You are on page 1of 11

Separation and Purification Technology 57 (2007) 209219

Effect of molecular weight and concentration of PEG additives


on morphology and permeation performance of
cellulose acetate hollow fibers
Wen-Li Chou a , Da-Guang Yu a , Ming-Chien Yang b, , Chi-Hsiung Jou c
a

Department of Materials and Fibers, Nanya Institute of Technology, Chung-Li, Tao-Yuan 320, Taiwan, ROC
Department of Polymer Engineering, National Taiwan University of Science and Technology, Taipei 106, Taiwan, ROC
c Department of Materials and Textiles, Oriental Institute of Technology, Pan-Chiao, Taipei County 220, Taiwan, ROC
Received 30 September 2006; received in revised form 29 March 2007; accepted 12 April 2007

Abstract
Asymmetric cellulose acetate (CA) hollow fiber membranes were prepared by dry/wet spinning process from a dope composed of cellulose
acetate (CA), N,N-dimethylformamide (DMF), and polyethylene glycol (PEG). Herein, PEG was the additive; DMF was the solvent; whereas
water was the nonsolvent. The spinning parameters in this study were the contents and molecular weights of PEG and the external coagulation
temperature. The surface and cross-section morphology of the resulting hollow fibers were examined using scanning electron microscopy (SEM).
The pure water permeability (PWP) and retention of dextran were also measured. The results showed that the addition of PEG would suppress the
formation of macrovoids. The effect on the suppression was more obvious for PEG with higher molecular weight and higher content. When PEG
was added, the outer surfaces changed from smooth to microporous, whereas the inner surface remained smooth and dense. The PWP increased
with the additive content but slightly decreased with the increase of molecular weight. Oppositely, the retention of dextrans decreased with the
increase of additive contents but increased with the molecular weight. When adding PEG and coagulating at higher temperature simultaneously,
the outer surface and cross-section of CA/PEG blended membrane exhibited macrovoids on outer surface and finger-like voids near the inner and
outer edges. The resulting membranes showed higher PWP (47 times) with slight decrease in the retention. Hence, adding PEG and elevating
coagulant temperature will promote the permeation performance of CA hollow fibers.
2007 Elsevier B.V. All rights reserved.
Keywords: Ultrafiltration; Hollow fiber; Cellulose acetate; PEG additive; Drywet spinning

1. Introduction
Since the successful development of cellulose acetate (CA)
asymmetric membranes with a very dense and thin active layer
on the top of a porous substrate [1], CA membranes have been
applied in reverse osmosis for converting sea water into freshwater. Furthermore, CA hollow fiber membranes are widely used
for clinical hemodialysis. For instance, Kell and Mahoney [2]
invented a CA hollow fiber suitable for use in artificial kidneys
to provide superior water and solute clearances.
Hemodialysis is one of the most important methods for blood
purification. In general the characteristics of a hemodialyzing
material are ultrafiltration rate, solute permeability, mechani-

Corresponding author. Tel.: +886 2 2737 6528; fax: +886 2 2737 6544.
E-mail address: myang@mail.ntust.edu.tw (M.-C. Yang).

1383-5866/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.seppur.2007.04.005

cal strength, and hemocompatibility. Because of CA hollow


fiber membranes have an excellent performance in the removal of low molecular weight toxic substances such as urea and
creatinine from patients and can reduce the activation of the
complement system and the leucopenia observed in the early
period of blood extracorporeal circulation. However, CA hollow fiber membranes have demonstrated insufficient removal
of lower molecular weigh proteins such as 2 -microglobulin
(2 -MG, 11,800 Da) which can cause amyloidosis. Besides, CA
membranes can easily cause serious protein adsorption and clot
formation when the membrane contacts blood as do other blood
purification membranes. These phenomena induce a decrease
in the water flux and solute permeability of the membrane and
require infusion of an anticoagulant into the patient during blood
purification therapy and these factors make serious complications for chronic renal failure patients who require continuous
and long time treatment [35]. Therefore, many researchers or

