup-
action parameters of water/NMP and water/DMAc pairs (Fig. 1). presses the formation of macrovoids [5,7,13]. The mixing tendency
This high mutual affinity between nonsolvent and solvent induces between the nonsolvent–solvent pair is reflected in the nonsol-
an instantaneous demixing in the aforementioned systems that vent/solvent interaction parameter of these components. Materials
leads to the formation of macrovoids in the resulting membranes with low g12 values have a higher tendency to mix with each
structure. other and materials with high g12 values have a lower tendency
A look at different morphologies of Fig. 3 shows that channel- to mix. As it has been shown in Fig. 1, in all compositions the
like structures with open ends are observed in membrane prepared g12 values of the water/NMP system are higher than those of the
from a solution with a polymer concentration of 10 wt% (Fig. 3a). water/DMAc system, which proves that there is a higher mutual
Membranes with these irregular morphological structures are not affinity between water and DMAc in comparison to water and
suitable membranes regarding mechanical strength and handling NMP. Therefore, according to the concept of macrovoid suppres-
[48]. As the polymer concentration increases to 12 wt%, the pores sion in nonsolvent–solvent pairs with lower tendency of mixing,
structure changes to regular channels with closed ends (Fig. 3b). more sponge-like structures with smaller fingers and tears must
At a concentration of 18 wt%, the size of the channels decreases be observed in membranes prepared from water/NMP/PES system
and more sponge-like structures are observed in the membrane but, completely opposite results are observed in Figs. 3 and 4.
structure (Fig. 3c). In high concentration of 20 wt%, the structure To find an explanation for these unusual observations, it’s
changes again and small channel-, finger-, and tear-like structures necessary that mechanisms of physical gelation of amorphous
are observed in the membrane structure (Fig. 3d). polymers in presence of a solvent and nonsolvent are considered.
In membranes prepared from DMAc (Fig. 4), a similar trend can According to the gelation mechanism proposed by Arnauts and
be observed. Albeit, in this case more sponge-like structures are Berghmans [49], gelation of an amorphous polymer in a solvent
observed in the membrane structures compared to the correspond- results from liquid–liquid phase separation arrested by vitrifica-
ing concentrations in the NMP/PES membranes. tion of the polymer-rich phase. Li et al. [50] extended this concept
As a general rule, in membranes prepared from nonsol- of gelation to a ternary system consisting of a nonsolvent, a solvent,
vent/solvent/polymer systems, increasing the polymer concentra- and an amorphous polymer. According to this extended concept,
tion in the casting solution, which leads to an increase in the the structure of a phase-separated polymer solution is set when
96 J. Barzin, B. Sadatnia / Journal of Membrane Science 325 (2008) 92–97

mass transfer of solvent into the nonsolvent bath and the transfer of
nonsolvent into the casting solution. This phenomenon suppresses
the formation of a significant supersaturated region near the inter-
face, which Pekny et al. claim is essential for promoting macrovoid
growth.

4. Conclusions

Morphological studies on membranes prepared from


water/DMAc/PES and water/NMP/PES systems showed that
both systems exhibit an instantaneous demixing that leads to the
formation of macrovoids in the resulting membranes. Neverthe-
less, in spite of higher mutual affinity between water and DMAc,
it was observed that membranes prepared from DMAc had more
sponge-like structures compared to those prepared from NMP. It is
contrary to this widely accepted concept that nonsolvent–solvent
pairs with lower tendency of mixing with each other suppress the
formation of macrovoids. To find an explanation for this unex-
Fig. 5. Theoretical binodal curve, vitrification boundary, and gelation boundary for pected observation, the vitrification lines and theoretical binodal
water/DMAc/PES (—) and water/NMP/PES (– –) systems. curves and gelation boundaries of these systems were determined
and compared with each other. It was shown that despite higher
mutual affinity between water and DMAc compared to water
the polymer-rich phase of demixed polymer solution reaches its and NMP, the vitrification boundary in the former case intersects
glass state (i.e. vitrifies). In a ternary phase diagram, a line that the binodal of this system in lower polymer concentration com-
separates the one-phase solution region from the one-phase glass pared to the latter case. This phenomenon leads to the earlier
region is the vitrification boundary [51]. The intersection of the vitrification of the polymer-rich phase of the water/DMAc/PES
vitrification temperature curve and binodal curve is called Bergh- system compared to that of the water/NMP/PES system, which
mans’ point and the tie line that crosses through the Berghmans’ suppresses the formation of macrovoids in the former system.
point determines the gelation boundary. Li et al. [50] have per- The results of this study show that in membrane-forming systems
formed comprehensive studies concerning the physical gelation of consisting two nearly similar solvents, in addition to the extent
PES in a mixture of different solvents and water systems. Accord- of nonsolvent–solvent pairs affinity, the position of vitrification
ing to the differential scanning calorimetry (DSC) measurements, and gelation boundaries play an important role in determining the
these authors determined the polymer concentration at which the type of macrovoid structures in the resulting membranes.
