Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

European Polymer Journal 36 (2000) 767777

Photooxidative degradation of poly(alkyl methacrylate)s


Halina Kaczmarek a,*, Alina Kaminska a, Alex van Herk b
a
Faculty of Chemistry, Nicolaus Copernicus University, Gagarin 7, 87-100 Torun, Poland
Department of Polymer Chemistry, Eindhoven University of Technology, P.O. Box 513, 5600 MB, Eindhoven, The Netherlands

Received 13 November 1998; received in revised form 22 February 1999; accepted 23 April 1999

Abstract
Photooxidative degradation of four dierent poly(alkyl methacrylate)s: poly(methyl methacrylate) (PMMA),
poly(ethyl methacrylate), poly(n-butyl methacrylate) and poly(n-hexyl methacrylate) have been investigated using
FTIR, UVVis spectroscopy and gel permeation chromatography. The inuence of the ester group size on the
course of photochemical reactions in poly(alkyl methacrylayte)s has been estimated. It has been found that PMMA
undergoes slower photooxidation but faster photodegradation than those in higher poly(alkyl methacrylate)s. The
dierent behavior of the polymers studied is caused by the dierent reactivity of the macroradicals inuenced by the
dierent exibility and mobility of macrochains at room temperature. # 2000 Elsevier Science Ltd. All rights
reserved.

1. Introduction
Poly(alkyl methacrylate)s, among which poly(methyl methacrylate) (PMMA) is most popular and
best known, have been studied during several decades because of their broad practical applications.
Owing to their excellent properties (high transparency, light weight, good mechanical and electrical
properties, great resistance to high temperature,
aging and chemicals, easy formability, poly(alkyl
methacrylate)s are used in architecture, industry,
motorization (as constructional materials and organic
glasses in buildings, cars, ships, aircrafts), in agriculture, medicine, pharmacy as well as in textile, paper
and paint industry [1,2]. The ability of alkyl methacrylates to copolymerize with other monomers allows
to obtain special copolymers and terpolymers which
are used for modication of polymers such as poly(vinyl chloride), polyacrylonitrile, polystyrene, poly-

* Corresponding author. Fax: +48-56-6542477.

ethylene, phenolic resins [14] and also natural


polymers (as starch or jute) [5,6].
Physical properties of poly(alkyl methacrylate)s
change with the size of ester group. PMMA is a
hard, rigid material with high glass transition temperature Tg 11058C). Polymers become more soft
and exible (Tg decreases) with the increase of the
length of the side ester chain up to certain limit.
Polymerization kinetics of methacrylates is well
known specially in case of methyl methacrylate.
Intensive studies of polymerization kinetics of other
monomers from this family (as ethyl methacrylate,
butyl methacrylate, hexyl methacrylate, dodecyl
methacrylate) are carried out in dierent laboratories
and results are collected by the IUPAC Working
Party on Modeling of kinetics and processes of polymerization [711].
Photochemical properties of poly(methyl methacrylate) [1224] and their blends [2530] are described in
detail in the present literature but there are limited informations about the behavior of other poly(alkyl
methacrylate)s [12,13]. Dierent physical properties of

0014-3057/00/$ - see front matter # 2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 4 - 3 0 5 7 ( 9 9 ) 0 0 1 2 5 - 1

768

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

polymers varying the size of side group suggest that


also photochemical stability of these polymers should
be dierent.
Photochemical properties of these polymers are important with respect to aging but also should be considered when high power monochromatic lasers are
used in the pulsed laser polymerization technique [8].
Their simultaneous photodegradation and photopolymerization could interfere with an accurate determination of propagation rate coecients.
The aim of this work was to study the photooxidative degradation of four polymers obtained under the
same conditions: poly(methyl methacrylate) (PMMA),
poly(ethyl methacrylate) (PEMA), poly(n-butyl methacrylate) (PBMA) and poly(n-hexyl methacrylate)
(PHMA). General chemical formula of studied polymers is:

ring during 8 h. Because of the 4 h half-life time of


AIBN (at 708C), after a 4 h period, a new portion of
initiator was added. Polymers were precipitated two
times by methanol and dried under vacuum at room
temperature. The degree of conversion, estimated gravimetrically, was 9098%.
2.4. Photodegradation
For UV-irradiation, thin polymer lms were
obtained by solvent evaporation from 2% benzene solutions. Samples with the same thickness (10 mm) and
surface 2  2 cm) were selected and exposed to UV-irradiation in air atmosphere at room temperature. A
low pressure mercury vapour lamp (TUV-30W,
Philips, Holland) emitting mainly 254 nm radiation
was used as a light source. Light intensity at sample
position was measured by IL 1400A Radiometer
(International Light, USA) and equaled 2.65 mW/cm2.
2.5. Analysis

