Sie sind auf Seite 1von 9

POLYMERS FOR ADVANCED TECHNOLOGIES

Polym. Adv. Technol. 13, 513521 (2002)


Published online in Wiley InterScience (www.interscience.wiley.com). DOI:10.1002/pat.219

Synthesis of Poly[(2-oxo-1,3-dioxolane4-yl)methyl methacrylate-co-styrene] by


Addition Reaction of Carbon Dioxide
and its Compatibility with Poly(methyl
methacrylate) or Poly(vinyl chloride)
Sung-Young Park1, Hee-Young Park1, Hang-Soo Woo1, Chang-Sik Ha2 and
Dae-Won Park1*
1
2

Department of Chemical Engineering/RIIT, Pusan National University, Pusan 609-735, Korea


Department of Polymer Science & Engineering, Pusan National University, Pusan 609-735, Korea

ABSTRACT W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W
We investigated the chemical fixation of carbon dioxide
(CO2) to a copolymer bearing epoxide and the application
of the cyclic carbonate group containing copolymer to
polymer blends. In the synthesis of poly[(2-oxo-1,3dioxolane-4-yl)methyl methacrylate-co-styrene] [poly(DOMA-co-St)] from the addition of CO2 to poly(glycidyl methacrylate-co-styrene) [poly(GMA-co-St)], quaternary ammonium salts showed good catalytic activity
at mild reaction conditions. The CO2 addition reaction
followed pseudo first-order kinetics with the concentration of poly(GMA-co-St). In order to expand the
applications of the CO2 fixed copolymer, polymer blends
of this copolymer with poly(methyl methacrylate)
(PMMA) or poly(vinyl chloride) (PVC) were cast from
N,N'-dimethylformamide (DMF) solution. Miscibility of
blends of poly(DOMA-co-St) with PMMA or PVC have
been investigated both by differential scanning calorimetry (DSC) and visual inspection of the blends, and the
*Correspondence to: Dae-Won Park, Department of Chemical
Engineering/RIIT, Pusan National University, Pusan 609-735,
Korea.
E-mail: dwpark2@pusan.ac.kr

Copyright 2002 John Wiley & Sons, Ltd.

blends were miscible over the whole composition ranges.


The miscibility behaviors were also discussed in terms of
FT-IR spectra. Copyright 2002 John Wiley & Sons,
Ltd.
KEYWORDS: polymer blends; carbon dioxide; poly(glycidyl methacrylate-co-styrene) [poly (GMA-co-St)];
quaternary ammonium salt; miscibility

INTRODUCTION W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W
The chemical fixation of carbon dioxide (CO2) has
received much attention from an environmental
protection view point. An attractive strategy to deal
with this situation is converting CO2 into valuable
substances. A useful method may be the application of CO2 as a monomer for the synthesis of
polymer materials [1, 2]. This could reduce the
greenhouse effect while reusing the carbon source
for useful polymer materials. The reaction of
epoxide polymers with CO2 is one of the most
inexpensive methods to incorporate CO2 into
organic compounds because of easy separation of
Received 25 September 2001
Accepted 24 December 2001

514 / Park et al.

SCHEME 1. Synthesis of poly(DOMA-co-St) via incorporation of CO2 into poly(GMA-co-St).

