You are on page 1of 11

Journal of Membrane Science 363 (2010) 149159

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Pervaporation dehydration of ethylene glycol through polybenzimidazole (PBI)-based membranes. 1. Membrane fabrication
Yan Wang a , Michael Gruender b , Tai Shung Chung a,
a b

Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, 10 Kent Ridge Cresent, Singapore 117576, Singapore PBI Performance Products, Inc., 9800-D Southern Pine Boulevard, Charlotte, NC 28273, USA

a r t i c l e

i n f o

a b s t r a c t
Ethylene glycol is an important commodity in chemical industries and dewatering is a critical process in the production and recycle of ethylene glycol. In this work, we have developed dual-layer polybenzimidazole/polyetherimide (PBI/PEI) hollow ber membranes for ethylene glycol dehydration via pervaporation. Three types of membranes have been prepared; namely, (1) PBI at dense membranes; (2) PBI single-layer hollow ber membranes; and (3) PBI/PEI dual-layer hollow ber membranes. PBI at dense membranes have the lowest separation performance due to severe swelling. PBI single-layer hollow ber membranes show better separation performance in terms of permeation ux and separation factor but have very low tensile strains. The dual-layer PBI/PEI hollow ber membranes have the best separation performance due to (1) unique combination of the superior physicochemical properties of the PBI selective layer and the less swelling characteristics of the PEI supporting layer, and (2) synergistic effects of molecularly designed membrane morphology via dual-layer co-extrusion. The effects of spinning parameters of PBI single-layer and PBI/PEI dual-layer hollow ber membranes on pervaporation performance have been investigated. A thermal treatment of PBI/PEI dual-layer hollow ber membranes at 75 C can signicantly enhance the separation performance. Compared with other polymeric membranes, the newly developed PBI/PEI dual-layer hollow ber membranes have much better separation factors and slightly lower uxes for the ethylene glycol dehydration. It is believed that the science and engineering of designing PBI/PEI duallayer hollow ber membranes with an ultra-thin functional separation layer and a synergic supporting layer may open new perspective for the development of next-generation high-performance multilayer membranes for liquid separations. 2010 Elsevier B.V. All rights reserved.

Article history: Received 18 May 2010 Received in revised form 12 July 2010 Accepted 14 July 2010 Available online 21 July 2010 Keywords: Pervaporation Polybenzimidazole Dual-layer hollow ber membrane Ethylene glycol Dewater

1. Introduction Ethylene glycol is an important chemical widely used as the non-volatile antifreeze and de-icing agent because of its low freezing point. It is also an important raw material for the manufacture of many industrial products [1,3]. In addition, ethylene glycol has been working as an ideal absorbent for natural gas dehydration due to its good afnity to water and high boiling point (197 C) [1]. All the aforementioned applications involve separation of water from ethylene glycol. Whats more, during the production of ethylene glycol, a direct oxidation of ethylene to ethylene oxide is followed by the hydrolysis of ethylene oxide, thus a large excess of water is used in the hydrolysis reaction to enhance the conversion. The excess water must be removed later to concentrate the product. Therefore, dewatering becomes a critical issue in the production and recycle process of ethylene glycol. Table 1 lists

Corresponding author. Tel.: +65 6516 6645; fax: +65 6779 1936. E-mail address: chencts@nus.edu.sg (T.S. Chung). 0376-7388/$ see front matter 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.memsci.2010.07.024

some physicochemical properties of ethylene glycol and water [24]. Currently, the ethylene glycol dehydration is done by conventional multi-stage evaporation units followed by distillation columns, characterized by high capital costs. Evaporation is suitable for concentrating ethylene glycol up to a concentration of 70 wt.%. However, beyond this point, the quality of the overhead product diminishes and the energy demand increases signicantly. Distillation can be used to further concentrate ethylene glycol beyond 70 wt.% and solve the quality problem, but still consumes a large amount of energy. Actually, the separation of ethylene glycolwater mixture is ranked the eighth most energy intensive distillation processes in the chemical industry because the high boiling point of ethylene glycol requires a high-pressure steam of the reboiler [5,6]. Therefore, more cost effective technologies are urgently needed in the chemical industry in order to stay competitive. Pervaporation is the most promising technology for molecularscale liquid/liquid separations existing in biorenery, petrochemical, and pharmaceutical industries because of its highly selective, economical, energy efcient and eco-friendly characteristics. Since

150

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159

Table 1 The physicochemical properties of ethylene glycol and water. Ethylene glycol Water

Chemical structure Formula CAS number Molar mass (g/mol) Density (g/cm3 ) Molecular volume (3 )a Appearance

Melting point ( C) Boiling point Polarity parameter ET (30) (kcal/mol) [2] Solubility parameter (MPa)1/2 [3] 17.0 d 11.0 p 26.0 h 33.0 sp Vapor pressure (mm Hg) 0.07 25 C 1.34 60 C

C2 H4 (OH)2 107-21-1 62.068 1.113 92.4 Odorless, colorless, syrupy liquid 12.9 197.3 56.3

H2 O 7732-18-5 18.015 0.998 29.9 Colorless, odorless liquid 0 100.0 63.1

15.5 16.0 42.3 47.8 23.76 149.44

a The molecular volume is calculated by the molecular weight divided by the density and the Avogadro number [4].

only part of the feed is vaporized and the remaining feed does not undergo phase change, the pervaporation process is much more energy efcient than the conventional distillation process. Hence, pervaporation has gradually gained acceptance in industries to solve the quality and energy problems in ethylene glycol dehydration in the past 20 years. In the case of pervaporation dehydration, signicant attention has been given to highly hydrophilic polymers in the earlier stage, such as poly(acrylic acid) (PAA), poly(vinyl alcohol) (PVA), polyacrylonitrile (PAN), sodium alginate, and chitosan, etc. [7]. However, these materials lack mechanical strength and stability in aqueous solutions due to the excessive swelling, leading to a drastic decrease in separation performance [8,9]. Cross-linking modications to depress the membrane swelling may not be very desirable since the permeation ux is often compromised. The additional treatments also incur extra costs and prolong production durations. Another feasible method is to employ the composite membrane comprised of a thin active layer and supported microporous substrate(s) made of less swelling material, where the outer surface layer provides the selectivity, while the porous substrate layer offers mechanical strength and high permeability. Composite membranes for pervaporation applications have been reported extensively in recent years. Ceramic-supported polymeric membranes have attracted much attention recently [1015] because of the unique chemical, mechanical and thermal stability of the ceramic substrates. Many polymeric composite membranes also showed much improved separation performance for ethanol dehydration [1619]. Especially, co-extrusion dual-layer hollow bers studied by our group as pervaporation membranes recently [2023] have shown promising pervaporation performance. Synergistic separation performance for organic dehydration can be achieved if inner- and outer-layer materials are properly selected even without intensive thermal or chemical treatment [2023]. Polybenzimidazole (PBI) is a high-performance aromatic polymeric material, suitable for many aggressive environments because of their outstanding chemical resistance, mechanical strength and thermal stability. It is a glassy polymer with high Tg (417 C) [24] with stable mechanical properties up to 350 C. In addition, PBI possesses both donor and acceptor hydrogen-bonding sites [25,26], which are capable of participating in specic interactions [27,28].