210

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

companies have pay attention to improve its performance by


added additive or plasticizer [2,68].
Cellulose acetate hollow fibers are produced by the drywet
spinning technique, and often have asymmetric structure with
dense inner and outer skin layers, therefore its performance
for water treatment or hemodialysis is always undesirable. Furthermore, the dry/wet spinning is a complex process since it
involves many process variables, such as the dope extrusion rate
[911], the flow rate of core liquid [9,12,13], the composition
and temperature of the core liquid and coagulant [12,1418],
post-drawing and on-line drawing [19,20], the length of the air
gap [9,10,13,2126], and additive agent or nonsolvent to dope
[8,27], that would influence both the geometrical characteristics
and the permeation properties of the hollow fibers.
The addition of organic or inorganic components as a third
component to a casting solution has been one of the important
techniques used in membrane preparation. However, the role of
organic and inorganic additives, such as poly(vinylpyrrolidone)
(PVP), polyethylene glycol (PEG), water, LiCl, and ZnCl2 ,
has been reported as a pore-forming agent to enhance the permeation properties. This behavior was explained in terms of
their water-soluble characteristics [3,28]. Although their works
are important and interesting, their theories might need further
modifications to be applied to hollow fiber membrane fabrication. It is due to the fact that it is difficult to simulate the
hollow-fiber spinning process by adopting the process conditions developed for asymmetric flat membranes. In addition,
Merrill et al. [29] reported that added PEG additive in the polymer solution or grafted PEG on the polymer surface would
improve anticoagulation.
In order to suppress macrovoids in the hollow fiber membranes, Liu et al. added high concentration of PEG400 to the
dope of polyethersulfone (PES) and using PEG400 aqueous
solution as the coagulant to control the morphology of the resulting hollow fiber [30]. Li et al. employed a co-extrusion spinneret
to prepare dual-layer Matrimid/PES hollow fibers [31]. They
also added high concentration (66%) of PEG (MW 100 kDa) to
the spinning dope in order to suppress the macrovoids of the
PES layer.
In this paper, in order to increase the removal of lower molecular weight proteins such as 2 -microglobulin, CA asymmetric
hollow fiber membranes were prepared from polymer solutions
composing of CA/PEG/DMF via drywet spinning method. By
varying the coagulant temperature and the molecular weight and
content of PEG, the morphological and transport properties were
systemically investigated. The method of this research not only
added additive to the dope, but also varied the coagulant temperature. In so doing, the permeation performance can be improved
with a single-pass manufacture procedure.
2. Experimental
2.1. Materials
Cellulose acetate (number average molecular weight 50,000
with acetyl content 39.8%) was purchased from Aldrich, USA.
N,N-dimethylformamide (DMF), sulfuric acid, and phenol were

purchased from Acros, USA. PEG (number average molecular


weight 1, 10 and 40 kDa) were purchased from Merck, USA.
Dextran with molecular weights of 10.2, 35.0, and 66.7 kDa were
purchased from Sigma, USA and stored at a suitable temperature
before use.
2.2. Spinning conditions
A homogenous spinning solution was prepared by adding
25 wt% CA polymer powder in 75 wt% solvent of DMF. In the
dope, the PEG was added 7, 14 and 21% with respect to the
weight of CA, respectively. All the ingredients were added in
the flask and were stirred at room temperature for 24 h until the
CA polymer and PEG additive were entirely dissolved to form a
homogenous spinning dope. Afterward, the dope was degassed
in a vacuum oven at room temperature for 2 h and extruded under
100 kPa through a hollow fiber spinneret with an inner diameter
of 0.4 mm and an outer diameter of 0.6 mm. Spinning of the
CA hollow fiber was based on the drywet technique. Both the
core liquid and the external coagulant were pure water, and the
external coagulant temperature was controlled at either 25, 50,
or 75 C. The air gap which is the distance between the tip of
spinneret and the surface of external coagulant was kept 20 cm.
After coagulation, the resulting CA hollow fiber was wound
up with a take-up roller and rinsed with water for 2 days to
remove residual DMF. Then the fibers were immersed in 50 wt%
glycerol solution for another 24 h and then were dried in air
at room temperature for ultrafiltration testing and morphology
examination. The spinning conditions and detailed parameters
were listed in Table 1.
2.3. Morphology studies
In order to examine the morphology of hollow fiber membranes, dried samples of hollow fibers were broken in liquid
Table 1
Process parameters and spinning conditions
Process parameters/spinning
conditions

Value

1. Polymer composition
2. PEG molecular weight (Dalton)
3. PEG concentration (PEG vs. CA by
weight)
4. CA concentration (by weight)
5. Spinneret temperature ( C)
6. Dope extrusion rate (g/min)
7. Spinneret OD/ID
8. Bore liquid
9. Bore liquid flow rate (ml/min)
10. Dope pressure (kPa)
11. Air-gap distance (cm)
12. Take-up velocity (m/min)
13. Coagulant composition
14. Coagulant temperature ( C)
15. Drying procedure

CA/DMF/PEG
1 k, 10 k, 40 kDa
7%, 14%, 21%
25%
Room temperature
11
0.6 mm/0.4 mm
H2 O
7.6
100
20
35
Tap water
2570
A few days in deionized water. One
day in 50% (w/w) glycerin aqueous
solution at room temperature and
dried at room temperature