2. Experimental
2.1. Materials
The following substances were used in this study:
. Monomers (stabilized 100 ppm MEHQ; Rohm
GmbH, Germany):
methyl methacrylate (MMA),
ethyl methacrylate (EMA),
butyl methacrylate (BMA),
hexyl methacrylate (HMA).
. Initiator:
2,2'-azobis(isobutyronitrile),
AIBN
(Perkadox, Akzo Nobel, Holland).
. Solvents: toluene, methanol, tetrahydrofuran, benzene (Merck p.a.).

. UVVis spectra were recorded on an UV-1601PC


Spectrometer (Shimadzu, Japan). Thin polymeric
lms were supported by quartz plates.
. IR spectra were obtained using a FTIR Spectrum
2000 (Perkin Elmer, USA). The polymer solution
was cast directly onto a KBr window. The amount
of photoproducts in degraded samples was calculated as a surface area of an absorption band (integral intensity) or as a height of a peak. For
comparison of the rate and eciency of photoreactions in dierent samples, percentage changes were
calculated.
. Gel Permeation Chromatography (GPC) was made
on Waters chromatographic system (USA) equipped
with four linear columns (PLgel Mixed-B) and
Waters Model 410 refractive index detector.
Polymer solutions in tetrahydrofuran (1% w/v) were
prepared, and then ltrated over 0.2 mm lters.
Stabilized tetrahydrofuran was used as the eluent at
a ow rate of 1 ml/min. Narrow-distribution polystyrene standards (Polymer Laboratories) were used
for calibration.

2.3. Polymerization
Polymers were prepared by monomer polymerization
in toluene solutions at 708C using AIBN as initiator.
Monomers were puried by shaking with inhibitor
remover (Aldrich). The concentration of monomers
was 20 wt%, concentration of initiator 5 mmol/l.
Monomer solution was deareated by bubbling with
nitrogen for 10 min. Polymerization was carried out in
nitrogen atmosphere with a reux condenser. Reagents
were heated and continuously mixed by magnetic stir-

3. Results and discussion


3.1. Absorption spectra in UVVis region
UV-irradiation of poly(alkyl methacrylate) lms
causes an increase of absorbance at 230400 nm range.
The absorbance at 275 and 300 nm were chosen for
comparison of behavior of dierent samples. Around
275 nm a new absorption band is formed in all cases

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

769

than those at 275 nm but the shape of the kinetic


curves is similar. PEMA, PBMA and PHMA do not
show plateaus on the kinetic curves contrary to
PMMA. It indicates that in poly(n-alkyl methacrylate)s
with higher ester groups, processes of chromophore
formation are autoaccelerated after longer irradiation
times.
The increase of absorption in 230400 nm region
can be explained by formation of an unsaturated
group in combination with carbonyl groups [12,13].
3.2. FTIR spectra analysis

(example is shown in Fig. 1). Absorbance at 275 nm


increases fastest in PMMA, especially in the rst step
of the irradiation (to about 4 h) but after longer exposure (420 h), it does not change signicantly (Fig.
2). In PEMA, A(275) increases slower than that in
PMMA. In case of PBMA and PHMA, the increase of
A(275) precedes a small initial drop of this value.
These results indicate that chromophore formation
is retarded in PEMA, PBMA and PHMA comparing
to PMMA up to about 12 h of UV-irradiation.
However, after longer irradiation time (>15 h), the
increase of A(275) is faster in PBMA and PHMA than
those in PMMA and PEMA.
A similar trend is observed in Fig. 3, where the
changes of absorbance at 300 nm are plotted versus irradiation time. Here, the values of DA(300) are smaller

Photochemical reactions in poly(n-alkyl methacrylate)s can easily be detected by FTIR. Signicant


changes caused by UV-irradiation were observed practically in the whole IR region (example of IR spectra is
shown for PBMA in Fig. 4).
The new band at 31003600 cm1 (with maximum
at 3480 cm1) appears in UV-irradiated PMMA,
PEMA, PBMA and PHMA. Such a broad band is
typical for free (non-associated) and hydrogen bonded
hydroxyl/hydroperoxide groups, which cannot be distinguished. Kinetic curves in Fig. 5 (where the integral
intensity of band in 31003600 cm1 was plotted
against time of irradiation) indicate clearly that PBMA
and PHMA are more susceptible to photooxidation
leading to OH/OOH groups formation than PMMA
and PEMA.
The results of a spectral analysis in the carbonyl
region are more complicated. The total amount of carbonyl groups changes irregularly during UV-irradiation in all samples studied and any reasonable

Fig. 2. Changes of absorbance at 275 nm during UV-irradiation for PMMA, PEMA, PBMA and PHMA.

