Vous êtes sur la page 1sur 10

ARTICLE IN PRESS

Chemical Engineering Science 65 (2010) 4378–4387

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Water sorption and diffusion through saturated polyester and their


nanocomposites synthesized from glycolyzed PET waste with
varied composition
Sunain Katoch a, Vinay Sharma b, P.P. Kundu a,b,
a
Department of Chemical Technology, Sant Longowal Institute of Engineering and Technology, Sangrur, Punjab 148106, India
b
Department of Polymer Science and Technology, University of Calcutta, 92, A.P.C. Road, Kolkata 700009, India

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

Article history: A new system of saturated polyester and their nanocomposites synthesized from glycolyzed PET with
Received 7 July 2009 varied composition is investigated for the sorption and diffusion studies in water. The kinetics of
Received in revised form sorption is studied by using the equation of transport phenomena. The values of ‘n’ from transport
29 January 2010
equation are found to be below ‘0.5’, showing the non-Fickian or pseudo-Fickian transport in the
Accepted 29 March 2010
polymer. The dependence of diffusion coefficient on composition and temperature has been studied for
Available online 1 April 2010
all polymeric samples. The diffusion coefficient of saturated polyester samples decreases with an
Keywords: increase in glycolyzed PET contents. The nanocomposite samples show less diffusion coefficient than
Water transport pristine polymer and it decreases with an increase in nano-filler up to 4 wt%. The diffusion coefficient
Diffusion
increases with an increase in temperature for all the samples. The sorption coefficient shows a little
Sorption
change with variation in composition as well as temperature for all the samples and it is in a range of 1.
Glycolyzed PET
Pseudo-Fickian transport The activation energy for diffusion and permeation is positive for all the samples. The heat of sorption is
Activation energy also positive for all the samples, indicating Henry type mode of sorption.
Heat of sorption & 2010 Elsevier Ltd. All rights reserved.

1. Introduction improve properties and reduce cost; however, there are limitations
in their application due to phase separation, particle agglomera-
Usual reactants for the saturated polyesters are a glycol and an tion, and heterogeneous distribution in the product (Carrado,
acid or anhydride. It is the family of polyesters in which the 2003). Small size of clay minerals and clay surface treatments
polyester backbones are saturated and hence unreactive as with chemicals in order to create an affinity between the clay
compared to the more reactive, unsaturated ones. Saturated surface and the polymer, can reduce those problems. According
polyesters consist of low molecular weight liquids used as to Tsai (2000) the addition of layered silicates in PET acts as a
plasticizers and as reactants in forming urethane polymers, and heterogeneous nucleating agent, which increases the overall
linear, high molecular weight thermoplastics such as polyethy- crystallisation rate and the crystalline fraction. The fact that clay
lene terephthalate (Mylar and Dacron). Poly(ethylene terephtha- particles are impermeable is also expected to improve barrier
late) (PET) is a semicrystalline polymer, with only minor properties of the PET nanocomposite to gases and water vapour
differences in molecular weight and modifications, PET is used (Wang et al., 2004).
in textiles (clothes, curtains, and furniture upholstery), reinforce- When water in liquid or vapor form is absorbed into the
ment of tires and rubber goods, and food and beverage packaging polymer; water molecules fill the voids that are formed between
(water, soft drink and isotonic beverage bottles, sauce and jam polymer chains and then induce relaxation, or swelling of the
jars, etc.) (Imai et al., 2002). The incorporation of clay into PET can polymer (Adhikari and Majumdar, 2004; Kondratowicz et al.,
result in outstanding property improvement in terms of decreasing 2001; Buchold et al., 1999; Pradas et al., 2001; Turner, 1982).
water permeability in food packaging, increasing flame resistance Water affects polymers in ways that are unique in comparison to
in textiles, and increasing the modulus in injection molded parts. any other substance (Pradas et al., 2001; Nogueira et al., 2000).
This can be done with less clay content than used in most The study of water sorption in polymers is important for many
conventional composites. Conventional fillers have been used to applications. For example, polymers are used in the coatings
industry. Coatings are also important in the food industry (Lange
and Wyser, 2003; Chirife and Buera, 1996; Arvanitoyannis, 1999).
 Corresponding author at: Department of Polymer Science and Technology, Water is probably the single most important factor in governing
University of Calcutta, 92, A.P.C. Road, Kolkata 700009, India.
microbial spoilage in foods. Therefore polymers that are excellent
E-mail address: ppk923@yahoo.com (P.P. Kundu). barriers for water are needed for packaging.

0009-2509/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2010.03.050
ARTICLE IN PRESS
S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387 4379

Sorption and diffusion of the solvents in and through polymers whole reaction was carried out in inert atmosphere (argon
have been widely investigated from both theoretical as well as atmosphere) under reflux with constant stirring.
experimental point of view (Gouanve et al., 2007; Hansen, 2004;
Tanigami et al., 1995; Lobo et al., 1996; Valente et al., 2005; 2.2.3. Synthesis of saturated polyester from glycolyzed PET
Pereira et al., 2009). The swelling technique is a commonly used Initially, a reference resin (blank) was synthesized by direct
method to determine various coefficients such as diffusion, esterification (Samant and Ng, 1999; Yang et al., 1996; Ravin-
sorption, and permeability coefficient (Zhu et al., 2006; Detallante dranath and Mashelkar, 1982a, b; Chegolya et al., 1979; Kemkes,
et al., 2002; Kumar et al., 1985; Han et al., 1995; Singh et al., 1970; Mellichamp, 1970). This reaction was a heterogeneous
2005). In swelling experiments, the polymer of known dimension reaction with monomers of phthalic anhydride and ethylene
is dispersed in a solvent, and the solvent mass uptake versus time glycol. The mixture of monomers was charged as slurry, because
is recorded and the data are used to calculate the various TA was hard to dissolve in EG. The TA:EG molar ratio used is
coefficients. These coefficients give an idea about the use of 1:1.1–2 and the reaction temperature was usually 150–180 1C.
polymers in various applications, such as membranes, ion- The reaction was carried out in a five-necked reaction kettle
exchangers, controlled release systems, packaging, microchip equipped with a mechanical stirrer, inlet to inert gas (argon),
manufacturing, etc. There are very few patents on the application thermometer, and condenser. The reactor was purged with argon
of PET nanocomposites for barrier properties (Barbee et al., 2000; for 15 min, before heating was switched on. Then, the contents
Williamson et al., 2006). were heated under argon atmosphere to 180 1C until approxi-
In the present study, the saturated polyester synthesized from mately 15 mL of water were collected. In the commercial reaction,
glycolyzed PET with varied compositions and their nanocompo- some PET prepolymer (BHET) is added in order to shorten the
site are studied using Fickian model. The activation energy for reaction time. Pre-polymer produced from the direct esterifica-
diffusion (ED) and for permeation (EP) is calculated by using tion reaction was gradually heated to 200 1C. In this step, EG was
standard Arrhenius relationship. The variation in sorption is collected as a byproduct. The overall reaction time, including the
studied with respect to time and temperature. The objective of the esterification and the polycondensation processes, was long and
present work is to study the sorption and diffusion. Kinetics of the usually varies from 5 to 10 h (Ravindranath and Mashelkar,
saturated polyester synthesized from glycolyzed PET waste and 1982a, b; Chegolya et al., 1979).
their nanocomposites with an alteration in the composition and The above experiment was repeated by replacing the ethylene
clay contents. The effect of various parameters on the polymer glycol with glycolyzed PET (GPET) by 100:0, 80:20, 60:40, 50:50, and
samples are studied by using transport equation and Arrhenius 40:60 percent, respectively. The nomenclature used in this work is
relationship. based on the original composition of reactants (shown in Table 1).