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

nitrogen and then sputtered with a thin layer of gold using Jeol
JFC-11-E sputtering device. The outer and inner surfaces of the
hollow fibers were examined using a field emission scanning
electrical microscopy (FESEM) (S-800, Hitachi, Tokyo, Japan).
2.4. Ultrafiltration experiments
The experimental setup for measuring the pure water permeability (PWP) and retention of dextran for CA hollow fiber
was shown in our previous work [20]. Hollow fibers were tested in bundles of 10 fibers of 15 cm in length. Each end of the
module was a piece of Teflon tubing, 6.35 mm (1/4 in.) in diameter and 2.5 cm in length. The fibers were potted at both ends
with epoxy resin and mounted in a test module. While not taking
measurements, the modules were preserved in 50 wt% glycerol.
The preserving solution was flushed out first with water before
taking data. The feed was held in the hopper under a pressure of
50 kPa. To prevent concentration polarization [32], the feed was
recirculated at a flow rate of 25 ml/min with a metering pump
(RH1CKC, Fluid Metering, Inc., USA) through the lumen of the
hollow fibers, and the permeate was collected from the shell side
to simulate hemodialysis. The PWP was determined when the
flow rate stabilized. The pore size distribution was determined
using dextran of different molecular weights (10.2k, 35k, 66.7k)
to simulate the 2 -MG (MW: 11,800), pepsin (MW: 35,000) and
bovine serum albumin (MW: 68,000) protein. The dextran solutions concentration in the feed solution was kept at 1000 ppm.
The concentrations of dextran solutions were determined by
sulfuric acid phenol solution titration [33,34] and the absorbance at 490 nm was determined using an ultraviolet-visible
spectrophotometer (CE7200, Aquarius, England). The PWP, or
the ultrafiltration rate (UFR), was calculated by the following
equation:
PWP =

Q
Q
=
AP
nDi LP

(1)

where Q is the volumetric flow rate (L/h), A is the inner surface


area of the hollow fiber membrane (m2 ), n is the numbers of
fibers, Di is the inner diameter of hollow fiber, L is the length of
the hollow fiber and P is the applied pressure of ultrafiltration

211

experiment. The retention R was calculated as follows:


R=1

Cp
Cf

(2)

where Cf and Cp are the solute concentrations in the feed


and permeate, respectively. All experiments were performed in
hollow-fiber modules. Three modules were prepared for each
hollow-fiber sample.
3. Result and discussion
3.1. The effect of molecular weight and content of PEG on
morphology
Fig. 1 shows the SEM photos of CA and CA/PEG hollow
fiber membranes via drywet spinning under the same spinning conditions. The molecular weight and content of PEG were
40 kDa and 21 wt%-CA, respectively. The CA membrane in
Fig. 1(a) exhibits an asymmetrical structure consisting of a dense
top-layer and a porous sub-layer with droplet morphologies near
the inner edge, whereas the CA/PEG membrane in Fig. 1(b) also
has an asymmetrical structure but without observable droplet
morphologies.
In order to understand the reasons that the macrovoids of
membranes would be reduced by the addition of PEG, the molecular weight and content of PEG in the dope were varied. Fig. 2
shows the SEM photos of partial cross-section. Those hollow
fiber membranes were prepared with three different contents (7,
14, 21 wt%-CA) of the same molecular weight (10 kDa). These
images indicated that the droplet morphologies enclosed in the
inner edge gradually disappeared with increasing PEG content
and that the thickness of the membranes increased with the PEG
content. The reason may be that the solid content of the dope
was increased due to the addition of PEG. Because PEG can
increase the viscosity of the solvent, this hindered the exchange
of solvent and nonsolvent, thus higher PEG content led to denser
structure. The other possible reason is that some part of the PEG
chains on the membranes surface penetrate into the pores and
are attached inside of the pores [35].

Fig. 1. SEM photograph of hollow fiber membrane: (a) CA/DMF/water system (magnification: 140) and (b) CA/PEG/DMF/water system (magnification: 150).

212

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

Fig. 2. Effect of the PEG (MW 10 kDa) contents (PEG vs. CA weight ratio) on the morphology of cross-section of CA hollow fiber membranes: (a) 0%; (b) 7%; (c)
14% and (d) 21%.