Fig. 3. Changes of absorbance at 300 nm during UV-irradiation for PMMA, PEMA, PBMA and PHMA.

Fig. 1. Absorption spectra of unirradiated PEMA (0) and


after 1, 4, 6, 14, 16, 18, and 20 h of UV-irradiation.

770

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

Fig. 4. FTIR spectra of PBMA after 0, 1, 2, 4, 6, 8, 10, 12, 14 and 16 h UV-irradiation: (a) hydroxyl/hydroperoxide + methyl/
methylene stretching vibrations, (b) carbonyl stretching vibrations, (c) ether group stretching vibrations + CH bending + skeletal
region (Q indicates the increase of absorption; q indicates the decrease of absorption with increasing time of irradiation).

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

771

Fig. 5. Kinetics of hydroxyl/hydroperoxide groups formation


in PMMA, PEMA, PBMA and PHMA (changes of integral
intensity in 31003700 cm1 region).

Fig. 7. Changes of the absorbance at 1710 and 1730 cm1 in


PMMA, PEMA, PBMA and PHMA caused by UV-irradiation.

conclusion cannot be drawn. It is obviously caused by


competitive reactions occurring simultaneously: decay
of ester groups and oxidation with formation of new
carbonyls. Both these processes are clearly seen from
the dierence spectra obtained by subtraction of spectra of UV-irradiated samples from spectra of the same
samples before degradation (Fig. 6). The absorbance
of the main carbonyl peak with a maximum at 1728
1731 cm1, attributed to ester groups, decreases continuously during UV-irradiation, but simultaneously
the development of carbonyl shoulders at 17071710
and 17541760 cm1 takes place. The rst absorbance
can be attributed to aliphatic ketones, the second one
to peracid and perester groups [32].

The data obtained by calculation of absorption


changes from dierence spectra are more clear (Fig. 7).
It is seen that ketones are created faster in higher
poly(alkyl methacrylates) than in PMMA. A similar relation (not shown here) was obtained for a second
increasing shoulder of the carbonyl band (01760
cm1). Above results suggest that susceptibility of
poly(alkyl methacrylates) to photooxidation increases
with the increase of the length of alkyl substituent in
ester group.
The same result was obtained on the base of mathematical analysis of absorption band in carbonyl
region (Fig. 8). Resolution of this complex band with
a computer program led to separation into four main

Fig. 6. FTIR dierence spectra of PHMA after 10 h UV-irradiation.

772

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

components of the carbonyl band. It was found that


the fraction of peracids/peresters (with absorption
maximum at 01760 cm1; peak 2 in Fig. 8) and the
fraction of ketones (with absorption maximum at
01710 cm1; peak 3 in Fig. 8) is larger in higher poly(alkyl methacrylate)s than those in PMMA after the

Fig. 9. Changes of the ether groups amount during UV-irradiation in PMMA, PEMA, PBMA and PHMA (absorbance
changes at 1240 cm1).

same irradiation time. For example, in PMMA


exposed during 16 h, the surface area of peak 2 and
peak 3 is about 15% and 16%, whereas in PHMA
irradiated 16 h, these amount 19% and 29%, respectively.
The destruction or abstraction of ester groups
during UV-irradiation, deduced from the decay of the
absorption at 1730 cm1, is slower in PMMA than
that in PEMA, PBMA and PHMA (Fig. 7).
Degradation of poly(n-alkyl methacrylate)s can also

Fig. 8. The components of IR carbonyl band obtained by resolution of original absorption spectrum for PMMA (a) and
PHMA (b) 16 h UV-irradiated (combination of Gaussian/
Lorentzian function and 25 iterations were applied).

Fig. 10. Changes of methyl/methylene groups amount during


UV-irradiation in PMMA, PEMA, PBMA and PHMA
(absorbance changes at 2950 cm1).