2. Experimental 2.2.4. Synthesis of saturated polyester/clay nanocomposites


The saturated polyester nanocomposites (GPET waste) were
2.1. Materials prepared by heating the desired mixture of phthalic anhydride,
ethylene glycol, GPET and modified montmorillonite clay in a five
necked reaction kettle. The modified nano-filler of a predeter-
Discarded PET bottles from soft drinks were procured from
mined quantity was dispersed in a reaction mixture. The detailed
scrapers, cleaned thoroughly and cut into small pieces (6 mm  6
compositions are reported in Table 1. The dispersion was
mm). Zinc acetate, minimum assay 99%, ethylene glycol (EG),
maintained by constant mechanical stirring at 500 rpm (overnight
diethylene glycol (DEG) and styrene were procured from E. Merck
for proper intercalation). The mixture was heated at 120 1C for 3 h,
(India) Pvt. Ltd., Bombay. Phthalic anhydride (TA) was obtained
followed by at 150–200 1C for 3–4 h. The whole mass was
from CDH (India). Montmorillonite (K-10), dodecyl trimethyl
transferred to an appropriate mould. The photographs of the
ammonium bromide (DTAB), cetyl-trimethyl ammonium bromide
nanocomposite prepared are shown in Fig. 1.
(CTAB) were purchased from Aldrich Chemical Company (Mil-
waukee, MI) and used as received.
2.2.5. Swelling experiments
The samples were cut into circular form using a die of 10 mm
2.2. Synthetic work
diameter. The thickness of the sample was measured by means of
a screw gauge. The dry samples were weighed on an electronic
2.2.1. Modification of montmorillonite
balance (Citizen, CX 220) and then kept in the solvent in screwed
Montmorillonite was modified by the same method used in
bottles. The samples were taken out of the solvent at specific
previous study (Sharma et al., 2008). Cetyl trimethyl ammonium
intervals and the excess solvent was rubbed off. The samples were
bromide (CTAB) or dodecyl trimethyl ammonium bromide (DTAB)
then weighed and again immersed in the solvent till equilibrium
was used for the modification of clay. The CEC (cation exchange
was attained (i.e. 72 h). The time for measuring weight of the
capacity) calculated from the titre value for CTAB was 29.92 and
sample was kept minimal (about 30 s), so that the escape of the
for DTAB 152.84 meq/100 gm of clay.
solvent from the sample remains negligible. Equilibrium swelling
experiments were carried out at 20, 30, 40 and 50 1C ( 71 1C) to
2.2.2. Glycolysis of PET waste study the effect of temperature on the swelling. For temperatures
Glycolysis of PET scrap has been done in a 1000 mL five-necked higher than room temperature, the samples were kept in a
reactor equipped with a reflux condenser, an inert gas inlet, a microprocessor controlled hot air oven.
mechanical PTFE blade stirrer and a thermocouple linked to a The mole percent uptake (Qt) at each time interval was
temperature regulation device. Molar ratio of PET repeating calculated by using (Ajithkumar et al., 1998).
unit to glycol has been taken 1:2, respectively. The mixture of
Mt 100
diols DEG:EG was charged in the ratio 50:50, respectively. Qt ¼  ð1Þ
Mr Mi
Zn(CH3COO)2 was used as trans-esterification catalyst. Then, the
reactants were heated in the temperature range from 120 to where Mt is the mass of the solvent taken up at time t, Mr is the
140 1C for first 3 h and then at 180 1C for subsequent 5 h. The relative molecular mass of the solvent and Mi is the mass of the
ARTICLE IN PRESS
4380 S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387

Table 1
Detailed composition of the saturated polyester and their nanocomposite with varied glycolyzed PET content.

Sample ID Glycolyzed PET/(%) Ethylene glycol/(%) Montmorillonite Clay/(%)

STD PET 0 100 0


GPET20 20 80 0
GPET40 40 60 0
GPET50 50 50 0
GPET60 60 40 0
GPET50(2)CTAB 50 50 2(CTAB)
GPET50(3)CTAB 50 50 3(CTAB
GPET50(4)CTAB 50 50 4(CTAB
GPET50(5)CTAB 50 50 5(CTAB
GPET20(4)CTAB 20 80 4(CTAB
GPET40(4)CTAB 40 60 4(CTAB
GPET(60)4CTAB 60 40 4(CTAB
GPET50(4)DTAB 50 50 5(DTAB)

Fig. 2. The sorption curve for saturated polyester with varied GPET composition at
25 1C.

in Fig. 2 for saturated polyester with varied GPET composition at


25 1C. It is observed that the swelling of the saturated polyester
decreases with an increase in the GPET contents. The sample
containing 50% GPET shows a minimum swelling, whereas the
neat sample (without GPET) shows a maximum swelling. In the
present case, the decrease in the water uptake with an increase in
the GPET contents can only be explained by an increase in the
rigid part with an increase in the GPET content, resulting in more
rigid chains. It is well known that incorporation of rigid bulky
groups such as an aromatic ring in the chain makes the polymer
more stiff and rigid. Moreover, the bulky group resulted in a great
hindrance to the chain rotation and thus, reduced the chain
flexibility. Depolymerization of PET with ethylene glycol (EG)
(Baliga and Wong, 1989; Goje and Mishra, 2003; Campanelli et al.,
1994; Kao et al., 1997; Chen and Chen, 1999; Chen et al., 2001)
Fig. 1. The photographs of prepared nanocomposites.
leads to bis hydroxy ethyl terephthalate (BHET) and PET
oligomers. The insertion of depolymerized PET, which is the
mixture of hydroxyl terminated monomer and oligomers, results
dry sample. The mole percent uptake at 72 h is taken as swelling in more rigid and stiff backbone of resulting polymeric chain. It is
at infinite time (QN). observed that beyond 50% GPET content the water uptake starts
increasing. It is possibly due to the decrease in the extent of
reaction with increase in the GPET content beyond 50%.
3. Results and discussion Fig. 3 shows the sorption curves for the nanocomposite
samples at 25 1C. Fig. 3a shows sorption for the saturated
3.1. Swelling of polymer samples polyester samples with varying GPET contents at 4% modified
montmorillonite clay. The maximum and minimum water uptake
The mole percent uptake of the solvent is plotted against is shown by samples having 60% and 50% GPET content at 4%
square root of time. The plot for the mole percent uptake is shown nano-clay (CTAB modified), respectively. The remaining samples
ARTICLE IN PRESS
S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387 4381

