Pervaporation Separation of Isopropanol-Water Mixtures Through Crosslinked Chitosan Membranes

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Membrane Science 262 (2005) 9199

Pervaporation separation of isopropanol/water mixtures through


crosslinked chitosan membranes
D. Anjali Devi a , B. Smitha b , S. Sridhar b , T.M. Aminabhavi a,
b

a Membrane Separations Division, Center of Excellence in Polymer Science, Karnatak University, Dharwad 580 003, Karnataka, India
Membrane Separations Group, Chemical Engineering Division, Indian Institute of Chemical Technology, Hyderabad 500 007, Andhra Pradesh, India

Received 16 August 2004; received in revised form 23 March 2005; accepted 26 March 2005
Available online 17 May 2005

Abstract
Chitosan membrane having 84% degree of deacetylation was crosslinked with toluene-2,4-diisocyanate and tested for the dehydration of
isopropanol by the pervaporation method. Pure and crosslinked membranes were characterized by Fourier transform infrared spectroscopy
and wide angle X-ray diffraction to study the intermolecular interactions and crystallinity, respectively. Dynamic mechanical thermal analyses
were undertaken to assess the thermal and mechanical stabilities of the membranes. Sorption studies were performed in water, isopropanol
as well as different composition feed mixtures to understand the polymerliquid interactions and pervaporation separation mechanism. The
membrane appears to have a good potential for breaking the aqueous azeotrope of 87.5 wt.% isopropanol with a high selectivity of 472 and
a substantial water flux of 0.39 kg/m2 h 10 m. The influence of operating parameters such as feed composition, membrane thickness and
permeate pressure on membrane performance like flux and selectivity was investigated.
2005 Elsevier B.V. All rights reserved.
Keywords: Pervaporation; Isopropanol/water azeotrope; Chitosan membrane; TDI

1. Introduction
Pervaporation (PV) has been widely used for the dehydration of aqueousorganic mixtures [13]. Chitosan (CS)
or poly(d-glucosamine), a natural biopolymer, obtained by
the deacetylation of chitin, has many inherent characteristics
such as hydrophilicity, biocompatibility, antibacterial properties, and a remarkable affinity for many substances [2]. Many
reports are available in the earlier literature on the PV separation of wateralcohol mixtures using CS membranes [311].
Detailed description of the preparation of acetic acid complex, carboxymethyl, cyanoethyl, amidoxime, carboxyethyl,
sulfonated and phosphorylated CS membranes are given
elsewhere [5,6]. Chitosan has many biomedical applications
This article is CEPS Communication #51.

Corresponding author. Tel.: +91 836 2771275;


fax: +91 836 2771275.
E-mail address: aminabhavi@yahoo.com (T.M. Aminabhavi).

0376-7388/$ see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2005.03.051

[1215], but in dehydration studies on aqueousorganic mixtures, its hydroxyl and amino groups can be modified easily.
Chitosan swells in water and hence, it can be crosslinked
by a suitable agent to improve its mechanical strength and
selectivity during PV experiments. The separation selectivity for wateralcohol mixture for CS itself is not high because
its free amine form is water insoluble. Since CS has both hydroxyl and amine groups, it can be modified chemically into
many forms. Glutaraldehyde (GA) has been the most commonly used crosslinking agent that forms Schiff base when
reacted with CS [13]. Reports on the PV performance of GA
crosslinked CS membranes [14] show higher selectivity and
permeation flux than those of glyoxal and terephthalaldehyde
crosslinked membranes, which results in higher flexibility induced by GA to the polymeric chain. Mochizuki et al. [16]
studied the PV separation of water-ethanol mixtures through
CS membranes neutralized by various acids and observed
high selectivity to water. They treated acetic acid complex
membranes in NaOH solution to make CS neutral and inves-

92

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199

tigated the effect of metal species such as cobalt on selectivity.