both original and produced polymers from the


reaction system as well as successful solid-state
reaction. Polymers containing cyclic carbonates can
be used as very useful functional polymers due to
their reactivity and high coordination ability [2].
The cyclic carbonates from the reaction of CO2
with epoxides have high polarity, high coordination ability and high reactivity, and therefore have
been widely used as aprotic dipolar solvents and
electrolytes for batteries [3]. Five-membered cyclic
carbonates can be readily prepared from oxiranes
and CO2 under mild conditions in quantitative
yield [46]. Thus, cyclic carbonates can be used as
synthetic equivalents of CO2, and production of
polymers using cyclic carbonate is one of the most
effective methods to incorporate CO2 into a polymer. The polymers containing pendant five-membered cyclic carbonate groups can also be
considered as starting materials for the synthesis
of functional polymers [4].
In our previous work [4, 6], we have reported
that poly(glycidyl methacrylate) (PGMA) was
converted to a polymethacrylate bearing a fivemembered cyclic carbonate group by the polymer
reaction with CO2 using a catalyst. In our present
work, the conversion of PGMA-like polymeric
epoxide based copolymers was attempted to the
corresponding polymers bearing five-membered
cyclic carbonate moieties by the reaction with CO2.
Blends of polymers bearing cyclic carbonate groups
with some commercial polymers have been developed in our laboratory for more versatile applications of the polymers [5, 6]. Polymer miscibility has
been the subject of considerable investigation from
both theoretical and practical standpoints. A
number of works of calorimetric, spectroscopic,
light or neutron scattering, and other experimental
techniques have been reported to investigate the
miscibility of polymer blends. Of particular interest
is the application of the FT-IR and differential
scanning calorimetry (DSC) to study the miscibility
of polymer blends [7, 8].
In this study, we report on the effect of the
structure of quaternary ammonium salt catalysts
on the kinetics in the synthesis of poly[(2-oxo-1,3dioxolane-4-yl)methyl
methacrylate-co-styrene]
Copyright 2002 John Wiley & Sons, Ltd.

[poly(DOMA-co-St)] from the addition of CO2 to


poly(glycidyl
methacrylate-co-styrene)
[poly(GMA-co-St)]. In order to expand the applications
of the CO2 fixed copolymer, blends of this
copolymer with poly(methyl methacrylate)
(PMMA) or poly(vinyl chloride) (PVC) were prepared and their miscibility behaviors were investigated by using DSC, FT-IR and the optical clarity
test.

EXPERIMENTAL W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W
Materials
Glycidyl methacrylate (GMA) and styrene monomers were washed with an aqueous sodium
hydroxide solution, rinsed with distilled water,
then dried over calcium chloride. Quaternary
ammonium salts, such as, tetrabutyl ammonium
chloride (TBAC), tetrabutyl ammonium bromide
(TBAB), tetrabutyl ammonium iodide (TBAI),
tetraoctyl ammonium chloride (TOAC) and tetrahexyl ammonium chloride (THAC) were all reagent grades (Aldrich) and were used as purchased
without further purification. N,N'-dimethylformamide (DMF, Junsei), dimethylsulfoxide (DMSO,
Junsei) and a,a'-azobisisobutyronitrile (AIBN, Junsei) were used as received. PMMA (Mw 120 000)
and PVC (Mw 106 000) were purchased from
Aldrich.
Synthesis of Poly(GMA-co-St)
A radical copolymer of GMA (20 g) with styrene
(5 g) [poly(GMA-co-St)] was prepared in DMSO
(250 ml) using AIBN (0.25g) as an initiator at 70 C
for 24 hr under nitrogen atmosphere, then the
solution was poured into distilled water to give a
precipitate. Polymers were recovered using an
excess of methanol, and dried in vacuum at 30 C
for 12 hr. The copolymer composition of poly(GMA-co-St) was determined from the ratio of area
in the copolymer using the 1H-NMR spectrum. The
ratio of area for the copolymer peak was 80.5 : 19.5
(GMA : St).
Polym. Adv. Technol., 13, 513521 (2002)

Synthesis of poly[(2-oxo-1,3-dioxolane-4-yl)methyl methacrylate-co-styrene] / 515

FIGURE 1. 1H-NMR spectrum of poly(DOMA-co-St).

Addition Reaction of Poly(GMA-co-St) and CO2


The synthesis of a copolymer [poly(DOMA-co-St)]
from poly(GMA-co-St) and CO2 was carried out
using quaternary ammonium salt as a catalyst, as
shown in Scheme 1. To a 300 ml three-neck pyrex
reactor containing a mixture of 16 g of poly(GMAco-St) and 150 ml of DMSO was introduced
0.5 mmol of catalyst, and the solution was heated
up to a desired temperature (100 C). The reaction
was started by stirring the solution under a slow
stream of CO2 (10 ml/min), and continued for
12 hr. The polymer obtained was purified by
precipitation in distilled water. The precipitate
was isolated by filtration, thoroughly washed with
distilled water, and dried under 30 C. Then, the
sample of the reaction mixture was taken and
analyzed to check the yield of CO2 in poly(GMAco-St) using gel permeation chromatography
(GPC). The yield of CO2 in poly(GMA-co-St) is
defined as follows:
Yield of CO2 addition %