In general, macrovoid formation during phase separation


occurs from freshly formed nuclei of the diluted phase when the
composition in front of the nuclei remains stable for a relatively
long period of time [36,37]. Due to diffusion of solvent expelled
from the surrounding polymer solution the macrovoid grows.
Macrovoids are generally formed in systems where instantaneous demixing takes place, except when the polymer additive
(e.g. PVP) concentration and/or the nonsolvent concentration in
the polymer solution exceed a certain minimum value [3639].
In addition, some researchers [29,40,41] have reported that the
macrovoid formation disappears as adding hydrophilic additive
to the casting solution [41] and that the macrovoids are suppressed by adding organic acids, because the acids form acidbase
complex with basic polar solvents such as NMP, DMF, and
DMAc [40]. Polymeric additives such as PVP and PEG are
widely used for the structure control for the fabrication of ultrafiltration and microfiltration membranes [29]. Based on those
SEM photos in the Fig. 2, our results are in accordance with
their reports.
The effect of molecular weight of PEG (1 k, 10 k and 40 kDa)
on the morphologies was also investigated, as shown in Fig. 3.

Fig. 3 indicates that the droplet macrovoids was obviously suppressed as higher molecular weight of PEG was added. Although
those hollow fiber membranes were spun from the same PEG
content (21%-CA), Fig. 3(a) and (b) (1 kDa PEG was added)
apparently show a few macrovoids near the inner edge, whereas Fig. 3(c) and (d) (10 kDa PEG was added) and Fig. 3(e)
and (f) (40 kDa PEG was added) show that macrovoids were
inhibited. The difference between Fig. 3(c)(f) was only in the
thickness, and the thickness of membrane increased with the
molecular weight (Fig. 3(f)). The effect of the contents and molecular weights of PEG on the thickness of membranes were also
exhibited in the Table 2.
Table 2 lists the dimensions of the resulting CA/PEG hollow fibers. In general, higher PEG content and higher molecular
weight of PEG led to larger wall thickness of the hollow fiber,
since the total polymer content (CA and PEG) was higher. In
addition, larger PEG molecules would be more difficult to be
leached, and leading to thicker hollow fibers.
It has been known that the formation of macrovoid is
influenced by the molecular weight of additives [42,43]. Lower
molecular weight PEG is more soluble in water than higher

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

213

Fig. 3. SEM overall and partial cross-section photographs of membranes made from of CA solutions containing various molecular weight and same content of PEG
(21 wt%): (a and b) PEG (1 kDa); (c and d) PEG (10 kDa) and (e and f) PEG (40 kDa).

molecular weight, thus low molecular weight PEG can be easily


washed out quickly by water during the formation of membrane.
Therefore, the thickness and the droplet-like macrovoids were
dependent on the solubility of PEG. In fact, intensive studies on
the effect of water-soluble additives to a dope solution on the
morphology of asymmetric membranes have been performed
[37,44,45]. Jung et al. [46] suggested that the top-layer thicknesses is not varied at low concentration and low molecular
weight of PVP, but the top layer thickness changes distinctly at
high concentration and high molecular weight of PVP. The main
reason is caused by residual content increase with increasing

concentration and molecular weight of PEG during the membrane formation. In summary, as more PEG is added, the number
of macrovoids gradually disappears, regardless of the molecular
weight of PEG. When the same amount of PEG is added, the
wall thickness increased with the molecular weight of PEG, as
shown in Figs. 2 and 3 and Table 2.
Fig. 4 shows that the CA hollow fiber membranes with or
without PEG additive exhibited a dense and sponge-like structure. Fig. 4(b)(d) clearly show that CA/PEG membranes were
denser than original CA membrane in Fig. 4(a). This is because
CA/PEG membranes were spun from a dope with higher poly-

214

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

Table 2
The dimensions of CA/PEG blend hollow fiber membranes
PEG content (wt%-CA)

PEG (Da)

Do (m)

Di (m)

dry (m)

wet (m)

0
7
7
7
14
14
14
21
21
21

1k
10k
40k
1k
10k
40k
1k
10k
40k

473.5 (5.1)
444.2 (3.3)
416.7 (5.1)
449.0 (4.9)
475.2 (6.2)
487.9 (7.3)
500.0 (4.6)
433.9 (3.3)
512.0 (3.8)
515.1 (4.3)

350.0 (4.3)
315.8 (2.4)
293.3 (4.8)
300.0 (2.3)
336.6 (5.1)
330.3 (4.9)
329.6 (6.3)
283.5 (6.5)
341.2 (4.0)
322.1 (5.3)

61.6 (2.0)
64.2 (2.1)
70.0 (4.1)
74.5 (3.4)
69.3 (2.6)
78.8 (3.7)
85.2 (3.5)
73.2 (2.1)
83.5 (2.2)
95.5 (2.2)

81.3 (2.7)
83.7 (2.5)
86.3 (3.3)
94.0 (2.6)
89.3 (2.3)
94.3 (3.1)
100.3 (2.8)
90.3 (2.3)
96.9 (2.9)
110.7 (3.7)