Divalent ions including SO4 2 were effective for increasing
the selectivity [17]. Dense CS membranes that are ionically
crosslinked by sulfuric acid showed the highest selectivity
with lower permeation rates.
Isopropanol, a widely used solvent in chemical and pharmaceutical industries, is known to form an azeotrope with
water, a characteristic that creates difficulties in its recovery by the conventional distillation [18]. The application of
PV as a means to achieve dehydration of solvents has received a widespread attention from chemical, petrochemical
and pharmaceutical industries. However, the recent technology improvements have led to a rapid commercialization of
several novel membranes that are economical, safe and clean
to be used in the separation of azeotropic and close by boiling liquid mixtures by the PV technique [19]. Considerable
literature on the successful dehydration of isopropanol by
PV is available. Table 1 gives a comparison of PV performance data of other membranes used in PV separation of
isopropanolwater mixtures [2024]. It is realized that performance of pure CS membrane is not satisfactory due to
larger free volume between the molecular chains. The membrane properties of CS can be improved by blending with
other polymers [25] or by incorporating a high selectivity
zeolite into the membrane [26]. In an effort to investigate
the effect of crosslinked structure on the PV performance of
CS membrane, we have used for the first time, toluene-2,4diisocyanate (TDI) as a crosslinking agent and the membrane
is used in the PV dehydration of isopropanol. The results are
compared with existing membranes.

to be the optimum reaction time, but higher concentrations


of TDI produced brittle membranes. The crosslinked membranes were washed with acetone and finally vacuum dried
for a period of 12 h. The dried films were utilized in PV
experiments. The stability of the membrane was analyzed
by bending the membrane before and after PV studies. The
experiment performed for the duration of 3 months during
which the membrane was found to be stable. After this period,
the membrane was bent to ensure its mechanical stability and
it was noted that despite bending completely the membrane
did not break. Hence, the membrane durability and stability
appears to be reasonably good [27].
2.3. Membrane characterization
2.3.1. FT-IR studies
Pure and crosslinked CS membranes were scanned in
the range 4004000 cm1 wave numbers using Nicolet-740,
Perkin-Elmer-283B FT-IR spectrophotometer.
2.3.2. Ion exchange capacity (IEC)
To determine the total number of interacting groups
present in the membranes, unmodified and crosslinked chitosan of similar weight were soaked in 50 mL of 0.01 M
sodium hydroxide solution for 12 h at ambient temperature.
Then, 10 mL of solution was titrated against 0.005 M sulfuric acid. The sample was regenerated with 1 M hydrochloric
acid, washed free of acid with water and dried to constant
weight. The IEC was calculated according to the equation
IEC =

B P MNaOH 5
m

(1)

B P (MH2 SO4 2) 5
m

(2)

2. Experimental

or

2.1. Materials

IEC =

Isopropanol, glacial acetic acid and TDI of purities


>99.9% were purchased from Loba Chemicals, Mumbai, India. Chitosan (flakes) of 84% degree of deacetylation was
received as a gift sample from the local market. Deionized
water of conductivity 20 S/cm was generated in the laboratory itself.

where B is the amount of 0.005 M sulfuric acid used to neutralize unmodified chitosan, P is the amount of 0.005 M sulfuric acid used to neutralize the crosslinked membranes, 5
represents the factor corresponding to the ratio of the amount
of NaOH taken to dissolve the polymer to the amount used
for titration, and m is the sample mass in g.

2.2. Membrane preparation

2.3.3. XRD analysis


A Siemens D 5000 powder X-ray diffractometer was used
to assess the solid-state morphology of the crosslinked CS
wavelength was
in a powder form. The X-rays of 1.5406 A
generated by a Cu K source. The angle of diffraction was
varied from 0 to 65 to identify the change in the crystal structure and intermolecular distances between the intersegmental chains after crosslinking.

Chitosan membranes were prepared by solution casting


and solvent evaporation method. The 3 wt.% solution of CS in
2% (v/v) aqueous acetic acid was prepared, stirred and filtered
to remove the undissolved matter. A bubble-free solution was
cast on a clean glass plate/petridish to the desired thickness
and dried in atmospheric conditions at room temperature followed by vacuum drying for a period of 5 h at the elevated
temperature (50 C) in an oven to remove the last traces of
solvent. The membranes thus prepared were crosslinked by
using TDI in a hexane bath containing two to three drops
of dibutyl tin dilaureate as a catalyst. Nearly 6 h were found

2.3.4. Thermal analysis by DMTA


The dynamic mechanical properties of CS specimen (storage and loss moduli, tan ) were measured by DMTA 1V instrument (Rheometric Scientific, USA) in tensile mode at a

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199

93

Table 1
Comparison of the pervaporation performance of the present crosslinked chitosan membrane with the literature data for waterisopropanol mixtures
Membrane

