Sie sind auf Seite 1von 22

10 THE USE OF ESR

SPECTROSCOPY FOR
STUDYING
POLYMERIZATION AND
POLYMER DEGRADATION
REACTIONS
T. G. CARSWELL, R. W. GARRETT, D. J. T. HILL,
J. H. O'DONNELL, P. J. POMERY and C. L. WINZOR
Polymer Materials and Radiation Group, Department of Chemistry, University
Queensland, Brisbane, QLD 4072, Australia

10.1 INTRODUCTION
Electron spin resonance spectroscopy offers a unique technique to study the role
of radical species as intermediates in both polymerization and polymer degrada-
tion processes. The technique has been developed significantly since its introduc-
tion to chemical applications in the 1950s [1], with major advances in the
stability of the magnetic field, in the sensitivity to low radical concentrations—
and hence the limit of detection and measurement—and in data collection and
manipulation. ESR spectrometry enables both the identification of radicals
and the measurement of their concentration. It is a non-destructive technique and
spectra can be recorded both during polymerization, and, in suitable circumstan-
ces, during degradation of polymers [2].
New quantitative studies of free radical polymerization kinetics are currently
being undertaken by a number of researchers using pulse laser techniques [3,4].
However, the pulsed laser methods cannot be used for crosslinking systems.
There has also been development of a number of new theoretical models for
polymerization [5-7]. The experimental information available comprises mono-
mer concentration and polymer molecular weight. ESR spectroscopy offers the
possibility of an additional piece of information—the radical concentration
during polymerization. There is increasing interest in polymerization to high
conversion, which is of great practical importance, and ESR spectra can be
Polymer Spectroscopy. Edited by Allan H. Fawcett
© 1996 John Wiley & Sons Ltd
obtained readily under these conditions when other methods can be difficult to
apply [8].
Degradation of polymers is often understood from a practical viewpoint as
a deterioration in the properties of polymer materials leading to failure in service.
Changes in the molecular structure, and particularly the molecular weight, of the
polymer is the fundamental degradation process. High-energy radiation, e.g.
gamma-rays and electron beams, is an important cause of controlled polymer
degradation [9]. The degradation reactions usually involve free radical inter-
mediates, and therefore ESR spectroscopy is a valuable technique for investigat-
ing the chemical mechanism of degradation.
We have studied the degradation by high-energy radiation of a number of
families of polymers by using a variety of techniques, including ESR spectroscopy
[10, H]. In this paper we show the similarities and differences in the role of free
radicals in the radiolysis of poly(methyl methacrylate), polystyrene, and random
copolymers of methyl methacrylate and styrene.

10.2 EXPERIMENTAL
The ESR spectra were recorded using a Bruker ER200D spectrometer fitted with
a liquid nitrogen dewar for measurements at 77 K, and with a variable tempera-
ture cavity having a heated supply of cold nitrogen for temperature control from
100 to 400K.
Methyl methacrylate (MMA) was distilled under a reduced pressure of nitro-
gen; ethylene glycol dimethacrylate (EGDMA) was purified by passage through
an alumina column. Polymerization mixtures containing azobisiso- butyronitrile
(AIBN) as initiator were sealed in 3 mm i.d. quartz tubes under vacuum. Some
measurements were also made in 1 mm i.e. tubes in order to minimize increases in
the temperature of the polymerization which could result from the exothermic
heat of polymerization during the Norrish-Trommsdorff region. The conversion
of monomer ( C = C concentration) was measured by near-infrared Fourier
transform spectroscopy using the C = C H vibration at 6152cm" l .

CH3 CH3 CH3


I I I
CH2=C CH2=C C=CH 2
COOCH3 CO CO
I I
0-CH2-CH2-O
MMA EGDMA
Poly(methyl methacrylate), polystyrene and their random copolymers were
prepared by free radical polymerization in vacuum using AIBN as initiator.
Samples of the polymers were evacuated with heating and sealed in high-purity
quartz tubes. Cobalt-60 gamma-irradiation was carried out in liquid nitrogen
(77 K) and at ambient temperature (300 K) at a dose rate of « 1 kGy/h.