Tg is equal to the system temperature (25 ◦ C) for DMAc/PES and
NMP/PES systems. These polymer concentrations can be used to Acknowledgement
determine the vitrification boundary in a ternary phase diagram.
In the present work we have shown the results of our The authors gratefully acknowledge support of this research by
numerical calculations regarding theoretical binodal curves for Iran National Science Foundation (INSF) grant 84122.
water/DMAc/PES and water/NMP/PES systems in Fig. 2. Based on
these calculations and data reported by Li et al. we have suc-
References
ceeded to construct complete ternary phase diagrams consisting
of the theoretical binodal curve, vitrification boundary, and gela- [1] R.W. Baker, Membrane Technology and Applications, John Wiley & Sons Ltd.,
tion boundary for water/DMAC/PES and water/NMP/PES systems England, 2004.
[2] M. Ulbricht, Advanced functional polymer membranes, Polymer 47 (2006) 2217.
and the results are presented in Fig. 5.
[3] C.A. Smolders, A.J. Reuvers, R.M. Boom, I.M. Wienk, Microstructures in phase-
According to this figure, a reasonable explanation for the unex- inversion membranes. Part 1. Formation of macrovoids, J. Membr. Sci. 73 (1992)
pected structures observed in Figs. 3 and 4 can be presented. As 259.
this figure shows the polymer-rich phase concentration at the [4] R. Matz, The structure of cellulose acetate membranes. 1. The development of
porous structures in anisotropic membranes, Desalination 10 (1972) 1.
intersection of binodal curve with the vitrification boundary is [5] M.A. Frommer, R.M. Messalem, Mechanism of membrane formation. VI. Con-
lower in the case of water/DMAc/PES system. On the other hand, vective flows and large void formation during membrane precipitation, Ind.
the binodal curve of this system is closer to the solvent–polymer Eng. Chem. Prod. Res. Dev. 12 (1973) 328.
[6] H. Strathmann, K. Kock, P. Amar, R.W. Baker, The formation mechanism of asym-
axis. Therefore, the vitrification curve of the water/DMAc/PES sys- metric membranes, Desalination 16 (1975) 179.
tem intersects the binodal curve of this system at lower polymer [7] H. Strathmann, K. Kock, The formation mechanism of phase inversion mem-
concentration which results in the gelation boundary of the sys- branes, Desalination 21 (1977) 241.
[8] I. Cabasso, E. Klein, J.K. Smith, Polysulfone hollow fibers. II. Morphology, J. Appl.
tem being closer to the solvent–polymer axis compared to the Polym. Sci. 21 (1977) 165.
water/NMP/PES system. Therefore, after liquid–liquid demixing the [9] D.M. Koenhen, M.H.V. Mulder, C.A. Smolders, Phase separation phenomena dur-
polymer-rich phase of the water/DMAc/PES system vitrifies earlier ing the formation of asymmetric membranes, J. Appl. Polym. Sci. 21 (1977)
199.
than that of water/NMP/PES system; this phenomenon suppresses
[10] C. Cohen, G.B. Tanny, S. Prager, Diffusion-controlled formation of porous struc-
the growth of macrovoids in the former system and membranes tures in ternary polymer systems, J. Polym. Sci. Polym. Phys. Ed. 17 (1979)
with more sponge-like structures are observed in this case. Recently 477.