Temperature
( C)

Thickness
(m)

Chitosan/TDI
PASA
PASA-bromo propane
PASA-bromooctane
PASA-(1-bromo-2phenylethane)
PASAmethylbromoacetate
PASA-ethylbromoacetate
PVA crosslinked with
glutaraldehyde
PVA crosslinked with
citric acid
Composite membrane of
NaAlg and chitosan
NaAlg
NaAlg/GG-g-pAAm
NaAlg/GG-g-pAAm
NaAlg/PVA (75:25)
NaAlg/PVA (50:50)
NaAlg/PVA (25:75)

30
20

50
30
30
30
30

Water in feed
(wt.%)
8.4
10

30

Flux (kg/m2 h)

Normalized
flux (10 m)

Selectivity

Reference

0.079
0.003
0.004
0.009
0.006

0.39
0.008
0.013
0.026
0.019

472
3686

111

Present work
[20]

0.008

0.024

116

30

30

10

0.003
0.194

0.010

30

05

0.095

60

45

10

0.554

2.493

30

10

0.058
0.062
0.043
0.025
0.034
0.039

10
10

[21]

741
2010

[22]

411
796
891
195
119
91

[23]

[24]

PASA: poly(amide sulfonamide); PVA: poly(vinyl alcohol); NaAlg: sodium alginate; GG: guar-gum; AA: acrylamide.

frequency of 10 or 1.0 Hz at the heating rate of 3 C/min to determine the viscoelastic behavior of the crosslinked and pure
CS membranes in nitrogen atmosphere over the temperature
range 35200 C.
2.3.5. Sorption studies
Weighed samples of cross-linked chitosan films (3 cm diameter) were soaked in pure water and isopropanol as well
as binary mixtures of different compositions. The films were
taken out at different soaking time intervals and weighed immediately after carefully wiping out the excess liquid to determine the amount of liquid sorbed by the film at that particular
time duration, t. The process was repeated until the films attained steady state as indicated by a constant weight after a
certain period of soaking time. The degree of swelling was
calculated from the equation:
Ms
degree of swelling =
Md

(3)

where Ms is the mass of the swollen polymer in g and Md


is the mass of the dry polymer in g. The % sorption was
calculated from the equation


M s Md
100
(4)
% sorption =
Md
2.4. Pervaporation experiments
2.4.1. Inuence of operating conditions
Pervaporation experiments were carried out on a 100 mL
batch level instrument with an indigenously constructed
manifold (Fig. 1a) operated at a vacuum as low as 0.05 mmHg
in the permeate line. Membrane area in the PV cell assembly

(Fig. 1b) was approximately 20 cm2 . The experimental


procedure is described in detail elsewhere [28]. Permeate
samples were collected after a period of 810 h. Tests were
carried out at room temperature (30 2 C) and repeated
twice using fresh feed solution to check reproducibility. The
collected permeates were weighed after allowing them to
attain room temperature in a Sartorius electronic balance
(accuracy, 104 g) to determine the flux. Feed as well as
permeate mixtures were analyzed by gas chromatography
(Nucon, Model 5765) to evaluate the membrane selectivity.
2.4.2. Flux and selectivity
In pervaporation, flux, J of a given species, say faster permeating component, i of a binary liquid mixture comprising
of i (water) and j (isopropanol) is given by
Ji =

Wi
At

(5)

Here Wi represents the mass of water in permeate (kg), A


is the membrane area (m2 ) and t represents the permeation
time (h). In the present study, even though different membrane thicknesses were utilized, the flux has been normalized
and reported for the thickness of 10 m for the effect of feed
composition and effect of permeate pressure. Membrane selectivity, is the ratio of permeability coefficients of water to
that of isopropanol, which is calculated from their respective
concentrations in feed and permeate as given below,
=

y(1 x)
x(1 y)

(6)

where y is the permeate weight content of water (%) and x is


the feed weight content of water (%).

94

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199

Fig. 1. Schematics of laboratory pervaporation unit.

2.5. Analytical procedure


Feed and permeate samples were analyzed using a Nucon Gas Chromatograph (GC Model 5765) installed with
thermal conductivity detector (TCD) and packed column
of 10% DEGS on 80/100 Supelco port of 1/8 in. i.d. and
2 m length. The oven temperature was maintained at 70 C
(isothermal), while the injector and detector temperatures
were maintained at 150 C each. The sample injection size
was 1 L and pure hydrogen was used as a carrier gas at a
pressure of 1 kg/cm2 . The GC response was calibrated for this
particular column and conditions with known compositions
of isopropanolwater mixtures. Calibration factors were fed
into the software to obtain the correct analysis for unknown
samples.