10.3 RESULTS AND DISCUSSION


10.3.1 FREE RADICAL POLYMERIZATION
10.3.1.1 Identification of the Radicals in the ESR Spectrum
The ESR spectrum is recorded in the first-derivative form as shown in Fig-
ure 10.1. A number of characteristics of the spectrum of a radical can be predicted
from its structure and used to identify the presence of the radical in an ESR
spectrum. These characteristics are illustrated in Figure 10.1, which is the ESR
spectrum of the methyl radical at 77 K. They include:
(1) g value—the centre of the spectrum—corresponding to the proportionality
between the magnetic field H and the microwave frequency, expressed in the
relationshipfcv= gfiH;
(2) the number of lines in the spectrum—resulting from interactions between
the unpaired electron spin on the radical and the nuclear spins of adjacent atoms,
particularly hydrogen;

Linewidth

Intensity

Lines

Figure 10.1 ESR spectrum of the methyl radical (CH3') showing the characteristic
features of g value, number of lines, relative intensities of the lines, hyperfine splitting (hfs),
line width, and line shape
(3) the relative intensities of the component lines of the spectrum of the
radical—frequently described by a binomial distribution of intensities;
(4) the hyperfine splitting, hfs, between the lines—this separation of the lines
depends on the electron spin on the radical site, the magnitude of interacting
nuclear spins and the conformation of the radical;
(5) the line widths—usually measured between the positive and negative peaks
of the derivative spectrum, or the width at half height of the absorption spectrum.
(6) the line shape—usually represented by a Gaussian or Lorentzian express-
ion, or a combination, depending on the curvature of the wings of the absorption
spectrum, reflecting the environment of the radical.

10.3.1.2 Measurement of Radical Concentration


ESR spectra are obtained as first-derivative spectra of signal intensity versus
magnetic field because of the method of observation of the absorption of
microwave power. Integration of the experimental spectrum gives the corre-
sponding absorption spectrum and a second integration gives the area of the
spectrum, which is proportional to the radical concentration provided that
microwave power saturation is avoided.
Saturation of the upper energy level of the unpaired spins can occur at high
microwave powers, the actual power depending on the relaxation time and hence
on the nature of the radical. Therefore for quantitative measurements of radical
concentrations the power dependence of the spectrum must be examined. Devi-
ation from linearity in a plot of spectrum area versus the square root of the
microwave power indicates the onset of microwave power saturation and the
upper limit for quantitative measurements of radical concentrations. Microwave
power saturation measurements for methacrylate propagation radicals during
polymerization is shown in Figure 10.2.
The area of the absorption spectrum does not yield an absolute value for
radical concentration. This must be obtained by calibration with a sample
containing a known concentration of radicals using standardized conditions of
measurement. A sample of pitch/KCl provided by Varian was used in the present
study, and this sample was calibrated with a measured concentration of recrystal-
lized diphenylpicrylhydrazine (DPPH) in benzene.

10.3.1.3 Monomer Concentration during Polymerization


We have utilized a variety of techniques for determination of the conversion of
monomer to polymer, by measurement of the concentration of C = C bonds at
different times during the polymerization of vinyl and allyl monomers. The 1 H
NMR spectra of samples quenched and dissolved in deuterated solvent showed
resonances due to the different H atoms of the monomer and polymer, and 1 H
NMR was a very good method for determination of conversion. However,
Area of ESR spectrum