show intermediate swelling. Fig. 3b shows the sorption curves for due to a quasi-equilibrium, which first reaches rapidly at the
the samples with varying montmorillonite contents from 0% to 5% polymer surface and then by simple diffusion throughout the
at similar composition. For the samples containing varying nano- polymer sample (Bagley and Long, 1955). The second stage of
clay at 50% GPET composition, the maximum and minimum sorption is associated with an increase in surface concentration,
swelling is observed for the samples with 5% and 4% clay content which occurs slowly compared to the diffusion process and is the
(modified by CTAB), respectively. It is observed that a sample with rate-determining factor for sorption. The concentration is vir-
5% nano-filler shows high solvent uptake. This may be due to tually uniform throughout the sheet and increases at a rate,
the high nano-clay content, resulting in the accumulation of independent of the thickness. From recent works, a better
clay at the interface, hence, reducing the barrier properties of rationalization of these anomalous behaviors has been achieved,
the nanocomposite. The clay (montmorillonite) is modified by where contributions from the effect of macroscopic elastic
two surfactants, cetyl trimethyl ammonium bromide (CTAB) and constraints arising during the swelling process (geometrical
dodecyl trimethyl ammonium bromide (DTAB). The sample effects) in adsorption experiments have been pointed out (Rossi,
modified by CTAB (surfactant) shows relatively less swelling 1996). Initially, the sample is completely dry and immersed in the
than GPET50 (4) modified by DTAB. It is because of inserting long wet environment (penetrant). The diffusion of penetrant starts
chain surfactant (CTAB) into the hydrophilic galleries of the native instantaneously and the sample starts swelling. The dry (non-
clay, the interlayer distance increases, and the surface chemistry swollen) core of the sample exerts a compressive stress on the
of the clay is modified. These newly rendered organophilic outer wet (swollen) layers, hindering the diffusion (Samus and
galleries show less affinity to water. Rossi, 1996). Since the core is still dry, the only direction to which
It is evident from Figs. 2, 3a and 3b, for all polymeric samples a the sample is allowed to swell is perpendicular to the surface
two stage sorption is followed (a non-Fickian diffusion case). It is (stage I). This is proved by the fast increase in thickness of the
sample, compared to the square root increase of the area. As the
diffusion front penetrates the sample, the reducing force is less
prominent and ultimately disappears, when the diffusion front
STDPET reaches middle of the sample. At this point, the swelling of the
0.7 GPET20-4 (CTAB)
GPET40-4 (CTAB) sample is free to commence in all directions (stage II) and the
GPET50-4 (CTAB) diffusion rate increases, until it levels off (close to the dissemina-
0.6 GPET60-4 (CTAB)
GPET50-4 (DTAB) tion point).
Fig. 4 shows the sorption curve for STDPET, sample with 50%
0.5 GPET content and its nanocomposite at different temperatures.
Qt / (mol %)

For STDPET sample the water uptake is a maximum at 55 1C


0.4 and decreases with decrease in temperature. The GPET-50%
sample shows a minimum swelling at 25 1C and a maximum
0.3 swelling at 55 1C. It is observed that the swelling increases with an
increase in temperature. This is due to the fact that the polymer
chain shows higher segmental mobility with an increase in
0.2
temperature, hence accommodating higher amount of the solvent.
The nanocomposite sample GPET-50-4% shows reverse order for
0.1
swelling with an increase in the temperature. It has maximum
swelling at low temperature and minimum swelling at high
0.0
0 1 2 3 4 5 6 7 8 9 10
(Time)1/2 / hr 0.9

0.8

0.7 GPET50
0.7
GPET50-2 (CTAB)
GPET50-3 (CTAB)
0.6 0.6
GPET50-4 (CTAB)
Qt / (mol %)

GPET50-5 (CTAB)
0.5
GPET50-4 (DTAB)
0.5
0.4
Qt / (mol %)

0.4
0.3

0.3 0.2

0.1
0.2
0.0
0.1 0 1 2 3 4 5 6 7 8 9 10
(Time)1/2 / hr
0.0 STDPET@25°C GPET50@25°C GPET50-4-@25°C
0 1 2 3 4 5 6 7 8 9 10 STDPET@35°C GPET50@35°C GPET50-4-@35°C
(Time)1/2 / hr STDPET@45°C GPET50@45°C GPET50-4-@45°C
STDPET@55°C GPET50@55°C GPET50-4-@55°C
Fig. 3. The sorption curves for nanocomposite samples at 25 1C. (a) For samples
with variation in GPET composition, and (b) for samples with variation in clay Fig. 4. The sorption curve for saturated polyester (STDPET, GPET50), and
contents. nanocomposite sample (GPET50-4) at different temperatures.
ARTICLE IN PRESS
4382 S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387

Solvent Table 2
Values of mole percent uptake at infinite time (QN), n, k and standard deviation
(SD) for different samples at 25 and 35 1C.

Temperature/(1C) Sample ID QN/(mol %) n K SD

25 STD PET 0.667 0.168  0.721 0.029


GPET20 0.659 0.173  0.752 0.036
GPET40 0.648 0.175  0.767 0.032
GPET50 0.505 0.141  0.599 0.014
GPET60 0.587 0.140  0.595 0.023
L
W GPET50(2)CTAB 0.497 0.158  0.677 0.016
d GPET50(3)CTAB 0.460 0.151  0.635 0.022
GPET50(4)CTAB 0.454 0.149  0.666 0.030
GPET50(5)CTAB 0.505 0.098  0.423 0.015
GPET20(4)CTAB 0.645 0.180  0.773 0.026
GPET40(4)CTAB 0.562 0.139  0.575 0.018
GPET(60)4CTAB 0.676 0.093  0.386 0.008
GPET50(4)DTAB 0.460 0.094  0.394 0.014