3. Results and discussion


Chitosan membrane was chosen for the PV studies of isopropanolwater mixtures on the basis of the
close proximity of its Hansens solubility parameter value

(43.04 J1/2 /cm3/2 ) [29] to that of water (47.9 J1/2 /cm3/2 ) [30]
as well as other useful features, such as hydrophilicity, good
mechanical strength and chemical resistance. Scheme 1 represents the crosslinking reaction between amino groups of
chitosan with the isocyanate group of TDI, resulting in the
formation of urea linkage. This is possible because the amino
group of chitosan is a stronger nucleophile than its hydroxyl
group. During the crosslinking reaction, the amino group of
chitosan interacts with the carbonyl group of TDI, resulting
in the formation of urea linkages. An estimation of the number of groups present before and after crosslinking gives an
idea of the extent of crosslinking.
3.1. Membrane characterization
3.1.1. Ion exchange capacity (IEC)
The amount of residual amine and hydroxyl groups present
after crosslinking was estimated from the IEC studies. It
was noted that unmodified chitosan showed an IEC of
0.42 mequiv./g, whereas the crosslinked polymer exhibited
an IEC of 0.2 mequiv./g. The IEC is equivalent to the total
number of free amino groups (considering the fact that amino

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199

95

Scheme 1. Crosslinking reaction of chitosan with 2,4-toluylene diisocyanate.

groups are more interactive than hydroxyl groups), R-NH2


present in the membrane which decreased upon crosslinking [31]. This shows that almost 50% of the amine groups
present in the unmodified chitosan have now formed the
crosslinks with TDI. Scheme 1 represents the crosslinking
reaction occurring between chitosan and TDI. The occurrence of crosslinking is proved by IEC and FT-IR studies.
Hence, TDI will establish a linkage with chitosan through
urea formation as confirmed by FT-IR (Fig. 2). To the best of
our knowledge, it is the first kind of study wherein TDI is employed as a crosslinking agent and the membrane could withstand the solvent environment and PV condition employed in
this study.
3.1.2. FT-IR studies
Fig. 2 shows the FT-IR spectra of the pure and the
crosslinked CS membrane. The spectrum of crosslinked CS
film shows peaks in the range 700850 cm1 , indicating the
presence of benzene ring. Reduced number of peaks in the
range of 10001200 cm1 compared to the spectra of pure
CS is due to the vibration of C O bond formed by the urethane linkage. The peak observed at 1640 cm1 shows the
presence of urea. The spectra of pure CS shows a broad peak
at wavenumbers 15701655 cm1 , which indicates the presence of amide I and II. The wavenumber 2833 cm1 corresponds to CH2 stretching, whereas that of free hydroxyl
group is observed at 3450 cm1 . FT-IR analysis confirms
the crosslinking of chitosan by the reaction of hydroxyl and
amine groups of CS with the carbonyl group of toluene-2,4diisocyanate.

Fig. 2. FT-IR spectra of (a) unmodified chitosan membrane and (b)


crosslinked chitosan membrane.

3.1.3. XRD studies


From the spectra obtained for pure and crosslinked CS
shown in Fig. 3, it is observed that XRD patterns of both
pure and crosslinked CS membranes appear to be semicrystalline. The broad peaks observed in the XRD pattern around
10 of 2 indicate the average intermolecular distance of the
amorphous part and relatively sharp semicrystalline peaks
are centered at around 20 of 2. From these observations, it
can be concluded that the average intermolecular distances
in CS and crosslinked CS are the same. It can be seen that

Fig. 3. XRD spectra of (a) unmodified chitosan membrane and (b)


crosslinked chitosan membrane.