PBI is also known to absorb 15 wt.% water at equilibrium and the water in PBI is mobile [29]. Water can preferentially permeate the PBI membrane due to the strong water afnity with PBI molecules and a much smaller molecular size of water. All the above characteristics of PBI material make it a promising pervaporation membrane material for the dehydration of various organics. However, in spite of so many advantages, the application of PBI material as pervaporation membranes is very limited [21,30]. This is possibly because the solvent-induced phase-inversion PBI membrane is very brittle after drying, which makes it difcult to be fabricated into self-standing membranes. Another disadvantage is that the PBI membrane with extreme hydrophilicity would be easily swollen in aqueous solutions, resulting in a decreased membrane selectivity and unstable long-term performance as stated above. Therefore, the aims of this work are to (1) investigate the science and engineering of fabricating dual-layer hollow ber membranes with a PBI outer layer and an anti-swelling polyetherimide (PEI) inner layer for ethylene glycol dehydration; (2) examine the materials synergism in terms of their unique physicochemical properties (i.e., superior hydrophilicity of PBI, extremely low water up-take of PEI, and their miscibility); and (3) identify the effects of key spinning parameters on membrane morphology and separation performance. We also fabricated PBI at dense membranes and single-layer hollow bers for comparison. This study may potentially open up new perspectives for the development of next-generation pervaporation membranes with an ultra-thin functional selective layer. In the following associated paper (Part 2), we will discuss the effects of pervaporation operation conditions (operation temperature, feed concentration and vacuum pressure) on the separation performance, in order to investigate the intrinsic properties of the membrane (permeability and selectivity). 2. Experimental 2.1. Materials The PBI polymer solution was provided by PBI Performance Products, Inc. with a composition of 26.2 wt.% PBI, 72.3 wt.% N,Ndimethylacetimide (DMAc), and 1.5 wt.% lithium chloride (LiCl). The LiCl serves the function of preventing PBI from phasing out of the solution. Ultem 1010 polyetherimide was purchased from GE plastics. Fig. 1 shows the chemical structures of PBI and PEI [8,24]. Polyvinylpyrrolidone (PVP) (Merck, Singapore) with an average Mw of 30 kDa was employed as an additive for the hollow ber spinning. Polymers were dried overnight at 120 C under vacuum before use. DMAc, employed as the solvent in bore uid and for membrane preparation, was supplied by Merck with analytical grade and used as received. Ethylene glycol of analytical grade was used to mix with deionized water to prepare the binary feed solution. 2.2. The fabrication of PBI at-sheet dense membranes Flat-sheet PBI dense membranes were cast from a 15 wt.% PBI/DMAc polymer solution. The polymer dope solution of PBI/DMAc/LiCl (15.0/84.1/0.9 wt.%) was prepared by diluting the original supplied PBI solution. The membrane was prepared by casting the polymer solution onto a glass plate with a casting knife of a 250 m thickness, and then placed on a hot plate preset at 75 C for 15 h, to allow the solvent evaporated slowly. The resultant lm was dried with two different drying protocols as described below. A Mitutoyo micrometer was employed to measure the nal membrane thickness. The 1st protocol: The resultant lm together with glass plate was immersed in a water coagulation bath. After the membrane separating from the glass plate automatically within several minutes, it was continuously immersed in water for 2 days with fresh water

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159

151

Fig. 1. The chemical structures of (a) polybenzimidazole (PBI) and (b) Ultem 1010 polyetherimide (PEI).

changed daily. The wet membrane was then dried in a Freeze-dryer (Thermo Electron Co. ModulyoD-230) with vacuum overnight after frozen in a freezer for 2 h. Using this drying protocol, the LiCl may be removed from the as-fabricated PBI dense membrane. The 2nd protocol: The resultant lm was peeled off carefully from the glass plate and then dried between two wire meshes in a vacuum oven, with temperature gradually increasing to 250 C at a rate of 12 C/20 min and hold there for 24 h to remove the residual solvents before cooling down naturally. The wire meshes not only prevented the membrane from sticking to the glass plate but also helped uniformly dry the membrane from both surfaces. With this drying protocol, the LiCl remains in the as-fabricated PBI dense membrane. 2.3. Spinning process and modules fabrication of hollow ber membranes The schematic diagrams of single-layer and dual-layer hollow ber spinning systems have been described elsewhere [22]. The polymer solution was degassed for 24 h before loading into a syringe pump (ISCO 1000). A mixture of 85/15 (w/w) DMAc/water
Table 2 The spinning parameters of single-layer hollow ber membranes. Parameters Dope solution composition (wt.%) Bore uid composition (wt.%) Dimensions of spinneret (mm) (inner diameter/outer diameter) Temperature ( C) External coagulant Dope ow rate (ml/min) Air gap distance (cm) Membrane ID (PBI-S-) Bore uid ow rate (ml/min) Take-up speed (m/min)

was employed as the bore uid in order to make a porous inner surface. The hollow ber was spun by extrusion of the polymer solution and bore uid out of the spinneret orice and subsequent phase inversion in a coagulant bath with a pre-set air gap. Both dope uid and bore uid were ltered through 15 m sintered metal lters before spinning. Tap water was used as the external coagulant at room temperature. The nascent bers were rolled up by a drum, cut into segments, and then rinsed in a clean tapping water bath for at least 3 days to remove the remaining DMAc. The as-spun hollow bers were dried in air naturally after freeze-drying, and then stored in ambient environment. For single-layer hollow bers, the dope solution contains 23 wt.% PBI diluted from the original supplied PBI solution. Table 2 lists the detailed spinning parameters of the single-layer coextrusion process with various bore uid ow rates and take-up speeds. For dual-layer hollow bers, the outer-layer dope was a 23 wt.% PBI polymer solution, the same as the single-layer dope solution; while the inner-layer dope was a 25 wt.% PEI polymer solution consisting of 25/5/70 wt.% PEI/PVP/DMAc. Table 3 lists the detailed spinning parameters for the dual-layer co-extrusion process. The 23 wt.% PBI/DMAc solution was chosen for the outer layer