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

773

Table 1
 w and polydispersity M
 w =M
 n in UV n ), weight-average molecular weight M
Changes of number-average molecular weight M
irradiated poly(alkyl methacrylate)s
Sample
PMMA
PEMA
PBMA
PHMA

Before degradation

After 5 h UV-irradiation

After 10 h UV-irradiation

n
M

w
M

 w =M
n
M

n
M

w
M

 w =M
n
M

n
M

w
M

 w =M
n
M

22,460
11,250
26,970
18,780

57,680
39,420
73,560
63,590

2.6
3.5
2.7
3.4

8800
8120
10,410
9810

24,500
22,560
79,180
134,510

2.8
2.8
7.6
13.7

5230
4290
5920
5700

18,880
13,200
23,700
62,220

3.6
3.1
4.0
10.9

easily be monitored by changes of the amount of C


OC groups absorbing in the 10001300 cm1 range.
The absorption at 1240 cm1, attributed to asymmetric
stretching vibration na of COC bonds [31,32],
decreases during exposure to UV (Fig. 9), which is a
simple evidence of destruction reactions in side groups.
The drop of the amount of COC groups is faster
and more ecient in PMMA than in PEMA, PBMA
and PHMA.
The opposite trend is observed on the basis of
absorbance changes at 2950 cm1 (Fig. 10). This band
is assigned to the following methyl and methylene
group vibrations: ns CH3 O na a-CH3 ns a-CH3
ns CH2 , where na is asymmetric stretching, ns is symmetric stretching. CH2/CH3 group amount decreases
faster in higher poly(alkyl methacrylate)s than that in
PMMA. The decrease of methyl/methylene groups in
polymers is due not only to degradation with side
groups abstraction but also to oxidation. Polymeric
oxidation products undergo further photolysis and vol-

atile low molecular weight products can be evolved


according to a mechanism proposed previously
[12,13,15,17,19,23].
3.3. Gel permeation chromatography results
The results obtained for poly(alkyl methacrylate)s
from GPC are listed in Table 1. An example of
changes in molecular weight distribution (MWD)
curves is shown in Fig. 11.
Undegraded samples are characterized by molecular
weights in approximately the same range (10,000
70,000). All samples exhibit signicant decrease of
number-average molecular weight after 5 and 10 h
UV-irradiation indicating random chain scissions.
Table 2 shows the average number of chain scissions
(S ) per macromolecule, calculated according to the following equation [33]:


 n 0 M
 n t 1,
S M
 n 0 and M
 n t are the number-average molwhere M
ecular weight of polymer before and after t hours of
UV-irradiation, respectively. It is seen, that the number
of chain scissions is very high in PMMA after 10 h
UV-irradiation. This is not surprising because PMMA
is known as a polymer that easily undergoes degradation and depolymerization upon UV-irradiation [17
19]. PHMA seems to be a polymer more resistant to
degradation under these conditions. The small value of
S for PEMA after 5 h UV-irradiation arises surely

Table 2
The average number of chain scissions (S ) per macromolecule
in UV-irradiated poly(alkyl methacrylate)s

Fig. 11. Dierential molecular weight distribution curves of


PMMA before (a) and after UV-irradiation during 5 h (b)
and 10 h (c).

Sample

S after 5 h UV

S after 10 h UV

PMMA
PEMA
PBMA
PHMA

1.55
0.38
1.59
0.91

4.29
1.62
1.56
1.29

774

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

Table 3
Glass transition temperature of poly(alkyl methacrylate)s
Sample

PMMA
PEMA
PBMA
PHMA

Tg (8C)
From Refs. [1,21]

From Ref. [34]

105
65
20
5

94
71
26
1

from its relatively low starting molecular weight (Table


1).
However, in PBMA and PHMA, an increase of
weight-average molecular weight was observed after 5
 w in PHMA
h exposure to UV-light (the increase of M
exceeds 110%). Prolongation of irradiation time (up to
 w : Polydispersity
10 h) causes again the drop of M


Mw =Mn in UV-irradiated polymers (with exception
of PEMA) increases, i.e., they become more non-uniform. This means that besides main chain scission,
opposite reaction leading to the increase of molecular
weight takes place. It cannot be crosslinking because,
as was checked, there are no insoluble gel in any UVirradiated samples. The increase of molecular weight
can be caused by recombination of macroradicals hav-