1/2
(Microwave power)
Figure 10.2 Microwave power saturation plot of the area of the ESR spectrum versus the
square root of the microwave power: (•) (experimental values; ( ) linear relationship
assuming no saturation, S is the maximum power level that can be used without saturation
occurring
Fourier transform near-infrared spectroscopy, using the C = C H band at
6152 cm" 1 for methacrylates, enabled the conversion to be followed in a single
sample throughout the polymerization to high conversion. Parallel experiments
were performed for near-infrared spectroscopic determination of monomer con-
centrations and for ESR measurements of radical concentrations. A further advan-
tage of the near-IR method is its applicability to insoluble crosslinked systems.
A typical conversion curve for the polymerization of methyl methacrylate is
shown in Figure 10.3. The three regions in the conversion curve correspond to (1)
the pre-gel region, (2) the Norrish-Trommsdorff region between the gel point and
the glass point, and (3) the glass region. The steep rise in the polymerization rate
above the gel point is attributed to a marked decrease in the termination rate
parameter, kv and the decrease in polymerization rate to near zero above the
glass point is attributed to a decrease of several orders of magnitude in the
apparent propagation rate parameter, kp.

10.3.1.4 Radical Concentration during Polymerization


Typical plots of radical concentration versus time during the polymerization of
methyl methacrylate are shown in Figure 10.4. The lower limit of sensitivity of
glass point
% Conversion

gel point

Time / min

Figure 103 Typical conversion curve for polymerization of methyl methacrylate show-
ing the gel and glass points and the (1) pre-gel, (2) Norrish-Trommsdorff, and (3) glass
regions. Polymerization temperature 45 0C; [AIBN] = 0.1 M

current ESR spectrometers is « 1 0 " 7 M, which is not sufficient to enable measure-


ment of the concentration of propagating radicals during the polymerization of
most monomers below the gel point. An ESR spectrum of propagating radicals
can be obtained below the gel point by quenching the polymerization system
in liquid nitrogen and accumulating spectra at 77 K, which also utilizes the
favourable Boltzmann distribution at this temperature. Radical concentra-
tions obtained below the gel point by the quenching technique are also shown
in Figure 10.4.
In favourable circumstances, the radical concentration below the gel point may
be obtained by accumulation of spectra in situ. However, the limited time
available for accumulation and the changing nature of the spectrum with
conversion militate against this procedure. Radical concentrations obtained
above the gel point by the quenching and in situ methods (without accumulation)
were in good agreement.
Comparison of the monomer concentration (Figure 10.3) and the radical
concentration (Figure 10.4) during the polymerization shows that, whereas the
rate of conversion decreases very suddenly at the glass point, the radical
concentration continues to increase, but at a steadily decreasing rate. At long
polymerization times the radical concentration reaches a maximum and then
decreases due to the depletion of initiator, indicating that termination reactions
continue above the glass point, although more slowly.
[FT)/ mol dm3

Time/ min

Figure 10.4 The variation of the radical concentration [R#] with time during the
polymerization of methyl methacrylate at 45 0C for different concentrations of AIBN
initiator: (A) 0.2 M; (o) 0.1 Af; (•) 0.05 Af, spectra obtained in situ during polymerization in
the ESR spectrometer. (•) 0.05 M AIBN, spectra obtained by quenching to 77 K, enabling
accumulation of scans

103.1.5 Correction for Changing Sensitivity of the Spectrometer


The dielectric constant of the polymerizing system decreases during the polymer-
ization owing to changes in the molecular mobility of the polar ester groups and
the conjugation between C=C and C = O . This effect will be greatest in the
Norrish-Trommsdorff region, between the gel and the glass point. The changes
in the dielectric constant cause changes in (1) the frequency of resonance, and (2)
the sensitivity, or Q value, of the spectrometer, i.e. in the area of the spectrum
which would be obtained from a constant radical concentration.
A procedure which can be used to measure the change in sensitivity of the
spectrometer during polymerization is to record the spectrum of a reference
unpaired spin species with an ESR spectrum which does not overlap the spectrum
of the methacrylate propagating radical. We have found that the Mn 2+ species
(conveniently found in MgO) provides a suitable spectrum [12]. The Mn 2+ (in
MgO) was coated on the outside of the ESR tube as an external standard by
Time / min