Range of variables PBI:DMAc:LiCl (23:75.67:1.33) DMAc:water (85:15) 1.05/1.60 Ambient (23 2) Water 2 2 A B C 1 1.25 1.5 5.15 5.15 5.15

D 2 5.15

E 1.5 4.60

F 1.5 9.59

G 1.5 16.24

H 1.5 21.79

Table 3 The spinning parameters of dual-layer hollow ber membranes. Parameters Outer-layer dope solution composition (wt.%) Inner-layer dope solution composition (wt.%) Bore uid composition (wt.%) Dimensions of spinneret (mm) External coagulant Temperature ( C) Outer-layer dope ow rate (ml/min) Inner-layer dope ow rate (ml/min) Bore uid ow rate (ml/min) Membrane ID (PBI-D-) Air gap distance (cm) Take-up speed (m/min) Range of variables PBI:DMAc:LiCl (23:75.67:1.33) PEI (Ultem 1010):PVP:DMAc (25:5:70) DMAc:water (85:15) OD1 /OD2 /ID (1.20/0.97/0.44) Water Ambient (23 2) 0.5 4 2 A B C 5 2 1 4.60 (free fall)

D 2 9.59

E 2 16.24

F 2 21.79

152

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159

have been described elsewhere [36]. A feed solution of 50/50 wt.% water/ethylene glycol was used for both pervaporation units unless otherwise stated. Since the feed concentration varies less than 0.5 wt.% during the entire experiment because of a large quantity of the feed solution compared to the permeate sample, it was considered to be constant. The operational temperature was 60 C. The feed ow rate was 1.38 l/min for the at-sheet dense membrane and 0.5 l/min for each hollow ber module. The permeate pressure was maintained less than 3 mbar by a vacuum pump. Retentate and permeate samples were collected after the membrane being conditioned for about 2 h. The ux J was determined by the mass of permeate divided by the product of the interval time and membrane area. The mass of permeate was weighed using a Mettler Toledo balance. The separation factor is dened by the equation below:
Fig. 2. Viscosity of PBI/DMAc solutions with different PBI concentrations.

yw,1 /yw,2 xw,1 /xw,2

(1)

because the critical concentration of PBI/DMAc solution was about 22.5 wt.% as shown in the viscosityconcentration curve in Fig. 2 provided by PBI [31(. Generally concentrations near or higher than the critical value were chosen for the preparation of asymmetric gas membranes to make sure a quasi-dense skin layer with less defects [3234] but the concept is also applicable to asymmetric pervaporation membranes [8,20,22,23]. The pervaporation modules were prepared by loading one piece of hollow ber membranes into a peruoroalkoxy tubing connecting with two Swagelok stainless steel male run tees with an effective length of around 20 cm. Both ends were sealed by epoxy and cured for 24 h at ambient temperature. At least two pervaporation modules were tested for each membrane sample. Thermal treatment of hollow bers was carried out before module fabrication if it was applied. All thermal treatments were conducted by immersing the hollow ber membranes in a hot water bath at 75 C for 2 h, followed by naturally cooling down in the water bath and air-drying after that. 2.4. Pervaporation study A static pervaporation cell made according to Prof. Matsuuras design [35] was used to test the pervaporation performance of the at-sheet membrane at room temperature. The design schematic is shown in Fig. 3. A testing membrane was placed in the stainless steel permeation cell with an effective surface area of 15.2 cm2 . For the characterization of hollow ber membranes, a laboratory scale pervaporation unit was employed and the details of the apparatus

where subscripts 1 and 2 refer to water and ethylene glycol, respectively; yw and xw are the weight fractions of component in the permeate and feed, and were analyzed through a Hewlett-Packard GC 7890 A with a HP-INNOWAX column (packed with cross-linked polyethylene glycol) and a TCD detector. 2.5. Membrane characterization The morphology of hollow ber membranes was observed using a JSM-6700F eld emission scanning electron microscope (FESEM). The hollow ber sample for SEM observation was prepared by fracturing the membrane strip in liquid nitrogen and then coating it with platinum. Photographs of the PBI dense membrane and hollow ber membrane after pervaporation were taken by conventional camera. The mechanical properties of hollow bers were tested using a tensile meter INSTRON 5542 and analyzed with the Bluehill 2 software. The tests were carried out at room temperature (25 C) and 80% relative humidity. Each hollow ber sample was clamped at the both ends with an initial gauge length of 50 mm and the test method involved stretching at a rate of 10 mm/min until failure. At least three samples were tested for each membrane. 3. Results and discussion 3.1. Pervaporation performance of the PBI dense membrane The at-sheet dense membrane is studied rstly to investigate the pervaporation performance of the neat PBI material for the ethylene glycol dehydration. The separation performance of PBI

Fig. 3. Schematic of the batch pervaporation separation system for at-sheet membranes.

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159 Table 4 Pervaporation performance of PBI dense membranes. Membrane Membrane 1 (freeze dried) Membrane 2 (thermal treated)
a b

153

Operation mode Normal mode Reverse modeb Normal modea Reverse modeb
a

Permeate (EG, wt.%) 14.04 10.08 12.10 7.05

Normalized total ux (g m/m2 h) 1232 1131 966 784

Separation factor (water/EG) 6.0 9.0 7.1 13.1

Membrane is mounted with the air-side facing the feed; Membrane is mounted with the glass-side facing the feed.

dense membranes prepared by different drying protocols and operated in different modes is shown in Table 4. The results show the thermal treated dense membrane is of higher separation factor but lower permeation ux, which is mainly due to a smaller free volume and the elimination of micro-defects in the thin separating layer, as well as a higher transportation resistance. Table 4 also compares the separation performance of dense membranes as a function of operation modes. The separation properties depend upon the membrane-mounting mode (i.e., using air-side or glass-side of the membrane against the feed during pervaporation tests). Both membranes exhibit much higher separation factors with somewhat lower uxes when the bottom side is placed facing against the feed. This phenomenon is consistent with previous studies [8,3739]. The possible causes are (1) the bottom surface has a higher hydrophilicity because the hydrophilic-end groups of PBI molecules tend to localize at the bottom toward the glass plate, and (2) a less swelling of the dense-selective layer in the reversed operation mode, as reported in our previous study [8]. Nevertheless, the pervaporation performance of at-sheet PBI dense membranes is very poor. The separation factor is around 10 and the normalized total ux is about 1 kg m/m2 h. Compared to polyimide materials investigated in our previous study [8], PBI dense membranes show an extreme low separation factor towards water but a higher ux. This poor performance may be due to the serious swelling in aqueous feed solutions because PBI has impressive water afnity [29]. In agreement with our hypothesis, severe swelling and wrinkles can be observed on the membranes after pervaporation tests as shown in Fig. 4. To lower the swelling problem, PBI hollow ber membranes are prepared in order to achieve an enhanced pervaporation performance as discussed in the following sections. 3.2. Pervaporation performance of single-layer PBI hollow ber membranes Table 5 shows the pervaporation performance of single-layer PBI hollow bers spun with various bore uid ow rates and takeup speeds for ethylene glycol dehydration. Compared with the PBI dense membranes, the single-layer PBI hollow bers show much improved separation performance in both permeation ux and separation factor. The substantial enhancement in ux arises from many factors. The conguration of asymmetric hollow bers is much better than that of at dense membranes because the former provides a larger surface area, less transport resistance, and

Fig. 4. Photograph of the PBI dense membrane after pervaporation.