Note: The numbers in the brackets are the standard deviation (n = 46).

mer concentration (CA and PEG). As shown in Fig. 4, the


porosity of the membrane decreased with the increase of molecular weight of PEG. In general, the radius of gyration (Rg ) of
PEG chains increases with the molecular weight [47]:
Rg = 0.0215M0.583
(nm)
w

(3)

Smaller PEG molecules can quickly diffuse out to the coagulant during the drywet spinning process which would lead to

a looser structure, while larger PEG molecules were more difficult to be removed which would retard the exchanging rate of
solvent and nonsolvent and result in a denser structure. Furthermore, when PEG and CA polymer chains entangled with each
other, higher molecular weight would make the entanglements
tighter. Hence the wall structure of membrane appeared denser
when adding higher molecular weight PEG. On the other hand,
because higher molecular weight PEG would diffuse out slower
and were entrapped in the CA matrix, after leaching out the resi-

Fig. 4. SEM image of partial cross-section of outer edge of membranes made from various molecular weight and same content of PEG (14 wt%): (a) original; (b)
PEG (1 kDa); (c) PEG (10 kDa) and (d) PEG (40 kDa).

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

215

Fig. 5. SEM image of inner surface of membranes made from various contents and same molecular of PEG (10 kDa): (a) original (magnification: 50,000); (b) PEG
(7%, magnification: 1000); (c) PEG (14%, magnification: 1000); (d) PEG (21%, magnification: 1000); (e) PEG (14%, magnification: 10,000) and (f) PEG
(21%, magnification: 10,000).

dual PEG, the space originally occupied by PEG would become


larger pores in the membranes.
3.2. Effect of the external coagulant temperature on the
morphology of CA/PEG hollow fiber membranes
Figs. 57 show the SEM micrographs of the inner skin, outer
skin and cross-section of the CA and CA/PEG hollow fibers spun
at various coagulant temperatures, respectively. Fig. 5 depicts
the inner surface of various contents of PEG with a molecular
weight of 10 kDa. Those SEM images show that the inner surface was generally smooth except the ripple of water in various

contents of PEG by magnification 1k50k times. This is because


the bore liquid used pure water at room temperature and was
injected 7.6 ml/min by a gear pump. When the spinning dope
was extruded from the spinneret, the polymer solution immediately contacted with the bore liquid that led to the immediate
occurrence of phase inversion. Because the bore liquid flow rate
(4.03 m/s) is faster than dope extrusion rate (1.27 m/s), the friction exerted by the bore liquid to the nascent hollow fiber led to
the ripples at the inner surface.
Fig. 6 depicts the coagulant temperature on the morphology
of outer surface of the both pure CA and CA/PEG blended hollow fiber membranes. Apparently, the temperature is a strong

216

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

Fig. 6. SEM image of outer surface of membranes made from various coagulant temperature and same content molecular weight of PEG (14%, 10 kDa): (a) pure
CA, 25 C, magnification: 50,000; (b) pure CA, 75 C, magnification: 50,000; (c) 25 C, magnification: 10,000; (d) 50 C, magnification: 10,000; (e) 75 C,
magnification: 5000.

influence on the morphology of membranes, especially, the


CA/PEG blended hollow fiber membranes clearly exhibited
macrovoids at higher coagulant temperature. This is because the
nascent hollow fiber plunged into a coagulation bath of higher
temperature. Higher coagulant temperature cannot only increase
the exchange rate of the solvent and nonsolvent but also would
increase the humidity in the air gap region (the vapor pressure
of water was 3.17, 12.34, and 38.56 kPa at 25, 50, and 75 C,
respectively). Both would accelerate the phase separation. The
molecular mobility would be higher at higher temperature, and
hence the extensibility of the hollow fibers and the humidity
induced phase-separation process seems to be a much faster

process at higher external coagulant temperature. Therefore, raising the coagulant temperature could result in porous surface
due to faster phase separation at higher temperature. In general,
macrovoid formation occurs under rapid precipitation conditions and the precipitation is faster at higher temperature [36].
Besides, higher temperature would increase the effusion rate
of the solvent from the spinning jet, hence the polymer molecules precipitated faster that shortened the time for molecular
rearrangement.
The other important parameter affecting the outer surface
morphologies is that PEG additive is water-soluble, especially
at higher temperature. Higher coagulant temperature could

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

217

Fig. 7. SEM image of partial cross-section of membranes made from various coagulant temperature with same content and molecular weight of PEG (14%, 10 kDa):
(a) pure CA, 25 C, magnification: 1000; (b and c) 25 C, magnification: 700; (d) 50 C, magnification: 700; (e) 75 C, magnification: 700.