Figure 10.5 Simultaneous measurement of the ESR spectra of Mn2+ and methacrylate
propagating radicals during the polymerization of methyl methacrylate: Mn area, sensitiv-
ity of the spectrometer; Mn field, conversion: [R*], corrected concentration of MMA
radicals

exposing the tube to the flame from burning magnesium metal. The variation in
sensitivity is shown in Figure 10.5 as the area of one peak in the spectrum of
Mn2 + . The increase in sensitivity occurs mainly in the Norrish-Trommsdorff
region and corresponds to a factor of % 3 in the area. The measured area of the
spectrum of the methacrylate propagating radical can then be corrected accord-
ing to the change in area of the Mn2 + peak.
An additional aspect of using Mn2 + to measure the changing sensitivity of the
spectrometer during the polymerization of methyl methacrylate was found to be
that the field position of the Mn 2+ peak also varied, and this variation could be
correlated with the C = C conversion. The position of the propagating radical
showed an identical variation in resonance frequency. Thus, the concentrations
of both R* and C = C can be obtained throughout the polymerization from the
ESR spectra. Typical data obtained for the polymerization of MMA at 600C with
AIBN initiator are shown in Figure 10.5.

10.3.1.6 Kinetic Analysis


The mechanism of polymerization of methyl methacrylate initiated by AIBN
involves the three kinetic steps:
Initiation
I—>2Y
Propagation
P;+M^P; + 1
Termination

The instantaneous rate of polymerization and the net rate of formation of


radicals are given by the following equations
-d[M]/df = fcp[F][M]
d[P]/dr = 2fc d /[I]-2fc f [P-] 2
where [M], [P'] and [I] are the concentrations of monomer, propagating
radicals and initiator, respectively, / is the initiator efficiency and kd is the rate of
decomposition of the initiator.
The values of fcp and kt can be obtained directly from these equations using
the experimentally determined values for [M] and [P - ], provided that /
remains constant. We have shown that a suitable manipulation of the second
equation enables the values of / and kt to be derived for incremental increases
in conversion. Good agreement is obtained with current theories of free
radical polymerization for the polymerization of methyl methacrylate at 60 0C
[13].

10.3.1.7 Crosslinking Methacrylate Monomers


A major objective of the current research programme is to extend the treatment of
polymerization kinetics based on direct measurements of monomer and radical
concentrations to crosslinking systems. Conventional methods for measurement
of monomer concentrations are not suitable, as they require soluble polymer. We
have been able to apply our procedure for utilizing the near-infrared spectrum of
the C = C bond in methyl methacrylate to systems containing ethylene glycol
dimethacrylate (EGDMA) [14].
Figure 10.6 shows the variation of the concentration of C = C bonds during the
polymerization of a MMA-EGDMA mixture containing 36% of EGDMA. The
initial rate of polymerization is much higher than for MMA, and this can be
attributed to the absence of a pre-gel region. However, the plateau region of the
conversion is lower. This results from the crosslinking reaction, with the radicals
and C = C bonds immobilized in the network.
% Conversion

Time / min

Figure 10.6 Dependence of C = C concentration on polymerization time for (a) MMA


and (b) a MMA-EGDMA mixture (36% EGDMA): polymerization temperature 600C;
initiator 0.05 M AIBN
[R*]x10- 6 /moldnrr 3

Time / min

Figure 10.7 Dependence of radical concentration [R#] on polymerization time for (a)
MMA and (b) a MMA-EGDMA mixture (36% EGDMA): polymerization temperature
600C; initiator 0.05 M AIBN. Note that the concentration scales differ by a factor of 40