Fig. 5. Photograph of the PBI single-layer hollow ber after pervaporation.

lower swelling than the latter. In addition, hollow bers have selfcontained vacuum channels where the feed is supplied from the shell side while vacuum is applied on the lumen side. The porous and dry substructure of asymmetric hollow bers attributes to a less swelling of the selective layer, thus achieve a higher separation factor. As shown in Fig. 5, no severe swelling phenomenon of the single-layer hollow ber membrane after pervaporation is observed. During the whole pervaporation process of about 810 h, there is no obvious decline in pervaporation performance. Table 5 shows the effects of bore uid rate and take-up speed on the pervaporation performance of single-layer PBI hollow ber membranes. These effects have been studied extensively on membranes for gas separation [4042] and morphology manipulations [43,44]. Not only the bore uid rate affects the demixing process in

Table 5 Pervaporation performance of PBI single-layer hollow ber membranes with different spinning parameters. Membrane ID PBI-S-A PBI-S-B PBI-S-C PBI-S-D PBI-S-E PBI-S-F PBI-S-G PBI-S-H Bore uid ow rate (ml/min) 1 1.25 1.5 2 1.5 1.5 1.5 1.5 Take-up speed (m/min) 5.15 5.15 5.15 5.15 4.60 (free fall) 9.59 16.24 21.79 Permeate (H2 O, wt.%) 97.14 98.99 99.07 98.57 99.13 98.81 99.02 99.12 Total ux (g/m2 h) 348 449 490 817 592 1314 1372 1147 Separation factor 35 100 109 71 105 89 105 116

154

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159

Fig. 6. SEM morphology of PBI single-layer hollow ber membranes with different bore uid ow rates.

the inner skin and its morphology, but also plays an important role on the hollow bers inner diameter and wall thickness, as well as the molecular orientation of the outer skin in the hoop direction. This is due to the fact that the solidication rate at the inner surface may increase with increasing bore uid rate since the mass transfer is facilitated. Therefore, an increase in bore uid rate results in an increase in inner diameter with a reduced wall thickness of the as-spun ber and a slightly stretched outer skin in the hoop direction. Fig. 6 illustrates the SEM morphologies of hollow bers PBI-S-A to D with increasing bore uid rate. Increased inner diameter and reduced wall thickness can be observed. Consistent with morphological changes, clear differences can be observed in their separation performance in terms of ux and separation factor as tabulated in Table 5. Increasing bore uid rate results in an increase in ux possibly because of a thinner wall. The separation factor shows an up-and-down trend possibly because of the enhanced molecular orientation in the hoop direction induced by the bore uid initially, but minor defects are created if the bore uid rate is continuously increased. The stress-induced orientation and upand-down trends in separation performance have been reported for various membranes [4446]. However, to our best knowledge, the bore uid induced molecular orientation in the hoop direction and its effects on separation performance have not been reported for pervaporation membranes. Table 5 also compares the effect of take-up speed on pervaporation performance for membranes made under a constant bore uid rate. The ux increases with increasing take-up speed, while the separation factor does not show obvious trend. Generally, the elongational or extensional stress upon the nascent ber is the pre-

dominant external stress on the spinning line [43,46,47]. A high take-up speed not only reduces hollow ber diameter and wall thickness (as shown in Fig. 7), but also induces chain orientation at the outer selective skin. The reduced wall thickness is the main reason contributing to the enhanced ux, while the enhanced orientation may attribute to the improved separation factor. However, an increase in take-up speed also causes over-stretched polymer chains to break at some weak points, offsetting the benets of the chain orientation, which leads to nearly unchanged separation factor. Fig. 7 shows the SEM morphology of PBI single-layer hollow ber membranes as a function of take-up speed and conrms that the higher the take-up speed, the denser the outer edge and the looser the inner surface. However, in spite of the enhanced pervaporation performance, the major drawback of the PBI single-layer hollow bers is its brittleness. For practical pervaporation applications, it is desired that the hollow ber has adequate mechanical strength so that it can withstand harsh operation environments over numerous cycles. Table 6 tabulates the mechanical properties of the PBI hollow bers spun from different conditions and shows the maximum strains of all single-layer hollow ber membranes are very low, indicating the poor exibility of these bers for module fabrication. 3.3. Pervaporation performance of the dual-layer PBI hollow ber membranes To enhance mechanical properties and further improve separation performance of hollow ber membranes, PBI/PEI dual-layer hollow ber membranes are further developed with the aid of PEI

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159

155

Fig. 7. SEM morphology of PBI single-layer hollow ber membranes with different take-up speeds.

as the inner supporting layer. The PEI material is chosen because of the following reasons: (a) PEI has been studied as a membrane material for vapor permeation and pervaporation [48], and showed preferential water selectivity over organic chemicals. (b) PEI has shown very low water adsorption (1.9 wt.%) [20]. It hardly suffers from swelling like most other polymers, which makes it especially attractive to work as an anti-swelling supporting layer in the design of dual-layer hollow ber membranes. (c) The PEI inner layer can provide the required mechanical support for the relative brittle PBI outer layer. (d) An important factor worthy of mention is that PBI and PEI materials are miscible at the molecular level [28,49], thus good adhesion between these two layers may be achievable.