increase the exchange rate of the solvent and nonsolvent, hence


the PEG was more quickly effused out when the nascent hollow fiber membrane was immersed in the coagulation bath of
higher temperature. The outer skin layer of the resulted hollow fiber membranes would have less pore density but higher
pore sizes when coagulated at higher temperature. Those SEM
images in Fig. 6 were spun at various temperatures. Fig. 6(a)
and (b) are the images of pure CA hollow fiber membranes spun
at room temperature and 75 C, respectively. Fig. 6(b) shows
higher roughness and micropore surface than Fig. 6(a) under
a magnification of 50k times, whereas Fig. 6(c)(e) show the
CA/PEG hollow fiber membranes whose spun at various coagulant temperature. It is clearly that the pores or macrovoids and
pore size of membranes increased with increasing temperature.
In other words, either adding PEG or increasing coagulant temperature would be limited to changing or improving morphology
of membranes, but combining these two factors, the morphology
of membranes would be controlled more precisely.
Fig. 7 shows the cross-section of the CA/PEG hollow fibers.
In Fig. 7(a), when coagulating at 25 C without the addition of
PEG, CA hollow fiber appeared as a sponge-like structure with
few macrovoids near the inner surface. When blending with 14%
PEG, there was no macrovoids observable when coagulating at
25 C, as shown in Fig. 7(b). Macrovoids increased with the

coagulant temperature, as shown in Fig. 7(c) and (d). This can


be attributed to that higher coagulant temperature would lead to
lower viscosity and higher exchanging rate of DMF and water.
This gave rise to more macrovoids.
3.3. Effect of molecular weights and content of PEG and
coagulant temperature on the performance
Our previous work reported that a higher coagulant temperature would strongly affect the permeation performance of CA
hollow fiber membranes [18]. Based on these results, the effect of
PEG additive combining with spinning parameters were further
investigated in this present work.
Most membrane manufactures characterize their products by
the pore size or molecular weight cut-off value which is usually
obtained by measuring the retention (R) of macromolecules with
a series of hydrodynamic diameters or molecular weights. The
pure water permeability (PWP) and the retention of dextran can
be also regarded as the indices for the permeation performance
of the hollow fiber [48]. Thus, we measured these two terms to
evaluate the permeation performance of CA/PEG hollow fiber
membranes.
As mentioned in previous section, the thickness of the membrane and the morphology depend on the molecular weight and

218

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

Table 3
Effect of various content and molecular weight of PEG additive on the performance (the membranes were spun at 25 C)
PEG content (PEG vs. CA ratio)

0
7%
7%
7%
14%
14%
14%
21%
21%
21%

PEG (MW)

Pure CA
1k
10k
40k
1k
10k
40k
1k
10k
40k

PWP (L/m2 h atm)

5.6
4.4
3.9
12.4
10.1
8.6
18.4
15.9
14.0

concentration of PEG in the dope. Table 3 shows that the PWP


and R also depended on the molecular weight and concentration of PEG in the dope. The PWP and R of pure CA could
not be measured with our ultrafiltration experimental device.
This is because of the original CA has a sponge-like structure in
sub-layer and it is also found that both the inner and outer skin
layers are dense and smooth, as shown in the SEM images in
Figs. 5(a) and 6(a).
Figs. 5(b) and 6(b) show that the CA/PEG hollow fiber coagulated at 25 C exhibit an outer skin layer with micropores and
rough surface and an inner skin with dense and smooth surface.
For CA, water is a weak nonsolvent, which means that coagulation occurs slowly when the polymer solution is brought into
contact with water, and the resulting membranes have a spongelike sub-layer and almost no pore in on the surface. Although
PEG additive is water-soluble, the bore liquid in the lumen of
nascent hollow fiber membranes cannot dissolve all PEG at once.
When the nascent hollow fiber plunged into the coagulation bath,
a great quantity of water would replace DMF in the nascent hollow fiber and dissolve part of PEG on the outer surface of nascent
hollow fiber, hence the outer skin exhibited more micropores and
rough surface than the inner skin.
Table 3 also shows that PWP increases with increasing PEG
content and decreased with increasing molecular weight. It

Retention (dextran MW)