35 STD PET 0.723 0.177  0.791 0.033


GPET20 0.689 0.178  0.789 0.042
GPET40 0.675 0.182  0.806 0.032
GPET50 0.525 0.141  0.599 0.013
Scheme 1. The schematic representation of tortuosity-based model to describe GPET60 0.623 0.158  0.667 0.028
the solvent diffusion in the nano-clay filled polymer composite. (W is the width or GPET50(2)CTAB 0.465 0.177  0.749 0.020
thickness and L is the length of the filler platelet. d is the thickness of the polymer GPET50(3)CTAB 0.445 0.153  0.648 0.022
matrix, through which the solvent molecules diffuse). GPET50(4)CTAB 0.423 0.160  0.716 0.036
GPET50(5)CTAB 0.487 0.098  0.421 0.013
GPET20(4)CTAB 0.654 0.176  0.752 0.026
temperature. It is due to fact that the layered nano-fillers have
GPET40(4)CTAB 0.534 0.135  0.563 0.022
platelet like structure, which improves the barrier properties of GPET(60)4CTAB 0.657 0.123  0.530 0.012
the polymer (Lu and Mai, 2005). The platelets due to the rise in GPET50(4)DTAB 0.447 0.102  0.439 0.010
temperature are then evenly distributed into the polymer matrix,
creating multiple parallel layers. These layers force the solvent
molecules to flow through the polymer in a ‘‘torturous path’’, Table 3
forming complex barriers to the solvent molecules. The detailed Values of mole percent uptake at infinite time (QN), n, k and standard deviation
process is shown in Scheme 1. The filler platelets are (SD) for different samples at 45 and 55 1C.
impenetrable for the diffusing solvent molecules. Therefore,
Temperature/(1C) Sample ID QN/(mol %) n K SD
when compared to the parent polymer, a decrease in the
diffusion of the solvent in the nanocomposites is observed. 45 STD PET 0.754 0.176  0.808 0.041
In order to find out the mechanism of swelling, the diffusion GPET20 0.695 0.172  0.752 0.037
data are fitted into an empirical equation (Eq. (3)), George et al. GPET40 0.687 0.180  0.800 0.034
GPET50 0.556 0.133  0.572 0.011
(1999) derived from the equation of transport phenomena
GPET60 0.654 0.158  0.666 0.029
(Ajithkumar et al., 1998) GPET50(2)CTAB 0.447 0.178  0.754 0.022
GPET50(3)CTAB 0.427 0.148  0.622 0.025
Qt
¼ ktn ð2Þ GPET50(4)CTAB 0.411 0.141  0.633 0.031
Q1 GPET50(5)CTAB 0.456 0.098  0.419 0.012
GPET20(4)CTAB 0.634 0.182  0.783 0.029
GPET40(4)CTAB 0.511 0.126  0.517 0.021
Qt
ln ¼ ln k þ n ln t ð3Þ GPET(60)4CTAB 0.648 0.117  0.497 0.016
Q1 GPET50(4)DTAB 0.412 0.104  0.432 0.010

where Qt and QN are the mole percent uptake of solvent at time ‘t’ 55 STD PET 0.767 0.176  0.819 0.045
and at infinity or equilibrium. ‘k’ is a constant, which depends GPET20 0.721 0.179  0.805 0.041
GPET40 0.697 0.182  0.799 0.034
upon the solvent-polymer interaction and the structure of the
GPET50 0.565 0.141  0.619 0.016
polymer. For all the samples, the regression coefficient (r) varies GPET60 0.667 0.157  0.651 0.031
between 0.95 and 0.99. The values of constant ‘k’ and ‘n’ obtained GPET50(2)CTAB 0.423 0.185  0.782 0.025
from the Eq. (3) and QN are reported in Tables 2 and 3. The value GPET50(3)CTAB 0.401 0.150  0.628 0.025
of ‘n’ gives an idea of the mechanism of sorption (Crank, 1975). GPET50(4)CTAB 0.387 0.151  0.681 0.037
GPET50(5)CTAB 0.432 0.097  0.410 0.012
For the value of ‘n’ as 0.5, the mechanism of swelling is termed as GPET20(4)CTAB 0.612 0.191  0.835 0.040
Fickian transport. This occurs, when the rate of diffusion of GPET40(4)CTAB 0.498 0.138  0.574 0.015
solvent is smaller than polymer segmental mobility. If the value of GPET(60)4CTAB 0.667 0.121  0.541 0.021
‘n’ is not equal to ‘0.5’, then the transport is considered as non- GPET50(4)DTAB 0.398 0.093  0.407 0.012
Fickian. In particular, if ‘n ¼1’, the transport is called ‘case II’
transport (Crank, 1975). It is a special case, where the solvent
front moves with constant velocity. If ‘n’ lies between ‘0.5’ and ‘1’, From Tables 2 and 3, the values of ‘n’ are below ‘0.5’, indicating
then it is called anomalous transport (Crank, 1975). For non- the transport as non-Fickian or pseudo-Fickian transport.
Fickian transport, diffusion is more rapid than the polymer In glassy polymers, deviations from this ideal Fickian behavior
relaxation rate. For anomalous transport, the diffusion and are often observed. These deviations are generally believed to
relaxation rates are comparable (Crank, 1975). If sorption is less arise as a consequence of the finite rate of polymer structure
than 0.5, then it is termed as pseudo-Fickian transport. It is reorganization in response to penetrant-induced swelling during
characterized by initial curvature of the Qt versus t1/2 plots out of the sorption-diffusion process (Rogers, 1985). The penetrant may
the origin concave to the time axis (Rogers, 1985; Windle, 1985). sorb in the polymer in two stages, an initial Fickian-like stage
ARTICLE IN PRESS
S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387 4383