Table 6 Mechanical strength of PBI single-layer hollow bers. Membrane ID PBI-S-A PBI-S-B PBI-S-C PBI-S-D PBI-S-E PBI-S-F PBI-S-G PBI-S-H Max tensile stress (MPa) 12.7 12.9 11.7 12.4 15.9 16.3 15.0 10.1 Youngs modulus (MPa) 1614 1532 1506 1488 1436 1229 1175 1277 Max strain (mm/mm) 0.011 0.013 0.010 0.010 0.013 0.019 0.019 0.010

Fig. 8 shows the SEM morphology of the PBI/PEI dual-layer hollow ber spun from an air gap of 2 cm and a take-up speed of 4.60 m/min (PBI-D-B). The hollow ber has a diameter of about 1250 m. The dual-layer wall thickness is about 240 m and the outer layer thickness is about 16 m. Both inner layer and outer layer have asymmetric cross-section morphology. No delamination can be observed at the interface between the two layers, indicating a good adhesion. The outer surface of the outer layer reveals a dense structure at a high magnication (50,000), which is a defect-less selective layer, while the inner surface of the inner layer is very porous which is desirable since it does not constitute much transport resistance. The interfacial morphology is a unique feature in dual-layer composite membranes. A seamless interface may be obtained depending on the miscibility of both dopes and many other factors during phase inversion. If the solvents, non-solvents, polymers, and additives used in both dopes are thermodynamically compatible, both dopes may diffuse into each other before and during phase inversion because of chemical potential differences [23,50]. The formation of the desired interface between PEI and PBI layers in this study can be attributed to several factors: (1) the miscibility between these two polymers as proved by previous studies [28,49]; (2) inter-penetration occurs between these two dopes because of the interfacial diffusions and convections driven by chemical potential differences; (3) a common solvent (DMAc) used in the spinning solutions of both layers. Table 7 summarizes the pervaporation performance of duallayer PBI/PEI hollow ber membranes spun with different spinning parameters for ethylene glycol dehydration. Compared with singlelayer PBI hollow ber membranes, the dual-layer hollow ber

156

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159

Fig. 8. FESEM images of the membrane morphology of PBI/PEI dual-layer hollow bers (PBI-D-B).

membranes have overall much higher separation factors and comparable uxes. This improvement may arise from three factors: (1) the low swelling characteristics of the PEI inner layer; (2) the porous and dry status of the PEI inner layer; (3) the decoupling of shear and elongational stresses in the dual-layer hollow ber fabrication. The observation of the dual-layer PBI/PEI hollow ber membranes after pervaporation conrms our 1st and 2nd hypotheses because the bers remain their original straight shapes, indicating no severe swelling of each individual hollow ber membrane. Table 7 also shows us the effects of spinning parameters (i.e., air gap and take-up speed) on the pervaporation performance of dual-layer PBI/PEI hollow ber membranes. The effects of air gap on gas separation performance of hollow bers have been well studied [20,22,46,51] and can be summarized as follows: (1) a longer air gap distance may increase solvent evaporation, causing a denser outer layer; (2) the surrounding moisture may induce gelation and cause the outer surface to be viscous, thus minimizing non-solvent intrusion and forming a denser skin; (3) a longer air gap distance may create more stretching by both gravity and take-up speed, which results in a thinner selective layer with enhanced polymer orientation; (4) a defective selective skin may be formed if polymer chains are over-stretched and torn apart. The former three factors lead to a higher separation factor but a lower ux, while the last factor causes a decline in separation factor. As shown in Table 7, ux increases slightly with decreasing air gap distance for membranes PBI-D-A, B and C; while separation factor shows an up-and-down trend, with

the highest separation factor occurring at air gap of 2 cm (PBI-D-B) where its dense-selective layer becomes thinner and oriented. A further increase in air gap to 5 cm, the dense-selective layer may be over-stretched, causing minor defects and a lower separation performance. The effects of take-up speed have been discussed in the previous section of single-layer hollow bers. However, dual-layer hollow ber membranes are somewhat different from single-layer hollow ber membranes. The former shows increased ux but decreased separation factor with an increase in take-up speed, as observed from membranes PBI-D-B, D, E and F in Table 7. It is due to the fact that these two layers experience different shear and elongational stresses during spinning since they have quite different compositions and viscosities. Because the outer layer has a higher viscosity, it will bear more loads (i.e., elongational stresses). Thus it can be stretched, oriented, and become thinner simultaneously. Table 7 also lists the membrane diameter which decreases with increasing take-up speed. Since a thin selective skin generates in a high ux, while a high take-up speed produces more defects, a higher takeup speed for PBI/PEI dual-layer hollow ber membranes results in a thinner selective layer, a higher ux and a lower separation factor. In addition, it seems that take-up speed may affect membrane performance more drastically than air gap distance. Table 8 tabulates the tensile properties of dual-layer hollow ber membranes, while Fig. 9 compares the tensile behavior of dual-layer PBI/PEI and single-layer PBI hollow ber membranes

Table 7 Pervaporation performance of PBI/PEI dual-layer hollow ber membranes. Membrane ID PBI-D-A PBI-D-B PBI-D-C PBI-D-D PBI-D-E PBI-D-F Air gap (cm) 5 2 1 2 2 2 Take-up speed (m/min) 4.60 4.60 4.60 9.59 16.24 21.79 OD (m) 1222 1229 1226 899 686 597 ID (m) 721 752 725 589 425 376 Outer-layer thickness (m) 14.2 17.5 14.6 9.5 5.2 4.8 Permeate (H2 O, wt.%) 99.93 99.96 99.90 99.76 99.72 99.67 Total ux (g/m2 h) 232 241 266 492 596 732 Separation factor 2156 2288 1016 436 373 303

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159 Table 8 Mechanical strength of PBI/PEI dual-layer hollow bers. Membrane ID PBI-D-A PBI-D-B PBI-D-C PBI-D-D PBI-D-E PBI-D-F Max tensile stress (MPa) 11.1 12.7 11.0 9.4 12.7 13.2 Youngs modulus (MPa) 417 423 436 325 527 600 Max strain (mm/mm) 0.23 0.21 0.20 0.24 0.25 0.28

157

Fig. 10. A comparison of pervaporation performance for the dehydration of ethylene glycol with previous studies.

Fig. 9. The typical tensile behavior of single-layer and dual-layer hollow ber membranes.

PBI-D-C has a separation factor of 4500 with a comparable high ux (186 g/m2 h) for the dehydration of a feed solution containing 64 wt.% ethylene glycol. Similar phenomena have been observed by other researches [8,23,30,5254] for various alcohol dehydrations. Another important reason is that the hot water treatment washes away the residual DMAc solvent trapped inside the ber, where it acts as a plasticizer. The ber will shrink due to solvent loss, thus the separation performance will be impacted. 3.5. Comparison with previous studies on pervaporation dehydration of ethylene glycol Pervaporation dehydration of ethylene glycol had been studied by various researchers in recent years. Usually, there is a tradeoff between the permeability and selectivity in pervaporation although the up-bound line is not well dened yet. A high permeability is often accompanied with a low selectivity, and vice versa. Relatively, membrane materials with high selectivity are generally preferred, as the disadvantage of low ux can be compensated theoretically by introducing asymmetry and suitable fabrication optimization. A literature search reveals that hydrophilic chitosanand poly(vinyl alcohol)-based membranes are most widely used for ethylene glycol dehydration. Much research has been focused on new potential membrane materials as well as modication methods to improve membrane performance. Fig. 10 shows a comparison between the PBI/PEI dual-layer hollow ber membranes and other polymeric membranes in terms of separation factor versus permeation ux [1,5,5565] under similar operating conditions. Feed compositions of 50/50 wt.% and 64/36 wt.% ethylene glycol/water are chosen for the PBI/PEI duallayer hollow ber membranes. The PBI/PEI dual-layer hollow ber membranes, especially those thermal treated ones present much higher separation factors than most other polymeric membranes.