or PVC with a given composition were cast from 10


wt% solution in DMF. The films were dried under
vacuum for 3 days at room temperature. Miscibility
of the blends was investigated by using DSC, FT-IR,
and optical clarity test.
Techniques
IR spectra were obtained by an Analect FX6160 FTIR spectrometer. The molecular weight of polymers was determined from GPC (Waters 244). The
measurement was conducted using a RI detector,
and DMF as an eluent with a flow rate of 1.0 ml/
min at 25 C. Polystyrene was used as a standard
for calibration. Glass transition temperatures (Tg)
were measured using DSC calibrated with pure
indium as a standard. Experiments were carried
out in a nitrogen atmosphere. In order to avoid the
thermal history from the samples packed in the
aluminum pan and to eliminate any small traces of

Number of unit of cyclic carbonate group in poly(DOMA-co-St)


 100
Number of unit of epoxide group in poly(GMA-co-St)

The identification of poly(DOMA-co-St) was performed by FT-IR and 1H-NMR spectroscopy.


Preparation of Blend Films
To prepare blend films, weighed amounts of
poly(DOMA-co-St) (19.5 wt% styrene) and PMMA
Copyright 2002 John Wiley & Sons, Ltd.

solvent, samples were heated to 450 K at a heating


rate of 10 C/min. All Tg values were taken as the
half-height point of the heat capacity jump in the
second scan. 1H-NMR spectra were recorded with
a Bruker 300 MHz NMR spectrophotometer. For
the measurement, 1.5 mg of sample was dissolved
in 0.5 ml of solvent (DMSO-d6) in a 5 ml tube at
25 C.
Polym. Adv. Technol., 13, 513521 (2002)

516 / Park et al.

FIGURE 2. FT-IR spectra of poly(GMA-co-St) and poly(DOMA-co-St).

RESULTS AND DISCUSSION W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W


Synthesis of Poly(DOMA-co-St) from Poly(GMAco-St) and CO2
Structure Identication of Poly(DOMA-co-St). The
conversion of epoxide ring in poly(GMA-co-St) to
the ve-membered cyclic carbonate group in
poly(DOMA-co-St) could be identied by 1HNMR and FT-IR spectra shown in Figs 1 and 2.
The characteristic peaks are as follows: 4.2  4.3
(OCH2, in side chain), 5.0  5.1 (HCO, in

cyclic carbonate), 4.54.7 ppm (OCH2, in cyclic


carbonate). The IR-spectrum of the poly(DOMAco-St) exhibited an absorption peak at 1800 cm 1
(C=O of cyclic carbonate), whereas the peak was
not observed on the IR spectrum for poly(GMA-coSt). Table 1 lists the molecular weights and Tg
values of poly(GMA-co-St) and poly(DOMA-co-St)
as well as PMMA and PVC.
Structure-activity Relationship of the Catalysts. The
anion and cation of quarternary ammonium salt

TABLE 1. Characteristics of Polymer Samples

Sample

Mwa

Mna

Mw/Mna

Tg ( C)b

Poly(GMA-co-St)c
Poly(DOMA-co-St)d
PMMAe
PVCe

204 000
261 000
120 000
106 000

104 000
130 000
82 000
60 000

1.95
2.01
1.42
1.77

130
61
114
90

Measured by GPC in our laboratory.


Measured by DSC in our laboratory.
c
Prepared with 20 g of GMA, 5 g of styrene and 250 ml of DMSO at 70 C for 24 hr.
d
Prepared with 16 g of poly(GMA-co-St), 150 ml of DMSO and 0.5 mmol of TOAC at 100 C for 12 hr.
e
Supplied by Aldrich Chemical Co., Inc.
b

Copyright 2002 John Wiley & Sons, Ltd.