followed by a protracted drift towards a final equilibrium value motion of the polymer segments closes the connection between
(McDowell et al., 1999). In another departure from Fickian the two free volume elements and if the penetrant happens to be
sorption kinetics, the penetrant weight uptake may be a linear away from its original position, as shown in 2C, when the gap in
function of contact time until equilibrium is reached (Moaddeb the polymer matrix is closed, the penetrant will be trapped in
and Koros, 1995). When penetrant sorption is accompanied another free volume element in the polymer matrix and will have
by significant swelling of the polymer, any time dependent executed a diffusion step. The process shown in Scheme 2 has
resistance to changes in the volume of the polymer can lead to been called the ‘‘Red Sea mechanism of penetrant transport in
non-Fickian sorption kinetics. As the penetrant swells the dense polymers’’ (Muller-Plathe, 1994).
polymer, local stresses are built up when the chains disentangle The swelling data are used to calculate the diffusion coefficient
from each other. These stresses can be quite high, and can, infact, (D), which is a measure of the ability of the solvent molecules to
cause mechanical failure in the polymer. move through the polymer and the sorption coefficient (S), which
The rate limiting step for penetrant diffusion is the creation gives an idea about the equilibrium sorption. The diffusion
of transient ‘‘gaps’’ in the polymer matrix via local scale coefficient (D) is calculated as (Ajithkumar et al., 2000)
polymer segmental dynamics involving several polymer chains  2
(Muller-Plathe, 1994). Penetrant molecules vibrate inside local hy
D¼p ð4Þ
cavities in the polymer matrix at frequencies much higher than 4Q1
the frequency of polymer chain motion required to open a gap of where p ¼3.14; h is the thickness of the dry sample and y is the
sufficient size to accommodate the penetrant. These steps are slope of the initial linear portion of the curve Qt versus Ot; and QN
shown schematically in Scheme 2. In 2A, a penetrant molecule is is the mole percent uptake of the solvent at infinite time.
shown dissolved in a polymer matrix. The penetrant vibrates The sorption coefficient (S) is calculated as (Ajithkumar et al.,
inside a gap or molecular scale cavity in the polymer matrix 2000)
at very high frequency (ca. 1012 vibrations/s or 1 vibration/
M1
picosecond) (Muller-Plathe, 1994). The polymer molecules do not S¼ ð5Þ
occupy the entire volume of the polymer sample. Due to packing Mp
inefficiencies and polymer chain molecular motion, some of the where MN is the mass of the solvent uptake at equilibrium and Mp
volume in the polymer matrix is empty or ‘‘free’’ and this so- is the mass of the dry sample. The sorption and diffusion
called free volume is redistributed continuously as a result of the coefficients are used to calculate permeability coefficient (P) of
random, thermally stimulated molecular motion of the polymer the samples, which is given by (Ajithkumar et al., 2000)
segments (Ghosal and Freeman, 1994).
P¼DS ð6Þ
In 2B, local polymer segmental motion has opened a connect-
ing channel between two free volume elements in the polymer The values of these coefficients are reported in Tables 4 and 5.
matrix and the penetrant molecule can, as a result of its own It is observed that for saturated polyester samples synthesized by
Brownian motion, explore the entire corridor between the initial varied GPET content, STDPET sample shows the highest and the
free volume element which it occupied and the second free GPET50 sample shows the lowest diffusion coefficient. The
volume element which is connected to it via the opening of a diffusion coefficient of the vary GPET samples decreases with an
transient gap in the polymer matrix. Eventually, local segmental increase in the GPET content from 20% to 50% (Tables 4 and 5). It
is due to an increase in the rigid part with an increase in the GPET
content (a mixture of monomer and oligomer), resulting in more
rigid chains thereby, decreasing the diffusion of solvent in
the polymer network. The diffusion coefficient increases with
an increase in the temperature for all varied GPET samples.
The decrease in the diffusion coefficient is due to the insertion of
phthalic anhydride moiety in the polymer chain. An increase in
the GPET content, results in incorporation of rigid bulky groups,
such as an aromatic ring in the chain makes the polymer stiffer
and rigid, hence shows less affinity to water. The sorption
coefficient values decrease with an increase in the GPET
content. Likewise, diffusion coefficient, the sorption coefficient
values increase with an increase in the temperature. The increase
in sorption is quite small and almost equal to one for all samples
at all temperatures. This indicates that there is very less change in
the sorption properties with variation in GPET contents as well as
temperature. The permeability in the samples shows the same
trend as that for diffusion coefficient.
For the nanocomposite samples with varied GPET content at
fixed organically modified nano-clay (4%), the diffusion coefficient
decreases with an increase in GPET content (Tables 4 and 5). It is
observed that a maximum and a minimum diffusion coefficient
values are shown by the samples GPET60(4) (CTAB modify), and
GPET50(4), respectively. The sorption coefficient decreases with an
increase in the GPET content up to 50% and increases with an
increase in the temperature. The permeability in the samples is
observed to show a similar trend as that for the diffusion coefficient.
For the nanocomposite samples with varied nano-clay from 0%
Scheme 2. The schematic representation of the penetration of solvent molecules to 5% at fixed GPET content, the diffusion coefficient decreases
in the polymer matrix. with an increase in the nano-clay content up to 4%. GPET50(5)
ARTICLE IN PRESS
4384 S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387

Table 4 -5.30
Diffusion coefficient (D), sorption coefficient (S) and permeability coefficient (P) in STDPET
GPET20
water for different samples at 25 and 35 1C. GPET40
GPET50
-5.35 GPET60
Temperature/(1C) Sample ID D/(10  7 cm2/s) S (g/g) P/(10  7 cm2/s)

25 STD PET 33.50 0.120 4.02


-5.40

Log D
GPET20 33.10 0.118 3.90
GPET40 31.32 0.116 3.63
GPET50 31.13 0.091 2.83
GPET60 32.46 0.105 3.40
-5.45
GPET50(2)CTAB 29.48 0.089 2.62
GPET50(3)CTAB 29.16 0.082 2.39
GPET50(4)CTAB 27.63 0.081 2.23
GPET50(5)CTAB 33.02 0.090 2.97 -5.50
GPET20(4)CTAB 31.39 0.116 3.64
GPET40(4)CTAB 30.77 0.101 3.10
GPET(60)4CTAB 32.01 0.121 3.87 0.0030 0.0031 0.0032 0.0033 0.0034
GPET50(4)DTAB 30.32 0.082 2.48 1/T / (K-1)
35 STD PET 33.89 0.130 4.40
GPET20 33.56 0.124 4.16 STDPET
GPET20
GPET40 31.81 0.121 3.84 -6.25 GPET40
GPET50 34.76 0.094 3.26 GPET50
GPET60
GPET60 47.66 0.112 5.33 -6.30
GPET50(2)CTAB 30.00 0.083 2.49
GPET50(3)CTAB 29.87 0.080 2.38
-6.35
GPET50(4)CTAB 29.65 0.076 2.25

Log P
GPET50(5)CTAB 33.54 0.087 2.91
GPET20(4)CTAB 31.90 0.117 3.73 -6.40
GPET40(4)CTAB 32.96 0.096 3.16
GPET(60)4CTAB 33.55 0.118 3.95 -6.45
GPET50(4)DTAB 33.21 0.080 2.65
-6.50