using PBI-S-G and PBI-D-E membranes as examples. PBI-D-E duallayer hollow bers have much longer breaking strains than PBI-S-E single-layer hollow bers, although the single-layer hollow bers exhibit higher tensile stress. This is a distinct proof of the improved elongation and exibility by means of the dual-layer co-extrusion technology. 3.4. Effect of heat post-treatment on membrane performance Generally, asymmetric membranes are often exposed to postthermal treatments prior to industrial applications. Thermal treatments typically lead to enhanced separation performance because of the elimination of micro-defects in the thin separating layer. A treatment at high temperatures can promote thermal motion of polymer chains and their interactions, facilitating chain relaxation and rearrangement towards a denser and closer packing. As a consequence, the thermally treated membrane will have morphology with a smaller free volume and a higher transportation resistance. The effects of heat treatment might not be vivid if the as-spun membrane already has a high separation factor. Using membranes PBI-D-C and PBI-D-F as examples, Table 9 shows their pervaporation performance before and after thermal treatments at 75 C for 2 h. The ux declines, while the separation factor increases signicantly after the thermal treatment. The annealed membrane

Table 9 Effect of the thermal treatment of PBI/PEI dual-layer hollow ber membranes on the pervaporation performance. Membrane ID PBI-D-C PBI-D-C annealed PBI-D-F PBI-D-F annealed Air gap (cm) 1 2 Take-up speed (m/min) 4.60 21.79 Permeate (H2 O, wt.%) 99.74 99.96 99.69 99.81 Total ux (g/m2 h) 222 186 758 597 Separation factor 1047 4524 592 1004

Feed composition: EG/H2 O (64/36 wt.%)

158

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159

Compared to the commercially available GFT membrane [58], the newly developed PBI/PEI membranes have high separation factors but slightly low uxes. Further efforts should be devoted to enhance the permeation ux by surface modication or spinning optimization. 4. Conclusion In this study, PBI at-sheet dense membranes, single-layer and dual-layer hollow ber membranes are fabricated and studied for pervaporation dehydration of ethylene glycol. The following conclusions can be made from this study: (1) The PBI at-sheet dense membranes show very poor separation performance for ethylene glycol dehydration due to severe membrane swelling. Experimental results indicate that a higher separation performance is obtained if the pervaporation is conducted when the membrane skin cast on the glass plate is facing the feed. (2) Compared to the at-sheet dense membranes, the hollow ber membranes show much better separation performance in both permeation ux and separation factor because of thin selective layer and porous substrate layer with less transport resistance. However, the single-layer hollow bers are of poor toughness and exibility for module fabrication. (3) Dual-layer PBI/PEI hollow ber membranes are successfully developed with improved mechanical toughness and enhanced selectivity compared to the PBI single-layer hollow ber membranes. The superior performance is attributed to both the balanced properties of the PBI selective layer and the less swelling characteristic of the PEI supporting layer. The desirable membrane morphology with seamless interface is another important factor resulting in surprising synergism of the above two factors. (4) The effects of take-up speed and air gap distance on pervaporation performance of PBI/PEI dual-layer hollow ber membranes have been studied. An increase in air gap results in hollow bers with a higher separation factor but a lower ux; while an increase in take-up speed results in hollow bers with a lower separation factor but a higher ux. An optimal air gap of 2 cm and a low take-up speed of 4.60 m/min result in membranes with the highest separation performance. (5) A mild thermal treatment of PBI/PEI dual-layer hollow ber membranes at 75 C can enhance separation performance efciently. By comparison with previous works on pervaporation dehydration of ethylene glycol, PBI/PEI dual-layer hollow ber membranes developed in this study, especially the thermal treated membranes, show higher separation factor than most other polymeric membranes. Further efforts for ux enhancement are needed in the future. Acknowledgement The authors thank PBI Performances Products, Inc. (R-279-000279-597) for funding this research. Special thanks are given to Dr. Jing Cai Su and Dr. Kai Yu Wang for their help on hollow ber spinning. Thanks also due to Mr. Weijie Bernard Neo and Mr. Kun Cheng for their help on experiments.

OD PBI PEI PAA PVA PAN LiCl PVP EG FESEM xw,i yw,i ET (30) d p h sp

outer diameter of hollow ber polybenzimidazole polyetherimide poly(amic acid) poly(vinyl alcohol) poly(acryl nitrile) lithium chloride polyvinylpyrrolidone ethylene glycol eld emission scanning electron microscope separation factor weight fraction of component i in the feed (wt.%) weight fraction of component i in the permeate (wt.%) empirical solvent polarity parameter (kcal/mol) dispersion solubility parameter (MPa)1/2 polar force solubility parameter (MPa)1/2 hydrogen-bonding solubility parameter (MPa)1/2 total solubility parameter (MPa)1/2