ing unpaired electron on the chain-ends. Another


reason of the increase of molecular weight is further
polymerization of residual double bonds induced by
UV-irradiation. In case of the presence of double
bonds on the chain-ends such reactions cannot be
excluded. If two macromolecules react together, the
product has a combined molecular weight. In fact,
traces of unsaturated groups can be detected in undegraded samples in the IR spectra (at 947 and 967
cm1, Fig. 4(c)). However, the absorbance of C1C
groups was very low and its change was not considered. It is important to recall that double bonds are
formed during UV-irradiation of polymers, which was
monitored by UVVis spectroscopy.
Pairs of radical chain-ends may also disproportionate or undergo chain transfer reaction to another
macromolecule. These processes, however, do not lead
to the change of polymer molecular weight.
3.4. Discussion of the substituent role on photooxidative
degradation of poly(alkyl methacrylate)s
Both PMMA and PEMA exist in a glassy state at
room temperature because of the high glass transition
temperature (Table 3). As the length of the side chain
increases, Tg decreases and polymer becomes more
softer and extensible [1,2,34]. This partially explains

Scheme 1.

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

the dierent photochemical behavior of glassy polymers (PMMA and PEMA) as compared to rubberlike
polymers such as PBMA and PHMA. The main reason
for the lower oxidation rate in PMMA (in which Tg >
Toxidation is lower oxygen diusion rate. The exibility
of the macrochains in PBMA and PHMA is very high
at ambient temperature, thus, the diusion of oxygen
into the bulk polymer is easier than that in the rigid
matrix. Moreover, high mobility of macrochains and
macroradicals facilitates secondary reactions. This is
probably the other reason of the accelerated photooxidation occurring in these polymers.
Moreover, it is very probable that in the case of
poly(alkyl methacrylate)s with longer side chain,
photooxidation can lead to formation of new carbonyl
groups not only in main chain (Scheme 1, reaction 1)
but also in side groups (Scheme 1, reaction 2).
Faster photooxidation of higher poly(alkyl methacrylate)s can also be explained by more ecient

775

destruction (or abstraction) of ester groups in this


case. Macroradicals produced by ester group cleavage
easily react with atmospheric oxygen giving peroxy
radicals and nally, after secondary reactions, hydroxyl
and carbonyl groups are formed. Radical side chain
can abstract hydrogen atom from new macromolecules
and can induce deeper degradation (Scheme 2). It
seems that in PMMA such reactions are less ecient.
Macroradicals or small radicals formed in UV-irradiated PMMA undergo further reactions in the nearest
neighborhood because of the cage eect, typical for a
rigid matrix. This probably leads to the high eciency
of degradation observed in PMMA.
Most likely, the exibility and mobility of macromolecules have a greater inuence on the course of
photooxidative degradation of polymers than steric
factors. Steric factor and hindrance parameter increasing with the increase of substituent size [35] should
impede the oxygen access to polymer chains. However,

Scheme 2.

776

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777

Table 4
Eect of the length of ester substituent in poly(alkyl methacrylate)s on the rate and eciency of photoprocesses
Polymer

PMMA
PEMA
PBMA
PHMA

0
1
3
5

a
b

Photodegradation
(main chain scission
+ depolymerization)

Photodestruction or
abstraction of ester
groups

Photooxidation formation
of C1O, OH, OOH)

Chromophores
formation
(e.g. C1C)

Q q
a

In rst step of UV-irradiation.


In further step of UV-irradiation.

in the case of poly(alkyl methacrylate)s with the linear


substituent in the ester group, the eect of steric hindrance was not observed.
4. Conclusions
The increase of the side chain length in poly(alkyl
methacrylate)s causes not only large changes of physical properties but also signicant changes in their
photochemical resistance.
It was found that UV-irradiated PMMA undergoes
faster degradation and depolymerization but slower
oxidation and destruction of ester groups than those
observed in other poly(alkyl methacrylate)s. The
photochemical processes leading to chromophore formation depends also on the stage of UV-irradiation. In
the rst step of sample exposure to light (up to approximately 10 h), the rate and eciency of formation
of double bonds (and double bonds coupled with carbonyls) is larger in PMMA than in PEMA, PBMA
and PHMA. In the continued UV-irradiation (after
015 h), this reaction is slower and less ecient in
PMMA. This is schematically shown in Table 4.
We can conclude, that the higher exibility and
mobility of macrochains, caused by relatively long
alkyl chains in ester group, facilitates the oxygen diusion into bulk polymer, which is the main reason for
the acceleration of photooxidation in PBMA and
PHMA as compared to oxidation in PMMA and
PEMA. Moreover, owing to the large side group, the
mechanism of photooxidative degradation can be
expanded: oxidation can take place not only in the
main chain but also in the side substituents.
Acknowledgements
The authors wish to thank Ir Ingeborg Schreur and
Ing. Wieb Kingma from the Laboratory of Polymer
Chemistry (Faculty of Polymer Chemistry and

Technology, Eindhoven University of Technology) for


the help in synthesis of polymers and for the GPC
analysis.