10.2k

35k

66.7k

89.5
88.8
90.3
85.8
87.3
89.3
83.0
84.1
85.9

96.2
94.7
95.2
94.6
93.5
94.4
89.3
90.5
90.4

99.5
97.4
97.7
95.7
94.5
95.3
95.1
94.5
95.2

seems that the PEG content is more influential on the permeation performance than the molecular weight of PEG. Higher PEG
content would lead to higher PWP. This is because that PEG is
a pore-forming agent that creates pores in the polymer matrix
of CA. On the other hand, higher molecular weight would be
less soluble and more difficult to effuse out than a lower molecular weight of PEG. Therefore the resulting CA hollow fiber
would have less but larger pores. It has been generally accepted
that performance of hollow fiber membranes were mainly affected by both of inner and outer skin. Micropore was observed
on the outer surface but not on the inner surface, even though
the PWP increased from 0 to 18.4 L/m2 h atm. The retention of
dextran slightly decreased with the increase of PEG content but
increased with the increase of the molecular weight of PEG.
This implies that the size of micropores increased with the PEG
content. However, the number of micropore decreased with the
increase of the molecular weight of PEG.
During the coagulation of hollow fiber, a substantial portion
of PEG may be entrapped in the pores and the entrapping amount
depends on the molecular weight of PEG. When the hollow fibers
were coagulated at lower temperature, the resulting membranes
exhibited dense structure, higher tensile force, and lower PWP.
By destroying the dense skin in either surface, the PWP can
be improved while retaining the retention. This can be achie-

Table 4
Effect of the coagulant temperature on the performance of CA/PEG blend hollow fiber membrane
PEG (MW and content) (PEG vs. CA ratio)

Coagulant temperature ( C)

PWP (L/m2 h atm)

Retention (dextran MW)


10.2k

35k

66.7k

Pure CA

25
50
75

8.4
15.6

87.6
78.7

93.8
88.2

97.2
93.1

CA + PEG (1k, 14%)

25
50
75

12.4
25.8
45.9

85.8
83.5
72.3

94.6
90.2
85.3

95.7
93.1
90.5

CA + PEG (10k, 14%)

25
50
75

10.1
30.1
55.3

87.3
75.3
59.3

93.5
84.6
74.9

94.5
92.4
90.2

CA + PEG (40k, 14%)

25
50
75

8.6
35.6
62.5

89.3
74.9
41.9

94.4
80.8
65.3

95.3
90.6
88.5

W.-L. Chou et al. / Separation and Purification Technology 57 (2007) 209219

ved by using higher coagulant temperature combining with the


addition of PEG additive. Table 4 illustrates the effect of the
external coagulant temperature on the performance of. CA/PEG
hollow fiber membranes. The PWP was progressively increased
to 62.5 L/m2 h atm at the highest coagulant temperature (75 C),
which was about 47 times of that coagulated at 25 C. It is also
found that the PWP was decreased with increasing molecular
weight. This is because that larger PEG additive in the CA/PEG
membranes can be extracted readily at 75 C than at 25 C. The
results also show that the increase in PWP did not sacrifice to
much the retention for 66.7 kDa dextran, which dropped from
97% to 88%.
4. Conclusion

[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]

Asymmetric cellulose acetate hollow fiber membranes were


prepared by dry/wet spinning process from dope composed
of cellulose acetate, N,N-dimethylformamide, and polyethylene
glycol. The effect of the molecular weight and the content of
PEG in the dope combining with external coagulant temperature on the morphology and performance were investigated.
The addition of PEG can suppress the formation of macrovoids.
These phenomena became obvious with increasing content and
molecular weight of PEG. The PWP increased with the PEG
content and decreased with the increase of the molecular weight
of PEG. In addition, higher coagulant temperature also increased
the PWP and decreased the retention of membranes. Combining simultaneously the addition of PEG and higher coagulant
temperature, the resulting CA hollow fiber membranes would
improve greatly the permeation performance that would effectively remove the toxic substances such as 2 -microglobulin from
the blood.

[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]

References
[1] G.S. Loeb, S. Sourirajan, Adv. Chem. Ser. 38 (1963) 117.
[2] M.J. Kell, R.D. Mahoney, GB Pat., 2,000,722 (1979).
[3] R.E. Kesting, Synthetic Polymeric MembranesA Structure Perspective,
Wiley, New York, 1985.
[4] T. Tsuruta, T. Hayashi, K. Kataoka, K. Ishihara, Y. Kimura (Eds.), Biomedical Applications of Polymeric Materials, CRC Press, Boca Raton, FL,
1993.
[5] R. Deppisch, M. Storr, R. Buck, H. Gohl, Sep. Purif. Technol. 14 (1998)
241.
[6] D. Gustav, S. Lothar, A. Wolfgang, US Pat. 5,403,485 (1995).
[7] Teijin Ltd., EP 0,697,242 (1996).