Polym. Adv. Technol., 13, 513521 (2002)

Synthesis of poly[(2-oxo-1,3-dioxolane-4-yl)methyl methacrylate-co-styrene] / 517

FIGURE 3. Effect of anions of quaternary ammonium salt on the yield of CO2 addition.

signicantly affect the catalytic activity [5, 9]. In


order to understand the effects of the anions of the
catalysts in the reaction of poly(GMA-co-St) and
CO2, quaternary ammonium salt catalysts with
different anions (Cl , Br , I ) were used at 100 C.
A small portion of reaction mixture was taken and
analyzed to check the added amount of CO2 into
poly(GMA-co-St).
From Fig 3, for the reaction of poly(GMA-co-St)
and CO2, the yield of CO2 addition increases in
the order I < Br < Cl . When the different
halide ions were used for the quaternary ammonium salts, chloride was the most active. The
order of activity of halide anions is in accord with
the order of their nucleophilicity. The importance
of anion nucleophilicity confirms that the rate
determining step of the epoxide-CO2 reaction
involves nucleophilic attack of an anion to oxirane
[4, 10].
When different halide ions of quaternary
ammonium salts are used, the solvation of the
anions are an important factor for liquid phase
reactions [11]. In an aprotic solvent, like DMSO, in
the case of poly(DOMA-co-St) synthesis, stronger
solvation may be expected for a soft anion like I
than for a hard anion like Cl . Hence, the order of
the nucleophilicity of halides in SN2-type reaction
in an aprotic solvent, such as DMSO, will increase
in the order I < Br < Cl , which is in accordance
with the order of the reactivity of quaternary
ammonuim halides [5, 12]. It is suggested, therefore, that high nucleophilicity of Cl in an aprotic
Copyright 2002 John Wiley & Sons, Ltd.

solvent enhances the attack of the anion to epoxide


ring of poly(GMA-co-St).
In order to understand the effects of the cation
structure in the reaction of poly(GMA-co-St) and
CO2, quaternary ammonium chloride catalysts of
different alkyl cations [tetra octyl (TOAC), tetra
hexyl (THAC), and tetra butyl (TBAC)] were used.
Figure 4 shows that the yield of CO2 addition is
determined in the decreasing order of TOAC
> THAC > TBAC. The catalyst cation having a
large alkyl group is easily dispersed in organic
solvents because of the lipophilicity of the reaction.
A bulky quaternary salt, having longer distances
between cation and anion, is generally known to
exhibit higher activity in activating anions [4, 5].
For example, the cationanion interionic distances
of sodium bromide and tetrabutyl ammonium
, respectively
bromide ion pairs are 2.85 and 6.28 A
[11]. Although the difference in cationanion
interionic distance for the two ion pairs is small, it
can, when translated into differences in energies of
activation, be sufficient to cause the quaternary salt
to react up to four-orders of magnitude faster than
anions from the sodium salt. This trend explains
why TOAC is the most effective catalyst in the
nucleophilic attack of the anion to the oxirane of
poly(GMA-co-St).
Kinetic Studies in a Semi-batch Reactor. For the
addition reaction of CO2 to poly(GMA-co-St),
the following elementary reaction steps can be
proposed, where R = poly(GMA-co-St), P = polyPolym. Adv. Technol., 13, 513521 (2002)

518 / Park et al.

FIGURE 4. Effect of cations of quaternary ammonium salt on the yield of CO2 addition.

(DOMA-co-St), and QX = quaternary ammonium


salt catalyst.
k1

R QX RQX

k2

k3

RQX CO2 ! P QX

The reaction rate constants are k1, k2 and k3.


The rate of formation of P can be written as:
dP/dt k3 CO2 RQX

dP/dt k0 [R][QX]
3

Adopting steady-state approximation method for


the activated complex RQX*, the net rate of RQX*
formation can be written as eq. (4):


dRQX =dt k1 [R][QX]

k2 RQX

k3 CO2 RQX 0

Rearranging this equation, one can obtain eq. (5),


RQX k1 [R][QX]=k2 k3 CO2

Substituting eq. (3) with eq. (5), the rate of


formation of P can be written as
dP/dt k1 k3 RCO2 QX=k2 k3 CO2

When the addition reaction of CO2 to poly(GMAco-St) is carried out in a semi-batch reactor with a
constant flow of CO2, the concentration of polyCopyright 2002 John Wiley & Sons, Ltd.

(GMA-co-St) only varies since the concentration of


dissolved CO2 in the solvent can be assumed
constant.
Considering that [CO2] is constant in our semibatch operations, when the absorption rate of CO2
into the solvent can be assumed much faster than
that of CO2 addition reaction with poly(GMA-coSt):
7

where k' is k1k3[CO2]/(k2 k3[CO2]).