96

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199

3.1.5. Swelling characteristics


Swelling data of the crosslinked CS membrane in
isopropanolwater mixtures of different compositions are
presented in Table 2. The % sorption computed from Eq.
(4) in the feed composition at the azeotropic concentration
containing 12.5 wt.% of water was found to be low i.e., 15,
but sorption increased steadily with an increase in water composition, which indicates that CS membrane interacts extensively with water and is capable of selective absorption and
permeation of water molecules to bring about the separation.
However, absorption of a large amount of water at high feed
water concentrations could cause the membrane to swell excessively leading to a decrease in membrane selectivity.
3.2. Inuence of operating conditions
It is well known that separation characteristics of a membrane depend upon the interaction between solvent to be separated and the membrane matrix. Hydrophilic membrane like
CS can develop hydrogen bond interaction with water leading to preferential sorption and diffusion of water through the
barrier membrane [33]. The influence of feed composition,
membrane thickness and permeate pressure on membrane
performance has been examined in detail.

Fig. 4. DMTA tracings of (a) unmodified chitosan membrane and (b)


crosslinked chitosan membrane.

there are two distinct bands with their maxima at 2 = 79


and 2 at 20 , which are related to two types of crystals:
crystal 1 and crystal 2 [32]. Crystal 1, which corresponds
to the peak at 9 is responsible for the separation, since
it comprises functional groups like NH2 and OH and
has undergone significant change after crosslinking. A re for
duction in the effective d-spacing value from 12.07 A

pure CS to 7.8 A for the crosslinked CS gives a clear picture of the shrinkage in cell size or inter-segmental spacing, which would improve the selective permeation of CS
membrane.
3.1.4. Thermal analysis
Thermal properties of the pure and crosslinked CS
membranes were examined by DMTA. It was of particular interest to estimate how the thermal transition of
CS varied with crosslinking. DMTA tracings of CS and
its crosslinked structure are shown in Fig. 4a and b,
which reveal that pure CS has a glass transition temperature, Tg at 112.6 C, which gets enhanced to 176.65 C
for the crosslinked membrane, indicating that crosslinking with TDI might have caused a reduction in segmental
mobility.

3.2.1. Effect of feed composition


The influence of feed concentration on PV properties
was studied. Experiments were carried out after soaking the
membrane for 12 h in the feed mixture to ensure equilibrium.
The PV performance of chitosan membrane crosslinked with
TDI was studied for various feed compositions comprising
of 4.41539.672 wt.% water, while keeping other operating
parameters such as permeate pressure and membrane
thickness constant at 2 mmHg and 50 m, respectively.
Expectedly, a rise in feed concentration of water produced
an increase in water flux from 0.28 to 1.24 kg/m2 h (see
Table 2 and Fig. 5). Mass transport through a hydrophilic
CS membrane occurs by solutiondiffusion mechanism
[34]. As shown in Scheme 1, the structure of crosslinked CS
membrane has polar urethane and urea linkages. Moreover,
the residual OH and NH2 groups will also be available
for the interaction with water molecules through hydrogen
bonding. In addition to sorption data of the membranes
obtained for the binary feed mixtures presented in Table 2,
it was found that the crosslinked CS membrane has shown a
high degree of sorption in pure water (121%), but relatively
negligible sorption in pure isopropanol (0.05%). According
to the solutiondiffusion mechanism, membrane selectivity
depends on the partition of two components between the feed
solution and the upstream layer of the membrane as well as
on the difference of their diffusivities across the membrane.
The preferential affinity of the membrane towards water
causes swelling, which allows a rapid permeation of feed
molecules. The extent of sorption correspondingly rises with
an increase in feed water concentration resulting in enhanced
flux. However, increased swelling of the membrane has a

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199

97

Table 2
Effect of feed composition on separation performance of the crosslinked chitosan membrane (permeate pressure 2 mmHg; membrane thickness 50 m)
Feed composition
Water (x)

Isopropanol (1 x)

4.415
8.385
10.633
16.262
28.581
39.672

95.585
91.615
89.367
73.105
71.419
60.328

% Sorption in membrane

5.4
6.6
15.55
25.0
27.9
31.2

Permeate composition
Water (y)

Isopropanol (1 y)