[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]

219

J.J. Qin, Y. Li, L.S. Lee, H. Lee, J. Membr. Sci. 218 (2003) 173.
S.A. Mckelvey, D.T. Clausi, W.J. Koros, J. Membr. Sci. 124 (1997) 223.
T. Liu, D. Zhang, S. Xu, S. Sourirajan, Sep. Sci. Technol. 27 (1992) 161.
J.J. Qin, T.S. Chung, J. Membr. Sci. 157 (1999) 35.
T.S. Chung, E.R. Katchniski, J. Appl. Polym. Sci. 65 (1997) 1555.
T.S. Chung, S.K. Teoh, X. Hu, J. Membr. Sci. 133 (1998) 161.
T.S. Chung, E.R. Kafchinski, P. Foley, J. Membr. Sci. 75 (1992) 181.
M. Henmi, T. Yoshioka, J. Membr. Sci. 85 (1993) 129.
T.S. Chung, E.R. Kafchinski, R. Vora, J. Membr. Sci. 88 (1994) 21.
W.L. Chou, M.C. Yang, Polym. Adv. Technol. 16 (2005) 1.
W.L. Chou, D.G. Yu, M.C. Yang, J. Polym. Res. 12 (2005) 219.
W.L. Chou, M.C. Yang, J. Membr. Sci. 250 (2005) 259.
M.C. Yang, M.T. Chou, J. Membr. Sci. 116 (1996) 279.
P. Aptel, N. Abidine, F. Ivaldi, J.P. Lafaille, J. Membr. Sci. 22 (1985)
199.
X. Miao, S. Sourirajan, H. Zhang, W.W.Y. Lau, Sep. Sci. Technol. 31 (1996)
141.
G.C. East, J.E. McIntyre, V. Rogers, S.C. Senn, Proceedings of the Fourth
BOC Priestly Conference, Royal Society of Chemistry, London, 1986, p.
63.
H. Kim, Y.I. Park, J. Jagel, K.H. Lee, J. Appl. Polym. Sci. 57 (1995) 1637.
H.A. Tsai, D.H. Huang, S.C. Fan, Y.C. Wang, C.L. Li, K.R. Lee, J.Y. Lai,
J. Membr. Sci. 198 (2002) 245.
M. Khayet, Chem. Eng. Sci. 58 (2003) 3091.
Z.L. Xu, F.A. Qusay, J. Membr. Sci. 233 (2004) 101.
J.H. Kim, K.H. Lee, J. Membr. Sci. 138 (1998) 153.
E.W. Merrill, S. Wan, E.W. Salzman, Trans. Am. Soc. Artif. Intern. Organs
20 (1986) 1517.
Y. Liu, G.H. Koops, H. Strathmann, J. Membr. Sci. 223 (2003) 187.
D. Li, T.S. Chung, R. Wang, J. Membr. Sci. 243 (2004) 155.
H. de Balmann, V. Sanchez, Int. Chem. Eng. 32 (1992) 665.
M. Dubois, K.A. Gilles, J.K. Hamilton, P.A. Rebers, F. Smith, Nature 168
(1951) 167.
M. Dubois, K.A. Gilles, J.K. Hamilton, P.A. Rebers, F. Smith, Anal. Chem.
28 (1956) 350.
Y.S. Park, J. Won, Y.S. Kang, Langmuir 16 (2000) 9662.
C.A. Smolders, A.J. Reuvers, R.M. Boom, I.M. Wienk, J. Membr. Sci. 73
(1992) 259.
R.M. Boom, I.M. Wienk, T. van den Boomgaard, C.A. Smolders, J. Membr.
Sci. 73 (1992) 277.
J.Y. Lai, F.C. Lin, C.C. Wang, D.M. Wang, J. Membr. Sci. 118 (1996) 49.
K. Kimmerle, H. Strathmann, Desalination 79 (1990) 283.
M.J. Han, Desalination 121 (1999) 31.
I.M. Wienk, R.M. Boom, M.A.M. Beerlage, A.M.W. Bulte, C.A. Smolders,
H. Strathmann, J. Membr. Sci. 113 (1996) 361.
Z.L. Xu, T.S. Chung, Y. Huang, J. Appl. Polym. Sci. 74 (1999) 2220.
Z.L. Xu, F.A. Qusay, J. Appl. Polym. Sci. 91 (2004) 3398.
M.J. Han, S.T. Nam, J. Membr. Sci. 202 (2002) 55.
L.Y. Lafreni`ere, F.D.F. Talbot, T. Matsuura, S. Sourirajan, Ind. Eng. Chem.
Res. 26 (1987) 2385.
B. Jung, J.K. Yoon, B. Kim, H.W. Rhee, J. Membr. Sci. 243 (2004) 45.
K. Devanand, J.C. Selser, Macromolecules 24 (1991) 5943.
H. de Balmann, P. Aimar, V. Sanchez, J. Membr. Sci. 45 (1989) 17.

You might also like