Table 5
-6.55
Diffusion coefficient (D), sorption coefficient (S) and permeability coefficient (P) in 0.0030 0.0031 0.0032 0.0033 0.0034
water for different samples at 45 and 55 1C.
1/T / (K-1)
Temperature/(1C) Sample ID D/(10  7 cm2/s) S (g/g) P/(10  7 cm2/s)
Fig. 5. The plot of log D and log P versus inverse of temperature (1/T) for saturated
polyester with varied GPET composition.
45 STD PET 34.32 0.135 4.63
GPET20 34.42 0.125 4.30
GPET40 31.98 0.123 3.93
GPET50 35.89 0.100 3.58
GPET60 47.95 0.117 5.61 polymer with 5% filler shows high diffusion coefficient value even
GPET50(2)CTAB 30.43 0.080 2.43 more than that for the sample without clay (GPET50). This may be
GPET50(3)CTAB 31.38 0.076 2.38
due to the high nano-clay contents, resulting in the accumulation
GPET50(4)CTAB 30.61 0.073 2.23
GPET50(5)CTAB 38.44 0.082 3.15
of clay at the interface. This behavior is just like ordinary filler.
GPET20(4)CTAB 33.19 0.114 3.78 The higher filler contents (5%) reduce the barrier properties of the
GPET40(4)CTAB 34.08 0.091 3.10 nanocomposite. Therefore, the sample shows distortion, instead
GPET(60)4CTAB 34.73 0.116 4.02 of swelling after 12 h.
GPET50(4)DTAB 34.57 0.074 2.55
An increase in the temperature results in an increase in the
55 STD PET 35.77 0.138 4.93 diffusion coefficient for all the samples. The sorption coefficient
GPET20 34.51 0.129 4.45
for all the samples shows a decrease with an increase in GPET
GPET40 32.57 0.125 4.07
GPET50 37.76 0.101 3.81 content up to 50% and 4% nano-clay content. This indicates that
GPET60 48.47 0.120 5.81 the solvent assimilation or the ability to absorb solvent decreases
GPET50(2)CTAB 31.64 0.076 2.40 due to the incorporation of nano-filler. This indicates the good
GPET50(3)CTAB 33.34 0.072 2.40 compatibility of the nano-clay and polymer matrix. The perme-
GPET50(4)CTAB 32.21 0.069 2.22
ability coefficient shows similar trends as is for diffusion
GPET50(5)CTAB 38.85 0.077 2.99
GPET20(4)CTAB 33.73 0.110 3.71 coefficient.
GPET40(4)CTAB 33.52 0.089 2.98 Diffusion and permeation are thermally activated processes,
GPET(60)4CTAB 31.09 0.120 3.73 and their temperature dependence can be used to calculate the
GPET50(4)DTAB 35.12 0.071 2.49
activation energy for the process of water absorption. The
activation energy for diffusion (ED) and for permeation (EP) is
calculated by using standard Arrhenius relationship (Eq. (8))
(CTAB modify) shows a maximum and GPET50(4) shows a (George and Thomas, 2001; George et al., 1996).
minimum diffusion coefficient. With an increase in the nano-
X ¼ Xo ðe-EX=RT Þ ð7Þ
clay, the decreasing values of the diffusion coefficient are due to
the incorporation of the nano-clay that increases the barrier
log X ¼ log Xo -EX =2:303:RT ð8Þ
properties. The platelet like structure with a high aspect ratio can
be expected to improve the resistance towards low molecular where X represents either D or P, Xo is a constant representing
weight solvent molecules. It is expected that a high loading of either Do or Po. EX is either ED or EP, which depends upon the
nano-clay is more effective for the solvent resistance, but the swelling process under consideration. From the linear plot of log X
ARTICLE IN PRESS
S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387 4385

-5.40
GPET50
-5.42 GPET50-2 (CTAB)
GPET50-3 (CTAB)
GPET50-4 (CTAB)
-5.44 GPET50-5 (CTAB)
GPET50-4 (DTAB)
-5.46

Log D
-5.48
-5.50
-5.52
-5.54
-5.56

0.0030 0.0031 0.0032 0.0033 0.0034


1/T / (K-1)

STDPET
-5.6 GPET20-4 (CTAB) GPET50
GPET40-4 (CTAB) -5.6 GPET50-2 (CTAB)
GPET50-4 (CTAB) GPET50-3 (CTAB)
-5.8 GPET60-4 (CTAB) GPET50-4 (CTAB)
GPET50-4 (DTAB) -5.8 GPET50-5 (CTAB)
GPET50-4 (DTAB)
-6.0 -6.0
Log P
Log P

-6.2 -6.2

-6.4 -6.4

-6.6 -6.6

-6.8
-6.8
0.0030 0.0031 0.0032 0.0033 0.0034
0.0030 0.0031 0.0032 0.0033 0.0034
1/T / (K-1)
1/T / (K-1)
Fig. 7. The plot of log D and log P versus inverse of temperature (1/T) for saturated
Fig. 6. The plot of log D and log P versus inverse of temperature (1/T) for saturated polyester with varied nano-clay content at fixed composition.
polyester with varied GPET content at 4% nano-clay.

Table 6
Arrhenius parameters and heat of sorption for different polymers in water.
versus 1/T, the values of EX can be calculated. Figs. 5–7 give the
plot of log D and log P versus 1/T for different polymer samples. Sample ID ED/(102 kJ/mol) EP/(102 kJ/mol) DHS/(102 kJ/mol)
The values of ED and EP calculated from Figs. 5 and 6 are shown
STD PET 5.04 16.09 11.05
in Table 6. The regression coefficient values in estimation GPET20 5.07 10.42 5.35
are between 0.92 and 0.99. The heat of sorption (DHS) is also GPET40 8.98 8.90 5.92
calculated by using the Eq. (9) (George and Thomas, 2001; George GPET50 15.01 23.98 8.97
et al., 1996) GPET60 22.78 24.79 2.01
GPET50(2)CTAB 5.44 10.65 5.21
DHS ¼ EP -ED ð9Þ GPET50(3)CTAB 10.86 15.49 4.63
GPET50(4)CTAB 11.94 79.98 68.04
The values of DHS provide additional information about the GPET50(5)CTAB 15.11 17.69 2.58
transport of solvent molecules through the polymer matrix. The GPET20(4)CTAB 10.77 48.74 37.97
values of DHS are given in Table 6. All the values are positive, GPET40(4)CTAB 8.96 36.35 27.39
GPET(60)4CTAB 8.016 37.74 29.724
indicating Henry type sorption in the majority (endothermic GPET50(4)DTAB 11.76 12.24 0.48
contribution to the sorption process) (Aminabhavi et al., 1996).
This type of sorption is concerned with the formation of a hole of a
molecular size in the polymer matrix (Aminabhavi et al., 1996).
The activation energy for diffusion is the energy needed to enable 4. Conclusion
the dissolved molecules to jump into another hole. The positive
activation energy of diffusion for all the polymer samples A new system of saturated polyester and their nanocomposites
indicates a high energy requirement to jump from one hole to synthesized from glycolyzed PET with varied composition is
another. It is observed that for polyester samples with an increase investigated for the sorption and diffusion studies in water. The
in GPET content, the activation energy goes on increasing. STDPET kinetics of sorption is studied by using the equation of transport
sample has the lowest and GPET60 sample has the highest values phenomena. The values of ‘n’ in solvent transport equation are
of ED. For nanocomposite samples at fixed GPET (50%) content, ED found to be below ‘0.5’, showing the non-Fickian or pseudo-
value increases with an increase in the clay loading. Fickian transport in the polymer. The dependence of diffusion
ARTICLE IN PRESS
4386 S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387