96.02
97.73
98.08
98.49
98.70
88.89

3.98
2.27
1.92
1.51
1.3
11.11

negative impact on membrane selectivity since the swollen


and plasticized upstream membrane layer allows some
of the isopropanol molecules to escape into the permeate
side along with water molecules, indicating a drop in
selectivity from 523.33 to 12.25 as seen in Fig. 5. It is worth
mentioning that the membrane showed promising results for
dehydrating feeds containing 528% water. Selectivity,
is good for residual water concentrations below 28%. Thus,
azeotropic composition (87.5 wt.%) of isopropanol was
easily broken down by PV process. About 97.73% of water
was obtained in permeate with a selectivity of 472 and a flux
of 0.39 kg/m2 h. The results of this work show comparable
fluxes and selectivities with respect to other membranes as
compared in Table 1 for dehydrating isopropanol.
3.2.2. Effect of membrane thickness
The effect of membrane thickness on water flux and
selectivity were evaluated at constant feed composition
(azeotropic) and permeate pressure (2 mmHg) by casting
membranes of thickness ranging from 25 to 125 m. With an
increase in membrane thickness, a gradual reduction in flux
from 0.265 to 0.032 kg/m2 h can be clearly evidenced from
Fig. 6. Even though the availability of polar groups enhances

Fig. 5. Effect of feed composition on pervaporation performance of


crosslinked chitosan membranes (membrane thickness 50 m, permeate
pressure 1 mmHg).

Selectivity, =

523.33
472.00
429.25
336.27
190.31
12.25

y(1x)
x(1y)

Normalized flux (kg/m2 h)

0.28
0.39
0.45
0.61
0.94
1.24

with an increase in membrane thickness, the flux decreases


since diffusion of feed is retarded due to increased resistance
to mass transfer. The permeate concentration of water varied
from 88.89 to 96.02 wt.%, which means that selectivity has
increased from 12.2 to 523.3. In the PV experiment, upstream
layer of the membrane is swollen and plasticized due to sorption of feed liquid thus, allowing the unrestricted transport
of feed components. In contrast, the downstream layer is virtually dry due to continuous evacuation in the permeate side
and therefore, this layer forms the restrictive barrier, which
allows only the interacting and smaller size molecules such as
water to pass through. It is expected that the thickness of the
dry layer would increase with an increase in the overall membrane thickness, thereby resulting in an improved selectivity
as observed in the present case.
3.2.3. Effect of permeate pressure
The effect of permeate pressure on membrane performance of the crosslinked chitosan membranes was studied in
the range 0.520 mmHg at a constant membrane thickness of
50 m. As the permeate pressure decreases, the driving force
for diffusing molecules decreases, resulting in higher permeate rates. Fig. 7 exhibits a considerable increase in flux from
0.06 to 0.58 kg/m2 h. Similarly, selectivity decreased from
745 to 22. With the increasing permeate pressures, diffusion
of the feed molecules through the membrane, which is the

Fig. 6. Influence of membrane thickness on pervaporation performance of


crosslinked chitosan membranes (feed composition: azeotropic, permeate
pressure: 2 mmHg).

98

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199

tively combined with distillation to constitute an economical


hybrid process to achieve the desired purity levels of isopropanol.

Acknowledgments

Fig. 7. Effect of downstream pressure on pervaporation performance of


crosslinked chitosan membrane (feed composition: azeotropic, membrane
thickness: 50 m).

rate-determining step, becomes slow whereas, high vacuum


exerts a larger driving force. Under lower vacuum conditions,
the volatility of feed components governs the separation selectivity of the membrane. Isopropanol, being more volatile
than water, permeates competitively, thus lowering the
selectivity.

4. Conclusions
Hydrophilic pervaporation of this study has shown an
immense potential as a viable technique for the dehydration of isopropanol. Toluene diisocyanate-crosslinked chitosan membrane, used for the first time in the literature,
could easily break the azeotrope of isopropanolwater mixture, indicating that the membrane acts as a third phase and
selectively allows water molecules to pass through due to
preferential affinity. TDI appears to be a useful and novel
crosslinking agent by rendering chitosan highly selective
to water without compromising heavily on the flux. Characterization of the membranes by FT-IR and XRD confirmed the crosslinking of chitosan. The XRD results of
the crosslinked chitosan showed a reduction in the effective d-spacing value, which is an indication of shrinkage
in the cell size, which would in turn, improve the selective permeation property of the membrane. Crosslinking did
not bring about any major changes in the thermal and mechanical stability of the membrane. With an increase in feed
water concentration, the membrane performance was found
to be affected substantially due to an increase in the extent of swelling of the polymer, thereby resulting in an increase of flux, but at the expense of selectivity. Increasing
membrane thickness decreased the flux, but improved the
separation selectivity. On the other hand, higher permeate
pressure could result in a reduction of both flux and selectivity. It is demonstrated that pervaporation could be effec-