[11] L. Broens, F.W. Altena, C.A. Smolders, D.M. Koenhen, Asymmetric membrane
Pekny et al. [26] have studied macrovoid development in the
structures as a result of phase separation phenomena, Desalination 32 (1980)
water/acetone/CA solutions and hypothesized that macrovoid for- 33.
mation requires a significant supersaturated region exist within the [12] C. Friedrich, A. Driancourt, C. Noel, L. Monnerie, Asymmetric reverse osmosis
casting solution. The earlier vitrification phenomenon observed in and ultrafiltration membranes prepared from sulfonated polysulfone, Desali-
nation 36 (1981) 39.
the water/DMAc/PES system studied in the present work causes [13] P. Neogi, Mechanism of pore formation in reverse osmosis membranes during
solidification of the interface and therefore greatly reduces the the casting process, AIChE J. 29 (1983) 402.
J. Barzin, B. Sadatnia / Journal of Membrane Science 325 (2008) 92–97 97

[14] A.J. Reuvers, J.W.A. van den Berg, C.A. Smolders, Formation of membranes by [33] J.J. Shieh, T.S. Chung, Effect of liquid–liquid demixing on the membrane mor-
means of immersion precipitation. Part I. A model to describe mass transfer phology, gas permeation, thermal and mechanical properties of cellulose
during immersion precipitation, J. Membr. Sci. 34 (1987) 45. acetate hollow fibers, J. Membr. Sci. 140 (1998) 67.
[15] C.S. Tsay, A.J. McHugh, Mass transfer modeling of asymmetric membrane for- [34] Z. Li, J. Ren, A.G. Fane, D.F. Li, F.S. Wong, Influence of solvent on the structure
mation by phase inversion, J. Polym. Sci., Part B: Polym. Phys. 28 (1990) 1327. and performance of cellulose acetate membranes, J. Membr. Sci. 279 (2006)
[16] A. Bottino, G. Camera-Roda, G. Capannelli, S. Munari, The formation of microp- 601.
orous polyvinylidene difluoride membranes by phase separation, J. Membr. Sci. [35] J.H. Kim, B.R. Min, J. Won, H.C. Park, Y.S. Kang, Phase behavior and mechanism
57 (1991) 1. of membrane formation for polyimide/DMSO/water system, J. Membr. Sci. 187
[17] Y.S. Kang, H.J. Kim, U.Y. Kim, Asymmetric membrane formation via immersion (2001) 47.
precipitation method. I. Kinetic effect, J. Membr. Sci. 60 (1991) 219. [36] P.J. Flory, Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY,
[18] L. Zeman, T. Fraser, Formation of air-cast cellulose acetate membranes. Part I. 1953.
Study of macrovoid formation, J. Membr. Sci. 84 (1993) 93. [37] H. Tompa, Polymer Solutions, Butterworths, London, 1956.
[19] F.G. Paulsen, S.S. Shojaie, W.B. Krantz, Effect of evaporation step on macrovoid [38] M.H.V. Mulder, C.A. Smolders, On the mechanism of separation of ethanol/water
formation in wet-cast polymeric membranes, J. Membr. Sci. 91 (1994) 265. mixtures by pervaporation. I. Calculations of concentration profiles, J. Membr.
[20] Y. Termonia, Fundamentals of polymer coagulation, J. Polym. Sci., Part B: Polym. Sci. 17 (1984) 289.
Phys. 33 (1995) 279. [39] J.W. Qian, A. Rudin, Prediction of thermodynamic properties of polymer solu-
[21] J.M. Cheng, D.M. Wang, F.C. Lin, J.Y. Lai, Formation and gas flux of asymmetric tions, Eur. Polym. J. 28 (1992) 725.
PMMA membranes, J. Membr. Sci. 109 (1996) 93. [40] J. Barzin, B. Sadatnia, Theoretical phase diagram calculation and membrane
[22] S.A. McKelvey, W.J. Koros, Phase separation, vitrification, and the manifestation morphology evaluation for water/solvent/polyethersulfone systems, Polymer
of macrovoids in polymeric asymmetric membranes, J. Membr. Sci. 112 (1996) 48 (2007) 1620.
29. [41] L. Shuguang, J. Chengzhang, Z. Yuanqi, The investigation of solution thermo-
[23] F.C. Lin, D.M. Wang, C.L. Lai, J.Y. Lai, Effect of surfactants on the structure of dynamics for the polysufone–DMAc–water system, Desalination 62 (1987)
PMMA membranes, J. Membr. Sci. 123 (1997) 281. 79.
[24] D.M. Wang, F.C. Lin, T.T. Wu, J.Y. Lai, Formation mechanism of the macrovoids [42] G.E. Gaides, A.J. McHugh, Gelation in an amorphous polymer: a discussion of
induced by surfactant additives, J. Membr. Sci. 142 (1998) 191. its relation to membrane formation, Polymer 30 (1989) 2118.