References
[1] J.R. Du, A. Chakma, X. Feng, Dehydration of ethylene glycol by pervaporation using poly(N,N-dimethylaminoethyl methacrylate)/polysulfone composite membranes, Sep. Purif. Technol. 64 (2008) 63. [2] C. Reichardt, Solvents and Solvent Effects in Organic Chemistry, 2nd ed., VCH, Weinheim, 1988. [3] C.M. Hansen, Hansen Solubility Parameters: A Users Handbook, CRC Press, Boca Raton, Fla., 2007. [4] R.Y.M. Huang, N.R. Jarvis, Separation of liquid mixtures by using polymer membranes. II. Permeation of aqueous alcohol solutions through cellophane and poly(vinyl alcohol), J. Appl. Polym. Sci. 14 (1970) 2341. [5] X.S. Feng, R.Y.M. Huang, Pervaporation with chitosan membranes. 1. Separation of water from ethylene glycol by a chitosan/polysulfone composite membrane, J. Membr. Sci. 116 (1996) 67. [6] M.N. Hyder, R.Y.M. Huang, P. Chen, Composite poly(vinyl alcohol)poly(sulfone) membranes crosslinked by trimesoyl chloride: characterization and dehydration of ethylene glycolwater mixtures, J. Membr. Sci. 326 (2009) 363. [7] P. Shao, R.Y.M. Huang, Polymeric membrane pervaporation, J. Membr. Sci. 287 (2007) 162. [8] Y. Wang, L.Y. Jiang, T. Matsuura, T.S. Chung, S.H. Goh, Investigation of the fundamental differences between polyamide-imide (PAI) and polyetherimide (PEI) membranes for isopropanol dehydration via pervaporation, J. Membr. Sci. 318 (2008) 217. [9] X. Qiao, T.S. Chung, Fundamental characteristics of sorption, swelling, and permeation of P84 co-polyimide membrane for pervaporation dehydration of alcohols, Ind. Eng. Chem. Res. 44 (2005) 8938. [10] T.A. Peters, C.H.S. Poeth, N.E. Benes, H.C.W.M. Buijs, F.F. Vercauteren, J.T.F. Keurentjes, Ceramic-supported thin PVA pervaporation membranes combining high ux and high selectivity; contradicting the ux-selectivity paradigm, J. Membr. Sci. 276 (2006) 42. [11] T.A. Peters, N.E. Benes, J.T.F. Keurentjes, Hybrid ceramic-supported thin PVA pervaporation membranes: long-term performance and thermal stability in the dehydration of alcohols, J. Membr. Sci. 311 (2008) 7. [12] R. Kreiter, D.P. Wolfs, C.W.R. Engelen, H.M. van Veen, J.F. Vente, Hightemperature pervaporation performance of ceramic-supported polyimide membranes in the dehydration of alcohols, J. Membr. Sci. 319 (2008) 126. [13] Y.X. Zhu, S.S. Xia, G.P. Liu, W.Q. Jin, Preparation of ceramic-supported poly(vinyl alcohol)-chitosan composite membranes and their applications in pervaporation dehydration of organic/water mixtures, J. Membr. Sci. 349 (2010) 341. [14] J.D. Jou, W. Yoshida, Y. Cohen, A novel ceramic-supported polymer membrane for pervaporation of dilute volatile organic compounds, J. Membr. Sci. 162 (1999) 269. [15] W. Yoshida, Y. Cohen, Ceramic-supported polymer membranes for pervaporation of binary organic/organic mixtures, J. Membr. Sci. 213 (2003) 145. [16] J.H. Kim, K.H. Lee, S.Y. Kim, Pervaporation separation of water from ethanol through polyimide composite membranes, J. Membr. Sci. 169 (2000) 81. [17] H.M. Guan, T.S. Chung, Z. Huang, M.L. Chng, S. Kulprathipanja, Poly(vinyl alcohol) multilayer mixed matrix membranes for the dehydration of ethanolwater mixture, J. Membr. Sci. 268 (2006) 113. [18] H. Yanagishita, D. Kitamoto, K. Haraya, T. Nakane, T. Okada, H. Matsuda, Y. Idemoto, N. Koura, Separatioin performance of polyimide composite membrane prepared by dip coating process, J. Membr. Sci. 188 (2001) 165.

Nomenclature DMAc ID N,N-dimethylacetimide inner diameter of hollow ber

Y. Wang et al. / Journal of Membrane Science 363 (2010) 149159 [19] D.M. Sullivan, M.L. Bruening, Ultrathin cross-linked polyimide pervaporation membranes prepared from polyelectrolyte multilayers, J. Membr. Sci. 248 (2005) 161. [20] Y. Wang, S.H. Goh, T.S. Chung, N. Peng, Polyamide-imide/polyetherimide dual-layer hollow ber membranes for pervaporation dehydration of C1C4 alcohols, J. Membr. Sci. 326 (2009) 222. [21] K.Y. Wang, T.S. Chung, R. Rajagopalan, Dehydration of tetrauoropropanol (TFP) by pervaporation via novel PBI/BTDA-TDI/MDI co-polyimide (P84) dual-layer hollow ber membranes, J. Membr. Sci. 287 (2007) 60. [22] R.X. Liu, X.Y. Qiao, T.S. Chung, Dual-layer P84/polyethersulfone hollow ber for pervaporation dehydration of isopropanol, J. Membr. Sci. 294 (2007) 103. [23] L.Y. Jiang, H. Chen, Y.C. Jean, T.S. Chung, Ultra-thin polymeric interpenetration network with separation performance approaching ceramic membranes for biofuel, AIChE J. 55 (2009) 75. [24] Y. Wang, S.H. Goh, T.S. Chung, Miscibility study of Torlon polyamide-imide with Matrimid 5218 polyimide and polybenzimidazole, Polymer 48 (2007) 2901. [25] E.W. Neuse, Aromatic polybenzimidazoles: syntheses, properties, and applications, Adv. Polym. Sci. 47 (1982) 1. [26] A. Buckley, D. Stuetz, G.A. Serad, in: J.I. Kroschwitz (Ed.), Polybenzimidazoles, Encyclopedia of Polymer Science and Engineering, Wiley, New York, 1987, 572. [27] T.S. Chung, A critical review of polybenzimidazoles: historical development and future R & D, J. Macromol. Sci. Rev. Macromol. Chem. Phys. C 37 (1997) 277. [28] M. Jaffe, P. Chen, E.W. Choe, T.S. Chung, S. Makhija, High performance polymer blends, Adv. Polym. Sci. 117 (1994) 297. [29] N.W. Brooks, R.A. Duckett, J. Rose, I.M. Ward, J. Clements, An NMR study of absorbed water in polybenzimidazole, Polymer 34 (1993) 4038. [30] T.S. Chung, W.F. Guo, Y. Liu, Enhanced matrimid membranes for pervaporation by homogenous blends with polybenzimidazole (PBI), J. Membr. Sci. 271 (2006) 221. [31] PBI Performance Products, Inc., Solutions brochure of polybenzimidazole (PBI) S26 solution, http://www.pbiproducts.com/les/SolutionsMSDS.pdf. [32] T.S. Chung, S.K. Teoh, X. Hu, Formation of ultra-thin high-performance polyethersulfone hollow ber membranes, J. Membr. Sci. 133 (1997) 161. [33] J.Z. Ren, T.S. Chung, D.F. Li, R. Wang, Y. Liu, Development of asymmetric 6FDA26 DAT hollow ber membranes for CO2 /CH4 separation. 1. The inuence of dope composition and rheology on membrane morphology and separation performance, J. Membr. Sci. 207 (2002) 227. [34] Y. Li, C. Cao, T.S. Chung, K.P. Pramoda, Fabrication of dual-layer polyethersulfone (PES) hollow ber membranes with an ultrathin dense selective layer for gas separation, J. Membr. Sci. 245 (2004) 53. [35] T. Matsuura, Synthetic Membranes and Membrane Separation Processes, CRC Press, Boca Raton, 1994. [36] R.X. Liu, X. Qiao, T.S. Chung, The development of high performance P84 copolyimide hollow bers for pervaporation dehydration of isopropanol, Chem. Eng. Sci. 60 (2005) 6674. [37] G.W. Li, W. Zhang, J.P. Yang, X.P. Wang, Time-dependence of pervaporation performance for the separation of ethanol/water mixtures through poly(vinyl alcohol) membrane, J. Colloid Interface Sci. 306 (2007) 337. [38] H. Zhang, H.G. Ni, X.P. Wang, X.B. Wang, W. Zhang, Effect of chemical groups of polystyrene membrane surface on its pervaporation performance, J. Membr. Sci. 281 (2006) 626. [39] L.Y. Jiang, T.S. Chung, -Cyclodextrin containing Matrimid nanocomposite membranes for pervaporation application, J. Membr. Sci. 327 (2009) 216. [40] 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) 1. [41] T.S. Chung, E.R. Kafchinski, The effects of spinning conditions on asymmetric 6FDA/6FDAM polyimide hollow bers for air-separation, J. Appl. Polym. Sci. 65 (1997) 1555. [42] N. Widjojo, T.S. Chung, The thickness and air-gap dependence of macrovoid evolution in phase-inversion asymmetric hollow ber membranes, Ind. Eng. Chem. Res. 45 (2006) 7618. [43] N. Peng, T.S. Chung, K.Y. Wang, Macrovoid evolution and critical factors to form macrovoid-free hollow ber membranes, J. Membr. Sci. 318 (2008) 363. [44] C. Cao, T.S. Chung, S.B. Chen, Z.J. Dong, The study of elongation and shear rates in spinning process and its effect on gas separation performance of poly(ether sulfone) (PES) hollow ber membranes, Chem. Eng. Sci. 59 (2004) 1053.