Dr. T.M. Aminabhavi and Miss Anjali Devi thank the


University Grants Commission, New Delhi (Grant No.
F1-41/2001/CPP-II) for a financial support to establish
the Center of Excellence in Polymer Science (CEPS). Dr.
K.V.S.N. Raju, Organic Coatings and Polymers Division for
DMTA analyses. Ms. C.L. Kalpana and Ms. G. Dhanuja,
Membrane Separations Groups, IICT, are thanked for
their assistance. The help from Mr. Saibabu of Design
Division for tracing work is gratefully acknowledged. Dr.
M. Ramakrishna, Head, Membrane Separations Group,
deserves thanks for his support and encouragement. This
work represents a collaborative research under the MoU
between IICT, Hyderabad and CEPS, Dharwad.

References
[1] R.Y.M. Huang, Pervaporation Membrane Separation Processes, Elsevier Science Publishers, Amsterdam, The Netherlands, 1991.
[2] B. Vijaya Kumar Naidu, K.S.V. Krishna Rao, T.M. Aminabhavi, Pervaporation separation of water + 1,4-dioxane and water + tetrahydrofuran mixtures using sodium alginate and its blend
membranes with hydroxyethylcellulosea comparative study, J.
Membr. Sci., in press.
[3] B. Vijaya Kumar Naidu, M. Sairam, K.V.S.N. Raju, T.M. Aminabhavi, Pervaporation separation of water + isopropanol mixtures using
nanocomposite membranes of poly(vinyl alcohol) and polyaniline, J.
Membr. Sci., in press.
[4] U.S. Toti, T.M. Aminabhavi, Different viscosity grade sodium alginate and modified sodium alginate membranes in pervaporation
separation of water + acetic acid and water + isopropanol mixtures, J.
Membr. Sci. 228 (2004) 198204.
[5] Y.M. Lee, Modified chitosan membranes for pervaporation, Desalination 90 (1993) 277290.
[6] J.W. Baek, E.M. Shin, Y.M. Lee, Sorption and diffusion of water
and alcohol in chitosan complex membranes, Polymer (Korea) 14
(1990) 273281.
[7] Y.M. Lee, E.M. Shin, Pervaporation separation of waterethanol
through modified chitosan membranes; chitosan acetic acid and metal
ion complex membranes, Polymer (Korea) 15 (1991) 182189.
[8] Y.M. Lee, E.M. Shin, S.T. Noh, Pervaporation separation
of waterethanol through modified chitosan membranes: carboxymethyl, cyanoethyl and amidoxime chitosan membranes,
Angew. Makromol. Chem. 192 (1991) 169181.
[9] R.Y.M. Huang, R. Pal, G.Y. Moon, Crosslinked chitosan composite
membrane for the pervaporation dehydration of alcohol mixtures and
enhancement of structural stability of chitosan/polysulfone composite
membranes, J. Membr. Sci. 160 (1999) 1730.
[11] Y.M. Lee, S.Y. Nam, D.J. Woo, Pervaporation of ionically surface
crosslinked chitosan composite membranes for water-alcohol mixtures, J. Membr. Sci. 137 (1997) 103110.
[12] J. Jegal, K.H. Lee, Chitosan membranes crosslinked with sulfosuccinic acid for the pervaporation separation of water/alcohol mixtures,
J. Appl. Polym. Sci. 71 (1999) 671675.