The radical concentration in MMA-EGDMA mixtures might be expected to


be higher than in MMA owing to the retardation of the termination reaction by
the network. In Figure 10.7 the absence of a pre-gel region is indicated by the
increase in radical concentration from the beginning of the polymerization. The
most significant difference between the MMA-EGDMA mixture and MMA is
the radical concentration in the plateau region. It is approximately 40 times
greater in the mixture, and is almost millimolar.
Kinetic analysis of the propagation and termination rate constants in the
polymerization of MMA-EGDMA mixtures is more difficult than for MMA. An
understanding of the polymerization depends on knowledge of the individual
concentrations of C = C bonds that are (1) in monomer molecules and (2)
attached to polymer molecules. This information may be obtainable for NMR
spectra utilizing differences in the relaxation times of the two types of C=C
environments. However, it is likely that the polymerization is heterogeneous with
a non-spatially random distribution of crosslinks.

10.3.2 POLYMER DEGRADATION BY


HIGH-ENERGY RADIATION

The effects of high-energy radiation, principally y-rays and electron beams, on


polymers have been studied extensively for many years. The main interest has
been in the changes in material properties, such as strength and elongation. These
changes in properties have been related to changes in molecular weight of the
polymer molecules, either by main-chain scission or by chain crosslinking,
frequently with the formation of an insoluble gel fraction. The applications of
radiation effects on polymers have been two-fold: (1) the use of polymer materials
in radiation environments, such as nuclear reactors, and more recently in space,
(2) modification of the properties of polymer materials by reduction in molecular
weight or crosslinking, especially in the microelectronics industry as electron
beam resists, and perhaps in the future as X-ray resists.
Fundamental understanding of the mechanism of degradation of polymers by
high-energy radiation has been based mainly on structural changes observed in
the polymers, and to a much smaller extent on measurements of small molecule
products. ESR has been used since 1960 to observe radicals produced in
irradiated polymers, and hence to provide evidence for intermediate species in the
radiolysis. However, recent improvements in the stability and sensitivity of ESR
spectrometers and in computer manipulation of the spectra have enhanced the
use of this technique.
The capabilities of the ESR technique for providing fundamental information
about the mechanism of radiation degradation of polymers are shown in
observations on gamma-irradiated poly(methyl methacrylate), polystyrene and
their random copolymers.

10.3.2.1 Poly(methyl methacrylate)


The ESR spectrum of poly(methyl methacrylate) at 300K, shown in Fig-
ure 10.8(a), is well known. This characteristic 13 line (9 + 5) alternating spectrum
Figure 10.8 ESR spectra of poly(methyl methacrylate) after y-irradiation in vacuum:
radiation dose 1 kGy; radiation temperature (a) 300 K, (b) 77 K

has been the subject of much debate, but it is now generally accepted to be due to
the methacrylate propagating radical with two conformations.
The ESR spectrum after irradiation at 77 K, shown in Figure 10.8(b), is quite
different from the spectrum after irradiation at 300K. It is evidently due to
a number of radicals; the propagating radical is not a significant component at
this temperature, and must be formed by subsequent reactions of the radicals
produced at 77 K.

Analysis of spectra
A variety of techniques can be used to analyse for the component radicals in an
ESR spectrum. They include:
(1) dose saturation—spectra are obtained after irradiation to a series of
radiation doses. The yield of trapped radicals does not increase linearly with dose
above certain doses which are characteristic of particular radicals. In particular,
radical ions show 'dose-saturation' at low doses;
(2) microwave power saturation—the observation of an ESR spectrum de-
pends on the relaxation of the radicals (which are excited into the higher energy
level by the microwave radiation) back to the lower energy level in accordance
with the requirements of the Boltzmann distribution. Some radicals, and particu-
larly radical ions, have a slow relaxation and hence they will not be observed at
high microwave powers;
(3) Photobleaching—radical ions can be distinguished from neutral radicals
by irradiation with visible light above a critical wavelength, which is usually
«500nm. The radical ions are bleached and disappear, whereas the neutral
radicals are unaffected. The efficiency of this technique does depend on the
interaction of the light with the radicals and hence requires a transparent or finely
powdered sample;
(4) Fourier transform masking—the ESR spectrum can be converted to its
Fourier transform in the frequency domain. Lines in the original spectrum with
different line-widths can be separated by masking of different parts of the
spectrum and then conversion back into the original domain;
(5) accumulation of spectra—the signal to noise ratio must be high to enable
separation of different radicals in a spectrum, especially if some of the radicals are
present in small proportion of the total spectrum. Accumulation of the spectra,
possible with high stability of the magnetic field, enables the signal to noise ratio
to be enhanced. The effect of spectral accumulation is illustrated in Figure 10.9
for poly(a-methylstyrene);