159

[45] J.J. Qin, R. Wang, T.S. Chung, Investigation of shear stress effect within a spinneret on ux, separation and thermomechanical properties of hollow ber ultraltration membranes, J. Membr. Sci. 175 (2000) 197. [46] N. Peng, F T., S. Chung, The effects of spinneret dimension and hollow ber dimension on gas separation performance of ultra-thin defect-free Torlon hollow ber membranes, J. Membr. Sci. 310 (2008) 455. [47] T.S. Chung, The limitations of using FloryHuggins equation for the states of solutions during asymmetric hollow ber formation, J. Membr. Sci. 126 (1997) 19. [48] X.S. Feng, R.Y.M. Huang, Preparation and performance of asymmetric polyetherimide membranes for isopropanol dehydration by pervaporation, J. Membr. Sci. 109 (1996) 165. [49] T.S. Chung, Z.L. Xu, Asymmetric hollow ber membranes prepared from miscible polybenzimidazole and polyetherimide blends, J. Membr. Sci. 147 (1998) 35. [50] D.F. Li, T.S. Chung, R. Wang, Morphological aspects and structure control of dual-layer asymmetric hollow ber membranes formed by a simultaneous coextrusion approach, J. Membr. Sci. 243 (2004) 155. [51] T.S. Chung, X. Hu, The effect of air-gap distance on the morphology and thermal properties of polyethersulfone hollow bers, J. Appl. Polym. Sci. 66 (1997) 1067. [52] X. Qiao, T.S. Chung, K.P. Pramoda, Fabrication and characterization of BTDATDI/MDI (P84) co-polyimide membranes for the pervaporation dehydration of isopropanol, J. Membr. Sci. 264 (2005) 176. [53] Y.H. See-Toh, F.C. Ferreira, A.G. Livingston, The inuence of membrane formation parameters on the functional performance of organic solvent nanoltration membranes, J. Membr. Sci. 299 (2007) 236. [54] I. Pinnau, B.D. Freeman, Membrane formation and modication, American Chemical Society, Washington, DC, 1999. [55] R.L. Guo, X. Fang, H. Wu, Z.Y. Jiang, Preparation and pervaporation performance of surface crosslinked PVA/PES composite membrane, J. Membr. Sci. 322 (2008) 32. [56] C.L. Hu, B. Li, R.L. Guo, H. Wu, Z.Y. Jiang, Pervaporation performance of chitosanpoly(acrylic acid) polyelectrolyte complex membranes for dehydration of ethylene glycol aqueous solution, Sep. Purif. Technol. 55 (2007) 327. [57] C.L. Hu, R.L. Guo, B. Li, X.C. Ma, H. Wu, Z.Y. Jiang, Development of novel mordenite-lled chitosanpoly(acrylic acid) polyelectrolyte complex membranes for pervaporation dehydration of ethylene glycol aqueous solution, J. Membr. Sci. 293 (2007) 142. [58] M.C. Burshe, S.B. Sawant, J.B. Joshi, V.G. Pangarkar, Dehydration of ethylene glycol by pervaporation using hydrophilic IPNs of PVA, PAA and PAAM membranes, Sep. Purif. Technol. 13 (1998) 47. [59] R.Y.M. Huang, P. Shao, X. Feng, W.A. Anderson, Separation of ethylene glycolwater mixtures using sulfonated poly(ether ether ketone) pervaporation membranes: membrane relaxation and separation performance analysis, Ind. Eng. Chem. Res. 41 (2002) 2957. [60] W. Jehle, T. Staneff, B. Wagner, J. Steinwandel, Separation of glycol and water from coolant liquids by evaporation, reverse osmosis and pervaporation, J. Membr. Sci. 102 (1995) 9. [61] Y.S. Nam, Y.M. Lee, Pervaporation of ethylene glycol-water mixtures. I. Pervaporation performance of surface crosslinked chitosan membranes, J. Membr. Sci. 153 (1999) 155. [62] R.L. Guo, C.L. Hu, F.S. Pan, H. Wu, Z.Y. Jiang, PVA-GPTMS/TEOS hybrid pervaporation membrane for dehydration of ethylene glycol aqueous solution, J. Membr. Sci. 281 (2006) 454. [63] R.L. Guo, X.C. Ma, C.L. Hu, Z.Y. Jiang, Novel PVAsilica nanocomposite membrane for pervaporative dehydration of ethylene glycol aqueous solution, Polymer 48 (2007) 2939. [64] R.L. Guo, C.L. Hu, B. Li, Z.Y. Jiang, Pervaporation separation of ethylene glycol/water mixtures through surface crosslinked PVA membranes: Coupling effect and separation performance analysis, J. Membr. Sci. 289 (2007) 191. [65] F.R. Chen, H.F. Chen, Pervaporation separation of ethylene glycol-water mixtures using crosslinked PVA-PES composite membranes. 1. Effects of membrane preparation conditions on pervaporation performances, J. Membr. Sci. 109 (1996) 247.

You might also like