coefficient on composition and temperature has been studied for Chirife, J., Buera, M.D., 1996. Water activity, water glass dynamics, and the control
all polymeric samples. The diffusion coefficient in saturated of microbiological growth in foods. Critical Reviews in Food Science and
Nutrition 36, 465–513.
polyester samples decreases with an increase up to 50% Crank, J., 1975. The Mathematics of Diffusion, second ed. Clarendon Press, Oxford
glycolyzed PET content. The nanocomposite samples show less (Chapter 11).
diffusion coefficient than pristine polymer and it decreases with Detallante, V., Langevin, D., Chappey, C., Metayer, M., Mercier, R., Pineri, M., 2002.
Kinetics of water vapor sorption in sulfonated polyimide membranes.
an increasing in nano-filler up to 4% by weight. The diffusion
Desalination 148, 333–339.
coefficient increases with an increase in temperature for all George, S.C., Knorgen, M., Thomas, S., 1999. Effect of nature and extent of
polymer samples. The sorption coefficient shows a little change crosslinking on swelling and mechanical behavior of styrene-butadiene rubber
with variation in composition as well as temperature for all the membranes. Journal of Membrane Science 163, 1–17.
George, S.C., Thomas, S., 2001. Transport phenomena through polymeric systems.
samples and it is in a range of 1. STDPET sample has the lowest Progress in Polymer Science 26, 985–1017.
and GPET60 sample has the highest values of ED. For nanocompo- George, S.C., Thomas, S., Ninan, K.N., 1996. Molecular transport of aromatic
site samples at fixed GPET (50%) content, ED value increases with hydrocarbons through crosslinked styrene-butadiene rubber membranes.
Polymer 37, 5839–5848.
an increase in the clay loading. The activation energy for diffusion
Ghosal, K., Freeman, B.D., 1994. Gas separation using polymeric membranes: an
and permeation is positive for all the samples. The heat of overview. Polymers for Advanced Technologies 5, 673–697.
sorption is also positive for all the samples, indicating Henry type Goje, A.S., Mishra, S., 2003. Chemical kinetics, simulation and thermodynamics of
mode of sorption. glycolytic depolymerization of poly (ethylene terephthalate) waste with
catalyst optimization for recycling of value added monomeric products.
From the diffusion studies it can be concluded that these Macromolecular Materials and Engineering 288, 326–336.
polymer nanocomposites can find numerous applications in the Gouanve, F., Marais, S., Bessadok, A., Langevin, D., Metayer, M., 2007. Kinetics of
food processing industries, the barrier to the vapors is a water sorption in flax and PET fibers. European Polymer Journal 43,
586–598.
prerequisite. These nanocomposites can be used in rigid food Han, H., Gryte, C.C., Ree, M., 1995. Water diffusion and sorption in films of high-
and beverage containers due to their barrier properties, proces- performance poly(4,40 -oxydiphenylene pyromellitimide): effects of humidity,
sibility and formability, ecological and toxicological character- imidization history and film thickness. Polymer 36, 1663–1672.
Hansen, C.M., 2004. Aspects of solubility, surfaces, and diffusion in polymers.
istics, and economics.
Progress in Organic Coatings 51, 55–66.
Imai, Y., Nishimura, S., Abe, E., Tateyama, H., Abiko, A., Yamaguchi, A., Aoyama, T.,
Taguchi, H., 2002. High-modulus poly(ethylene terephthalate)/expandable
Appendix A. Supplementary material fluorine mica nanocomposites with a novel reactive compatibilizer. Chemistry
of Materials 14, 477–479.
Kao, C.Y., Cheng, W.H., Wan, B.Z., 1997. Investigation of catalytic glycolysis of
Supplementary data associated with this article can be found polyethylene terephthalate by differential scanning calorimetry. Thermochim
in the online version at doi:10.1016/j.ces.2010.03.050. Acta 292, 95–104.
Kemkes, J.F., 1970. Poly(ethylene terephthalate). US 3497473.
Kondratowicz, B., Narayanawamy, R., Persaud, K.C., 2001. An investigation into the
use of electrochromic polymers in optical fibre gas sensors. Sensors and
Actuators B 74, 138–144.
References Kumar, R., Srivastava, S.K., Mathur, G.N., 1985. Evaluation of swelling studies and
thermal characteristics of castor oil based polyurethane elastomeric networks.
Journal of Elastomers and Plastics 17, 89–96.
Adhikari, B., Majumdar, S., 2004. Polymers in sensor applications. Progress in Lange, J., Wyser, Y., 2003. Recent innovations in barrier technologies for plastic
Polymer Science 29, 699–766. packaging—a review. Packaging Technology and Science 16, 149–158.
Ajithkumar, S., Patel, N.K., Kansara, S.S., 1998. Sorption behaviour of interpene-
Lobo, V.M.M., Murtinho, D.B., Gil, M.H., Garcia, F.P., Valente, A.J.M., 1996. Methods
trating polymer networks based on polyurethane and unsaturated polyester.
of measuring diffusion coefficients of water and potassium chloride in aqueous
Polymer Gels and Network 6, 137–147.
solution in cellulose acetate membranes. International Journal of Polymeric
Ajithkumar, S., Patel, N.K., Kansara, S.S., 2000. Sorption and diffusion of organic
Materials 32, 221–233.
solvents through interpenetrating polymer networks (IPNs) based on
Lu, C., Mai, Y.W., 2005. Influence of aspect ratio on barrier properties of polymer-
polyurethane and unsaturated polyester. European Polymer Journal 36,
clay nanocomposites. Physical Review Letters 95 (1–4), 088303.
2387–2393.
McDowell, C.C., Freeman, B.D., McNeely, G.W., 1999. Acetone sorption and uptake
Aminabhavi, T.M., Phayde, H.T.S., Ortego, J.D., Vergnaud, J.M., 1996. Sorption/
kinetics in poly(ethylene terephthalate). Polymer 40, 3487–3499.
diffusion of aliphatic esters into tetrafluoroethylene/propylene copolymeric
Mellichamp Jr. D.A., 1970. Glycol terephthalate linear polyester by direct
membranes in the temperature interval from 25 to 70 1C. European Polymer
esterification of terephthalic acid. US 3496146.
Journal 32, 1117–1126.
Moaddeb, M., Koros, W.J., 1995. Effects of orientation on the transport of
Arvanitoyannis, I.S., 1999. Totally and partially biodegradable polymer blends
d-limonene in polypropylene. Journal of Applied Polymer Science 57, 687–703.
based on natural and synthetic macromolecules: preparation, physical
Muller-Plathe, F., 1994. Permeation of polymers—a computational approach. Acta
properties, and potential as food packaging materials. Polymer Reviews C 39,
205–271. Polymerica 45, 259–293.
Bagley, E., Long, F.A., 1955. Two-stage sorption and desorption of organic vapors in Nogueira, P., Ramirez, C., Torres, A., Abad, M.J., Cano, J., Lopez, J., Lopez- Bueno, I.,
cellulose acetate. Journal of American Chemical Society 77, 2172–2178. Barral, L., 2000. Effect of water sorption on the structure and mechanical
Baliga, S., Wong, W.T., 1989. Depolymerization of poly(ethylene terephthalate) properties of an epoxy resin system. Journal of Applied Polymer Science 80,
recycled from post-consumer soft drink bottles. Journal of Polymer Science: 71–80.
Part A Polymer Chemistry 27, 2071–2082. Pereira, R.F.P., Cerqueira, D.A., Valente, A.J.M., Polishchuk, A.Y., Burrows, H.D., Lobo,
Barbee, R.B., Matayabas, C., Trexler, J.W., Piner, R.L., 2000. Polyester nanocompo- V.M.M., 2009. Effects of pH and temperature on the sorption of sodium
sites for high barrier applications. USP 6034163. dodecyl sulfate by cellulose acetate/polyaniline blend membranes. Journal of
Buchold, R., Nakladal, A., Gerlach, G., Herold, M., Gauglitz, G., Sahre, K., Eichhorn, Applied Polymer Science 111, 1947–1953.
K.J., 1999. Swelling behavior of thin anisotropic polymer layers. Thin Solid Pradas, M.M., Ribelles, J.L.G., Aroca, A.S., Ferrer, G.G., Anton, J.S., Pissis, P., 2001.
Films 350, 178–185. Interaction between water and polymer chains in poly(hydroxyethyl acrylate)
Campanelli, J.R., Kamal, M.R., Cooper, D.G., 1994. Kinetics of glycolysis of hydrogels. Colloid and Polymer Science 279, 323–330.
poly(ethylene terephthalate) melts. Journal of Applied Polymer Science 54, Ravindranath, K., Mashelkar, R.A., 1982a. Modeling of poly(ethylene terephthalate)
1731–1740. reactors: 4. A continuous esterification process. Polymer Engineering and
Carrado, K.A., 2003. Chapter 10: Polymer-Clay Nanocomposites. In: Shonaike, G.A., Science 22, 610–618.
Advani, S.G. (Eds.), Advanced Polymeric Materials: Structure Property Ravindranath, K., Mashelkar, R.A., 1982b. Modeling of poly(ethylene terephthalate)
Relationships. CRC Press, Boca Raton, FL. reactors: 6. A continuous process for final stages of polycondensation. Polymer
Chegolya, A.S., Shevchenko, V.V., Mikhailov, G.D., 1979. The formation of Engineering and Science 22, 628–636.
polyethylene terephthalate in the presence of dicarboxylic acids. Journal of Rogers, C.E., 1985. Polymer Permeability. In: Comyn, J. (Ed.), first ed., Chapman
Polymer Science: Part A Polymer Chemistry 17, 889–904. Hall, UK, pp. 11–74.
Chen, C.H., Chen, C.Y., Lo, Y.W., Mao, C.F., Liao, W.T., 2001. Studies of glycolysis of Rossi, G., 1996. Macroscopic description of solvent diffusion in polymeric
poly(ethylene terephthalate) recycled from postconsumer soft-drink bottles. 1. materials. Trends in Polymer Science 4, 337–342.
Influences of glycolysis conditions. Journal of Applied Polymer Science 80, Sharma, V., Banait, J.S., Kundu, P.P., 2008. Spectroscopic characterization of linseed
943–948. oil based polymer nanocomposites. Polymer Testing 27, 916–923.
Chen, J.W., Chen, L.W., 1999. The glycolysis of poly(ethylene terephthalate). Samant, K.D., Ng, K.M., 1999. Synthesis of prepolymerization stage in poly-
Journal of Applied Polymer Science 73, 35–40. condensation process. A.I.Ch.E. Journal 45, 1808–1829.
ARTICLE IN PRESS
S. Katoch et al. / Chemical Engineering Science 65 (2010) 4378–4387 4387