Figure 10.9 ESR spectrum of poly(a-methylstyrene) after y-irradiation in vacuum at


300 K (dose 6 kGy): (a) one scan of 200 s; (b) 150 scans
(6) thermal annealing—when a number of different types of radicals are
trapped in a polymer during irradiation at a particular temperature, they will
react in different ways at different temperatures on subsequent heating. The
process of radical disappearance is known as thermal annealing, and can be used
to distinguish different radicals present at a lower temperature. The greatest
number of radicals will be produced by irradiation at the lowest possible
temperature. Usually, liquid nitrogen is used to enable irradiation at 77 K for this
reason, and the procedure is known as cryogenic trapping;
(7) subtraction techniques—recording of ESR spectra by computer enables
a variety of computational procedures to be used to identify and quantify the
component radicals and their reactions. In particular, subtraction of spectra
obtained after progressive stages of warming (thermal annealing after cryogenic
trapping) will frequently show a triplet or other spectrum characteristic of
a particular radical which has disappeared. The effectiveness of this procedure is
enhanced if the sample is cooled back to the same reference temperature to record
the spectrum after each warming step;
(8) simulation—confirmation of the presence of different types of radicals and
estimates of their proportions, and hence of their concentrations, can be obtained
by simulation of the ESR spectrum. This procedure requires values for the
parameters of the spectrum, Le. g value, number of lines, relative intensities of the
lines, hyperfine splittings, line-widths, line shape (Gaussian or Lorentzian or
a mixture). Simulated spectra can be computed for a wide variety of values for the
different parameters, the simulated ESR spectrum being matched to the experi-
mentally observed spectrum.
We have utilized all of these techniques to analyse the ESR spectrum of
poly(methyl methacrylate) at 77 K after y-irradiation. The progressive disappear-
ance of different types of radicals, and the formation of the chain scission radical
(which is the same radical as the propagating radical), eventually as the only
species, during thermal annealing after cryogenic trapping, are shown in Figure
10.10.
We have identified seven different radical species A-G, including the polymer
radical anion G, in the ESR spectrum of poly(methyl methacrylate) at 77 K after
y-irradiation in vacuum (the spectrum shown in Figure 10.8(b)), as follows:
CH3 I I CH3
I 1 . 1 I
CH3- -CHO -COOCH3 - C H 2 - C - -C—CH-C—CH2-C-
COOCH^ I ' COOCH2
A B C D E F
The progressive disappearance of these radicals, based on the spectra shown in
Figure 10.10, is shown in Figure 10.11. The five regions shown in Figure 10.11
correspond to the disappearance of different radicals as follows: stage (1) A, B and
C; stage (2) C and D; stage (3) E; stage (4) propagating radical (F) is the only
species present and is stable; (5) F.
The mechanism of the degradation of poly(methyl methacrylate) by y-radi-
ation can be deduced on the basis of the disappearance of radicals shown in
Figure 10.11. This mechanism (Scheme 1) is consistent with the formation of
molecular products, as previously reported.