D. Anjali Devi et al. / Journal of Membrane Science 262 (2005) 9199


[13] X.P. Wang, Z.Q. Shen, F.Y. Zhang, Pervaporation separation of water/alcohol mixtures through hydroxypropylated chitosan membranes,
J. Appl. Polym. Sci. 69 (1998) 20352041.
[14] H. Ito, T. Shibata, Y. Noishiki, H. Inagaki, Formation of polyelectrolyte complexes between cellulose derivatives and their blood compatibility, J. Appl. Polym. Sci. 31 (1986) 24912500.
[15] J.H. Kim, J.Y. Kim, Lee, K.Y. Kim, C.S. Cho, Y.K. Sung,
Controlled-release drug delivery through crosslinked poly(vinyl alcohol)/chitosan blend membranes, Polymer (Korea) 15 (6) (1991)
695701.
[16] A. Mochizuki, S. Amiya, Y. Sato, H. Ogawara, S. Yamashita, Pervaporation separation of water/ethanol mixtures through polysaccharide
membranes; the permselectivity of the neutralized chitosan membrane and the relationships between its permselectivity and solid
state structure, J. Appl. Polym. Sci. 37 (1989) 33853398.
[17] J.H. Kim, Y.M. Lee, Synthesis and properties of dimethylaminoethyl
chitosan, Polymer 34 (1993) 19531957.
[18] Y.M. Lee, S.Y. Nam, B.R. Lee, D.J. Woo, K.H. Lee, J.M. Won,
B.H. Ha, Dehydration of alcohol solutions through crosslinked chitosan composite membranes; preparation of chemically crosslinked
chitosan composite membranes and ethanol dehydration, Membrane
(Korea) 20 (1996) 3743.
[19] A. Mochizuki, Y. Sato, H. Ogawara, S. Yamashita, Pervaporation
separation of water/ethanol mixtures through polysaccharide membranes; the permselectivity of the chitosan membranes, J. Appl.
Polym. Sci. 37 (1989) 33753384.
[20] W.H. Chan, C.F. Ng, S.Y. Lam-Leung, X. He, Water/alcohol separation by pervaporation through chemically modified poly(amidesulfonamide), J. Membr. Sci. 160 (1999) 7786.
[21] M.C. Burshe, S.A. Netke, S.B. Sawant, J.B. Joshi, V.G. Pangarkar,
Pervaporation dehydration of organic solvents, Sep. Sci. Technol. 32
(1997) 13351349.
[22] G.Y. Moon, R. Pal, R.Y.M. Huang, Novel two-ply composite membranes of chitosan and sodium alginate for the pervaporation dehydration of isopropanol and ethanol, J. Membr. Sci. 156 (1999)
1727.

99

[23] U.S. Toti, T.M. Aminabhavi, Pervaporation separation of water


and isopropanol through blend membrane of sodium alginate and
guar gum-grafted-polyacrylamide, J. Appl. Polym. Sci. 85 (2002)
20142024.
[24] M.D. Kurkuri, U.S. Toti, T.M. Aminabhavi, Synthesis and characterization of blend membranes of sodium alginate and poly(vinyl
alcohol) for the pervaporation separation of water/alcohol mixtures,
J. Appl. Polym. Sci. 86 (2002) 36423651.
[25] M. Muncha, A. Pawalak, Thermal analysis of chitosan and its blends,
Thermochim. Acta 427 (2005) 6976.
[26] X. Chen, H. Yang, Z. Gu, Z. Shao, Preparation and characterization
of HY zeolite-filled chitosan membranes for pervaporation separation, J. Appl. Polym. Sci. 79 (2001) 11441149.
[27] B. Smitha, S. Sridhar, A.A. Khan, Polyelectrolyte complexes of chitosan and poly(acrylicacid) as proton exchange membranes for fuel
cells, Macromolecules 37 (2004) 22342239.
[28] S. Sridhar, R. Ravindra, A.A. Khan, Recovery of monomethyl
hydrazine liquid propellant by pervaporation technique, Ind. Eng.
Chem. Res. 39 (2000) 24852490.
[29] R. Ravindra, K. Kameswara Rao, A.A. Khan, Solubility parameter of chitin/chitosan, Carbohydr. Polym. 36 (1998) 121
127.
[30] A.F.M. Barton (Ed.), CRC Handbook of Solubility Parameters and
other Cohesive Parameters, CRC Press, Boca Raton, FL, USA,
1983.
[31] K. Igarashi, Y. Nakano, Fundamental adsorption properties of chitosan gel particles prepared by suspension evaporation method, J.
Appl. Polym. Sci. 86 (4) (2002) 901906.
[32] R.J. Samuels, Solid state characterization of the structure of chitosan
films, J. Polym. Sci. Polym. Phys. Ed. 19 (1981) 10811105.
[33] S. Sridhar, G. Susheela, G. Jayasimha Reddy, A.A. Khan,
Crosslinked chitosan membranes: characterization and study of
dimethylhydrazine dehydration by pervaporation, Polym. Int. 50
(2001) 11561161.
[34] J.G. Wijmans, R.W. Baker, The solutiondiffusion model: a review,
J. Membr. Sci. 107 (1995) 121.

You might also like