Samus, M.A., Rossi, G., 1996. Methanol absorption in ethylene-vinyl alcohol Valente, A.J.M., Polishchuk, A.Y., Burrows, H.D., Lobo, V.M.M., 2005. Permeation of
copolymers: relation between solvent diffusion and changes in glass water as a tool for characterizing the effect of solvent, film thickness and
transition temperature in glassy polymeric materials. Macromolecules 29, water solubility in cellulose acetate membranes. European Polymer Journal 41,
2275–2288. 275–281.
Singh, P., Kaushik, A., Gupta, P., 2005. Characterization of castor oil and Wang, T., Chen, L., Chua, Y.C., Lu, X., 2004. Crystalline morphology and
diphenyl methane diisocyanate-based polyurethane-polystyrene interpene- isothermalcrystallization kinetics of poly (ethylene terephthalate)/clay nano-
trating networks. Journal of Reinforced Plastics and Composites 24, composites. Journal of Applied Polymer Science 94, 1381–1388.
1479–1491. Williamson, D.T., Schleinitz, H.M., Hayes, R.A., 2006. Polyester clay nanocompo-
Tanigami, T., Yano, K., Yamaura, K., Matsuzawa, S., 1995. Anomalous swelling of sites for barrier applications. USP 0141183.
poly(vinyl alcohol) film in mixed solvents of dimethylsulfoxide and water. Windle, A.H., 1985. Polymer Permeability. In: Comyn, J. (Ed.), first ed., Chapman
Polymer 36, 2941–2946. Hall, UK, pp.75–118.
Tsai, T.Y., 2000. Polyethylene terephthalate-clay nanocomposites. In: Pinnavaia, Yang, K.S., An, K.H., Choi, C.N., Jin, S.R., Kim, C.Y., 1996. Solubility and esterification
T.J., Beall, G.W. (Eds.), Polymer-Clay Nanocomposites. John Wiley and Sons, kinetics of terephthalic acid in ethylene glycol III. The effects of functional
New York, pp. 173–192. groups. Journal of Applied Polymer Science 60, 1033–1039.
Turner, D.T., 1982. Polymethyl methacrylate plus water: sorption, kinetics and Zhu, Q., Shentu, B., Liu, Q., Weng, Z., 2006. Swelling behavior of polyethylenimine–
volumetric changes. Polymer 23, 197–202. cobalt complex in water. European Polymer Journal 42, 1417–1422.

Vous aimerez peut-être aussi