CH 3 CH 3 O
I I Il
^ C H 2 - C ^ - - C H 2 C - C H 2 - + - C - O — C H 3 + OTHERS
#CH
C=O 3
O -CHO
(LjJ3 -CO-O-CH2

CH 3 CH 3 CH 3
I I I
SCISSION
-CH2C-CH2- • -CH2-C- + CH2=C-CH2-
C=O
O
I
CH 3

-CO-O-CH3 CO + CO 2 + CH 4 + CH 3 OH
Scheme 1 Mechanism of degradation of poly(methyl methacrylate) by y-radiation de-
duced from ESR studies of radical intermediates

10.3.2.2 Polystyrene
The ESR spectrum of polystyrene after y-irradiation in vacuum at 77 K is shown
in Figure 10.12(a). The spectrum is quite different from that of poly(methyl
methacrylate). It can be assigned to three species: (1) the oc-carbon radical, (2) the
cyclohexadienyl radical, and (3) a radical anion. The proportions of cyclo-

a-carbon cyclohexadienyl
Figure 10.10 ESR spectra of poly(methyl methacrylate) after y-irradiation at 77 K and
progressive wanning to 300 K. All spectra (except that at 77 K) were recorded on cooling
back to 140K after 10 min at the specified temperature

hexadienyl radicals and of radical anions are strongly dose-dependent, which


provides a method for their assignment.
The two neutral radicals are consistent with crosslinking being the major effect
of radiation on polystyrene, in contrast to main-chain scission in poly(methyl
methacrylate).
The ESR spectrum after irradiation at 300K, shown in Figure 10.12(b), is
similar to the spectrum at 77 K except that the centre line is reduced, owing to the
absence of the radical anion, and the large outer peaks of the 'triplet' show greater
resolution into subsidiary peaks.

10.3.2.3 Random Copolymers of Methyl Methacrylate and Styrene


The effect of high-energy radiation on random copolymers of styrene and methyl
methacrylate provides an excellent system for testing hypotheses for intra-
Btfrel

T / K
Figure 10.11 Decrease in radical concentration in poly(methyl methacrylate) after 1 kGy
of y-irradiation at 77 K on progressive warming to 360K. The numbers refer to stages of
radical reactions during wanning

molecular and inter-molecular interactions of energy transfer and radical reac-


tions.
The ESR spectra of a series of copolymers of styrene and methyl methacrylate
across the composition range between the two homopolymers are shown in
Figure 10.13 after irradiation at 300K. The spectra show a progressive change
between the spectra of the homopolymers, but it is apparent, e.g., considering the
copolymers containing 20% and 50% of styrene, that the proportions of'styrene'
radicals in the spectra are greater than the proportions in the compositions of the
copolymers. Thus, there is a preference for the formation of styrene radicals. This
effect has one component occurring during irradiation at 77 K (attributed to
energy transfer) and another component occurring during warming to 300 K (or
occurring during irradiation at 300K), which we have attributed to radical
transfer reactions.
This preference for the formation of styrene radicals is a manifestation of
a protective effect by styrene units on the degradation of methacrylate units in the
copolymer. The protective effect is also shown by the variation in the yield of
radicals, G(R"), with the composition of the copolymer. Figure 10.14 shows how
the value of G(R*) in the copolymers is always much less than the value which
would be obtained from the additivity of the electron densities of the two
monomer units. The protective effect is even greater at 300 K than at 77 K, which
is consistent with the additional mechanism of protection which occurs above
77 K.

10.3.2.4 ESR and the Mechanism of Radiolysis


The number of different types of radicals observed in the ESR spectra of
irradiated polymers is always greater after irradiation at 77 K than at 300 K. The
Figure 10.12 ESR spectra of polystyrene after y-irradiation in vacuum. Radiation dose
10OkGy; radiation temperature: (a) 77 K, (b) 300K

Scheme 2 ESR procedure of cryogenic trapping and thermal annealing to provide an


understanding of reactions which occur rapidly during the irradiation of polymers at
0%STY 8% STY

20% STY 50% STY

100% STY

Figure 10.13 ESR spectra of random copolymers of styrene and methyl methacrylate
after y-irradiation in vacuum at 300K. Radiation dose 3kGy. The compositions of the
copolymers are specified in mol% styrene
G(R)
G(R-)