[25] W. Albrecht, Th. Weigel, M. Schossig-Tiedemann, K. Kneifel, K.V. Peinemann, [43] S.C. Pesek, W.J. Koros, Aqueous quenched asymmetric polysulfone membranes
D. Paul, Formation of hollow fiber membranes from poly(ether imide) at wet prepared by dry/wet phase separation, J. Membr. Sci. 81 (1993) 71.
phase inversion using binary mixtures of solvents for the preparation of the [44] L. Zeman, G. Tkacik, Thermodynamic analysis of a membrane-forming sys-
dope, J. Membr. Sci. 192 (2001) 217. tem water/N-methyl-2-pyrrolidone/polyethersulfone, J. Membr. Sci. 36 (1988)
[26] M.R. Pekny, J. Zartman, W.B. Krantz, A.R. Greenberg, P. Todd, Flow-visualization 119.
during macrovoid pore formation in dry-cast cellulose acetate membranes, J. [45] D.J. Lin, C.L. Chang, C.L. Chang, T.C. Chen, L.P. Cheng, Fine structure of
Membr. Sci. 211 (2003) 71. poly(vinylidene fluoride) membranes prepared by phase inversion from a
[27] J. Barzin, C. Feng, K.C. Khulbe, T. Matsuura, S.S. Madaeni, H. Mirzadeh, Charac- water/N-methyl-2-pyrollidone/poly(vinylidene fluoride) system, J. Polym. Sci.,
terization of polyethersulfone hemodialysis membrane by ultrafiltration and Part B: Polym. Phys. 42 (2004) 830.
atomic force microscopy, J. Membr. Sci. 237 (2004) 77. [46] Y.M. Wei, Z.L. Xu, X.T. Yang, H.L. Liu, Mathematical calculation of binodal curves
[28] J. Barzin, S.S. Madaeni, H. Mirzadeh, M. Mehrabzadeh, Effect of polyvinylpyrroli- of a polymer/solvent/nonsolvent system in the phase inversion process, Desali-
done on morphology and performance of hemodialysis membranes prepared nation 192 (2006) 91.
from polyether sulfone, J. Appl. Polym. Sci. 92 (2004) 3804. [47] F.W. Altena, C.A. Smolders, Calculation of liquid–liquid phase separation in a
[29] J. Barzin, S.S. Madaeni, H. Mirzadeh, Effect of preparation conditions on ternary system of a polymer in a mixture of a solvent and a nonsolvent, Macro-
morphology and performance of hemodialysis membranes prepared from molecules 15 (1982) 1491.
polyether sulphone and polyvinylpyrrolidone, Iran. Polym. J. 14 (2005) [48] C. Barth, M.C. Goncalves, A.T.N. Pires, J. Roeder, B.A. Wolf, Asymmetric polysul-
353. fone and polyethersulfone membranes: effects of thermodynamic conditions
[30] S.S. Madaeni, A. Rahimpour, J. Barzin, Preparation of polysulphone ultrafiltra- during formation on their performance, J. Membr. Sci. 169 (2000) 287.
tion membranes for milk concentration: effect of additives on morphology and [49] J. Arnauts, H. Berghmans, Amorphous thermoreversible gels of atactic
performance, Iran. Polym. J. 14 (2005) 421. polystyrene, Polym. Commun. 28 (1987) 66.
[31] S.S. Madaeni, M. Esmaeili, J. Barzin, Preparation and optimization of [50] S.G. Li, Th. van den Boomgaard, C.A. Smolders, H. Strathmann, Physical gelation
polyethersulfone-based composite membranes for air separation at low pres- of amorphous polymers in a mixture of solvent and nonsolvent, Macro-
sures, Polym. Polym. Compos. 15 (2007) 579. molecules 29 (1996) 2053.
[32] J. Barzin, S.S. Madaeni, S. Pourmoghadasi, Hemodialysis membranes prepared [51] P. van de Witte, P.J. Dijkstra, J.W.A. van den Berg, J. Feijen, Phase separation
from poly(vinyl alcohol): effects of the preparation conditions on the morphol- processes in polymer solutions in relation to membrane formation, J. Membr.
ogy and performance, J. Appl. Polym. Sci. 104 (2007) 2490. Sci. 117 (1996) 1.

Você também pode gostar