Since the catalyst concentration does not change
during the reaction, the pseudo first-order rate
equation can be applied.
dP/dt

dR/dt kR

Integration of eq. (8) gives eq. (9):


ln [poly(GMA-co-St)]0 =[poly(GMA-co-St)] kt
9
From the slope of the linear plot between ln
[poly(GMA-co-St)]0/[poly(GMA-co-St)]) v. time,
one can estimate the pseudo first-order rate
constant k. Figure 5 shows the plots of ln [poly(GMA-co-St)]0/[poly(GMA-co-St)]) v. time for the
different cations of quaternary ammonium salts.
Since good straight lines are obtained, the reaction
can be considered as pseudo first-order with
respect to [poly(GMA-co-St)]. From the slope, the
Polym. Adv. Technol., 13, 513521 (2002)

Synthesis of poly[(2-oxo-1,3-dioxolane-4-yl)methyl methacrylate-co-styrene] / 519

FIGURE 5. First-order plot of the CO2 addition reaction with various different cations of
quaternary ammonium salts at 100 C [k(hr 1) = rate constant].

reaction rate constant was determined as 0.417 hr 1


for TOAC, 0.324 hr 1 for THAC and 0.241 hr 1 for
TBAC, respectively. Similarly, the pseudo firstorder rate constant was determined as 0.126 hr 1
for TBAB and 0.096 hr 1 for TBAI.
The Miscibility of Poly(DOMA-co-St) and PMMA or
PVC. Blends of poly(DOMA-co-St) and PMMA or
PVC were prepared by the solution-casting method. The copolymers for blending were prepared
under the experimental conditions given in the
footnote of Table 1. In order to examine the degree
of miscibility of the poly(DOMA-co-St)/PMMA or
poly(DOMA-co-St)/ PVC blends, optical clarity
was rst investigated (Table 2). All the poly(DOMA-co-St) containing blends formed clear lms,

which seems to mean that the blends are miscible


over the whole concentration ranges. If the
refractive indices of the two polymers are not
sufciently different, however, a transparent blend
sometimes indicates that the size of any heterogenity present is much smaller than the wavelength of visible light. For a detailed study of
polymer miscibility, we measured the Tg values of
the blends. Each blend of different poly(DOMAco-St) compositions with PMMA or PVC exhibited
a single Tg value between the two Tg values of each
polymer. This result again conrms that these
blends are miscible over the entire composition
ranges.
FT-IR spectroscopy has been recognized as a
powerful tool for elucidation of structural informa-

TABLE 2. Optical Clarity and Glass Transition Temperatures of Poly(DOMA-co-St) with PMMA or PVC blends

Composition of
poly(DOMA-co-St)

Optical clarity
PMMA

Tg ( C)
PVC

PMMA

PVC

0
0.2
0.4
0.6
0.8
1.0

Clear
Clear
Clear
Clear

Clear
Clear
Clear
Clear

114
110
102
81
70
61

90
80
68
65
72
61

Copyright 2002 John Wiley & Sons, Ltd.

Polym. Adv. Technol., 13, 513521 (2002)

520 / Park et al.

FIGURE 6. FT-IR spectra of the carboxyl vibration region for poly(DOMA-co-St)/PMMA blends.

tion. The position, intensity, and the shape of


vibration bands are useful in clarifying conformational characteristics and determining the presence

of interactions between various groups due to the


sensitivity of the force constants to inter- and intramolecular interactions [13, 14].

FIGURE 7. FT-IR spectra of the carboxyl vibration region for poly(DOMA-co-St)/PVC blends.
Copyright 2002 John Wiley & Sons, Ltd.