% STYRENE

Figure 10.14 Protective effect against degradation by y-radiation provided by styrene


units in random copolymers of styrene and methyl methacrylate, shown by the radical
yields G(R#) derived from the ESR spectra, (a) Experimental values for y-irradiation at
77 K; (b) experimental values for y-irradiation at 300 K. The lines correspond to the G(R*)
values for the copolymers calculated from the G(R') values for poly(methyl methacrylate)
and polystyrene based on the additivity of electron densities
ESR spectra are usually similar after irradiation at 300K or irradiation at 77 K
and warming to 300 K, although the concentrations of radicals may be different.
A model for the mechanism of the radiolysis can then be deduced from the
radicals which are observed at 77 K and the reactions which they undergo on
warming. The procedure is outlined in Scheme 2.

10.4 CONCLUSIONS
ESR provides a powerful technique for developing a fundamental understanding
of the mechanism and kinetics of free radical polymerization and of the mechan-
ism of degradation of polymers by high-energy radiation. The assignment of ESR
spectra to component radicals and the measurement of the concentrations of
these radicals require a variety of experimental and computational procedures,
which have been greatly enhanced by improvements in spectrometer perform-
ance and computer capabilities.

10.5 ACKNOWLEDGEMENTS
The authors are grateful to the Australian Research Council and the Australian
Institute of Nuclear Science and Engineering for supporting this research, and to
the Australian Nuclear Science and Technology Organization for the provision
of irradiation facilities.

10.6 REFERENCES
[1] P.B. Ayscough, Electron Spin Resonance in Chemistry, Methuen, London, 1967.
[2] D.J.T. Hill, J.H. O'Donnell and PJ. Pomery, in Electron Spin Resonance, Royal
Society of Chemistry Specialist Periodical Reports, Vol. i3A, Cambridge, 1992,
p. 202.
[3] O.F. Olaj, I. Bitai and F. Hinkelman, Makromol. Chem., 1987,188,1689.
[4] T.P. Davis, K.F. O'Driscoll, M.C. Piton and M.A. Winnik, Macromolecules, 1989,
22, 2785.
[5] S.K. Soh and D.C. Sundberg, J. Polym. ScL, Polym. Chem. Ed., 1982,20,1345.
[6] MJ. Ballard, R.G. Gilbert, D.H. Napper, PJ. Pomery, P.W. O'Sullivan and J.H.
O'Donnell, Macromolecules, 1986,19,1303.
[7] G.T. Russell, D.H. Napper and R.G. Gilbert, Macromolecules, 1988, 21, 2141.
[8] R.W. Garrett, DJ.T. Hill, J.H. O'Donnell, PJ. Pomery and CL. Winzor, Polym.
Bull, 1989,22,611.
[9] M. Dole (Ed.), The Radiation Chemistry of Macromolecules, Academic Press, New
York, 1972.
[10] R.W. Garrett, DJ.T. Hill, TT. Le, J.H. O'Donnell and PJ. Pomery, Radiat. Phys.
Chem., 1992,39,215.
[11] DJ.T. Hill, S.Y. Ho, J.H. O'Donnell and PJ. Pomery, Radiat. Phys. Chem., 1990,36,
467.
[12] T.G. Carswell, DJ.T. Hill, D.S. Hunter, PJ. Pomery, J.H. O'Donnell and CL.
Winzor, Eur. Polym. J., 1990, 26, 541.
[13] T.G. Carswell, DJ.T. Hill, D.I. Londero, J.H. O'Donnell, PJ. Pomery and CL.
Winzor, Polymer, 1992,33,137.
[14] T.G. Carswell, DJ.T. Hill, R. Kellman, D.I. Londero, J.H. O'Donnell, PJ. Pomery
and CL. Winzor, Makromol. Chem., Macromol. Symp., 1991,51,183.

Das könnte Ihnen auch gefallen