References
[1] Mark HF. In: Bikales NM, Overberger Ch G, Menges
G, Kroschwitz JI, editors. Encyclopedia of polymer
science and engineering, vol. 1. New York: Wiley, 1985.
p. 234.
[2] Brandrup J, Immergut EH, editors. Polymer handbook,
3rd edn. New York: Wiley, 1989.
[3] Schmidt-Naake G, Stenzel M. Angew Makromol Chem
1998;254:55.
[4] Kim BS, Nakamura GI, Inoue T. J Appl Polym Sci
1998;68:1829.
[5] Ghos P, Dev D, Samanta AK. J Appl Polym Sci
1998;68:1139.
[6] Hebeish A, Beliakova MK, Bayazeed A. J Appl Polym
Sci 1998;68:1709.
[7] Hutchinson A, Paquet Jr DA, McMinn JH, Fuller RE.
Macromolecules 1995;28:4023.
[8] van Herk AM. JMS-Rev Macromol Chem Phys
1997;C37:633.
[9] Beuermann S, Buback M, Davis TP, Gilbert RG,
Hutchinson RA, Olaj OF, Russel GT, Schweer J, van
Herk AM. Macromol Chem Phys 1997;198:1545.
[10] Hutchinson A, Beuermann S, Paquet Jr DA, McMinn
JH. Macromolecules 1997;30:3490.
[11] Zammit D, Coote ML, Davis TP, Willett GD.
Macromolecules 1998;31:955.
[12] Rabek JF. Photodegradation, photooxidation and photostabilization of polymers: Principles and applications.
London: Wiley, 1975.
[13] Rabek JF. In: Polymer photodegradation: Mechanisms
and experimental methods. London: Chapman and Hall,
1995. p. 134.
[14] Chandra R, Saini R. J Macromol Sci -Rev Macromol
Chem Phys 1990;C30(2):155.
[15] Torikai A, Fueki K, Photochem J. Photobiol, Chem
1989;A49:273.
[16] Ueno N, Mitsuhata T, Sugita K, Tanaka K. ACS Symp
Series 1989;412:424.

H. Kaczmarek et al. / European Polymer Journal 36 (2000) 767777


[17] Gupta A, Liang R, Tsay FD, Moacanin J.
Macromolecules 1980;13:1696.
[18] Torikai A, Fueki K. Polym Photochem 1982;2:296.
[19] Torikai A, Ohno M, Fueki K. J Appl Polym Sci
1990;41:1023.
[20] Siampiringue N, Leca J, Lemaire J. Eur Polym J
1991;27:633.
[21] Mitsuoka T, Torikai A, Fueki K. J Appl Polym Sci
1993;47:1027.
[22] Lippert T, Wokaun A, Stebani J, Nuyken O, Ihlemann
J. Angew Makromol Chem 1993;213:127.
[23] Fritz PM, Zhu QQ, Schnabel W. Eur Polym J
1994;30:1335.
[24] Torikai A, Hattori T, Eguchi T. J Polym Sci: Part A:
Polym Chem 1995;33:1867.
[25] Torikai A, Segikawa Y, Fueki K. Polym Deg Stab
1988;21:43.
[26] Osawa Z, Fukuda Y. Polym Deg Stab 1991;32:285.

777

[27] Kaczmarek H, Decker C. Polym Networks Blends


1995;5:1.
[28] Kaminska A, Swia tek M, Kaczmarek H. Polish J Chem
1996;70:1257.
[29] Kaminska A, Swia tek M. J Therm Anal 1996;46:1383.
[30] Kaminska A, Swia tek M, Kaczmarek H. Polish J Chem
1997;71:1479.
[31] Nagai H. J Appl Polym Sci 1963;7:1697.
[32] Bellamy LJ. The infra-red spectra of complex molecules.
NY: Wiley, 1975.
[33] Sikkema K, Hanner MJ, Brennan DJ, Smith PB, Priddy
DB. Polym Deg Stab 1992;38:119.
[34] Donth E, Beiner M, Reissig S, Korus J, Garwe F,
Vieweg S, Kahle S, Hempel E, Schroter K.
Macromolecules 1996;29:6589.
[35] Elias HG. In: An introduction to polymer science.
Weinheim: VCH, 1997. p. 182.

You might also like