Polym. Adv. Technol., 13, 513521 (2002)

Synthesis of poly[(2-oxo-1,3-dioxolane-4-yl)methyl methacrylate-co-styrene] / 521

FT-IR spectroscopy is particularly suitable for


investigating carbonyl concentrations in blends
[13]. Figures 6 and 7 show typical peaks of C=O
band vibration region for poly(DOMA-co-St)/
PMMA and poly(DOMA-co-St)/PVC blends. The
characteristic peak of C=O bond vibrations
appeared; vibrations assigned to free C=O groups
and hydrogen bonded C=O groups around 1730
and 1670 cm 1, respectively [15].
In the blends of poly(DOMA-co-St)/PMMA
and poly(DOMA-co-St)/PVC, a competition for
hydrogen bonding exists; intra-association and
inter-association between poly(DOMA-co-St) and
PMMA or PVC groups. This causes the appearance
of a hydrogen bond, this hydrogen bonded
carbonyl vibration, at a lower wavenumber as can
be seen in Figs 6 and 7 for the two different polymer
blends. In the blends, on increasing the composition of poly(DOMA-co-St), the relative absorbance
of the hydrogen bonded carbonyl vibration increase. Therefore, it can be concluded that the
miscibility of the blends based on the optical clarity
and the Tg behavior is attributed to the intermolecular interaction between the components. The
interaction may come from both carbonyl group
and proton group in poly(DOMA-co-St) and
PMMA or PVC.

CONCLUSIONS W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W
In the synthesis of poly(DOMA-co-St)] from poly(GMA-co-St) and CO2, quaternary ammonium salts
showed good catalytic activity. Also, the quaternary ammonium salts of larger structure and more
nucleophilic counter anion exhibited higher CO2
addition to the epoxide groups of poly(GMA-coSt). The addition reaction of CO2 to poly(GMA-coSt) can be considered to be pseudo first-order with
the concentration of poly(GMA-co-St). An integrated process has been developed for the catalytic
conversion of CO2 to useful polymer materials by
blending of poly(DOMA-co-St) with PMMA or
PVC. All the poly(DOMA-co-St) formed clear films
when blended with PMMA or PVC, which seems to

Copyright 2002 John Wiley & Sons, Ltd.

mean that the blends are miscible over the whole


concentration ranges. The Tg behavior of the
polymer blends from DSC thermograms also
confirmed the miscibility of the blends. The good
miscibility of blends of poly(DOMA-co-St) with
PMMA or PVC were explained by the intermolecular interaction between the components due to
the presence of carbonyl group vibrations, based on
FT-IR spectra.

ACKNOWLEDGEMENTS W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W
This work was supported by the Korea Research
Foundation (KRF-E00347) through the Research
Institute of Industrial Technology of Pusan
National University. C. S. Ha thanks the Center of
Integrated Molecular Systems, POSTECH, Korea.

REFERENCES W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W W
1. Koning C, Wildeson J, Parton R, Plum P, Steeman P,
Darensbourg DJ. Polymer 2001; 42: 3995.
2. Nishikubo T, Kameyama A, Sasano M. J. Polym. Sci.,
Polym. Chem. Ed. 1994; 32: 301.
3. Kihara N, Hara N, Endo T. J. Org. Chem. 1993; 58:
6198.
4. Sung CK, Kim KH, Moon JM, Chun SW, Na SE, Park
DW. J. Kor. Ind. & Eng. Chem. 1999; 10: 129.
5. Park DW, Moon JY, Yang JG, Jung SM, Lee JK, Ha CS.
Studies in Surface Science and Catalysis 1997; 61: 315.
6. Park SY, Lee HS, Ha CS, Park DW. J. App. Poly. Sci.
2001; 81: 2161.
7. Mekhilef N, Hadjiandreou P. Polymer 1995; 36: 2165.
8. Li D, Brisson J. Polymer 1998; 39: 793.
9. Nishikubo T, Kameyama A, Yamashita I, Tomoi M,
Fukuda W. J. Polym. Sci., Polym. Chem. Ed. 1993; 31:
939.
10. Kihara N, Endo T. Makromol. Chem. 1992; 193: 1482.
11. Starks CM, Liotta CL, Halpern M. Phase Transfer
Catalysis, chapter 1. Chapman & Hall: New York,
1994.
12. Kihara N, Endo T. Macromolecules 1992; 25: 4824.
13. Vermeesch I, Groeninckx G. Polymer 1995; 36: 1039.
14. Xiao C, Lu Y, Liu H, Ehang L. J. Macromol. Sci., Pure &
Appl. Chem. 2000; A37: 1663.
15. Li D, Brisson J. Polymer 1998; 39: 801.

Polym. Adv. Technol., 13, 513521 (2002)

Das könnte Ihnen auch gefallen