Вы находитесь на странице: 1из 14

Makromol. Chem.

189,2381 - 2394 (1988) 2381

A new method for monitoring ultra-fast


photopolymerizations by real-time infra-red (RTIR)
spectroscopy
Dedicated to Prof. P. Rempp on the occasion of his 60th birthday

Christian Decker *, Khalil Moussa

Laboratoire de Photochimie General associC au CNRS, Ecole Nationale Superieure de Chimic:,


3 rue A. Werner, 68200 Mulhouse, France

(Date of receipt: March 29, 1988)

SUMMARY:
An original method based on IR spectroscopy has been developed in order to follow real-tirr.e
photopolymerizations that take place in less than one second. Conversion versus time curv(:s
were directly recorded for various multiacrylic monomers irradiated in condensed phase, thtls
allowing an immediate evaluation of both the rate of polymerization and the amount of residu il
unsaturation in the UV-cured polymer. This method proved well suited to study the effect c n
the polymerization rate of the photoinitiator efficiency, the monomer reactivity, the light inten-
sity, the film thickness, and the 0, inhibition. The dark polymerization that develops just aftcr
the UV exposure was investigated as a function of the degree of conversion and was shown 10
represent up to 60% of the total process. A comparison with other methods of kinetic analysis
shows the unique advantages of real-time infra-red (RTIR) spectroscopy which permits I o
follow quantitatively and in situ ultra-fast polymerizations.

Introduction
Multifunctional monomers are known to undergo a rapid polymerization whc n
they are exposed to intense UV radiation, even in the presence of air, leading to high y
crosslinked insoluble polymers'). This UV curing technology has found widespread
applications, in particular for the surface protection of different types of materials tly
fast drying clear coatings, lacquers, or printing inks, as well as for the production of
instantly hardening composites and adhesives2). Most of the UV-curable resins usc d
today are made of acrylic monomers and oligomers, owing to their high reactiviti,
moderate cost, and high quality of the final product, which exhibits remarkable
mechanical and optical properties.
Several methods have been employed in order to follow the kinetics of the photo-
polymerization process that transforms rapidly the liquid resin into a solid material.
They consist either of discrete measurements like in infra-red spectro~copy~,~),
13C NMR5), and FTIR p h o t o a c o ~ s t i c spectroscopy
~~~) or of a continuous monitoring
of the reaction in real time, like in methods based on dilato-
metry'1s12),IR radiometry13),laser nephel~metry'~*I~), and interferometry16).Exce3t
for the laser techniques, real-time methods are mainly used to follow polymerizations
performed at low light intensities, owing to their relatively long response, with typical
cure times in the range of 5 to 100 s. This requires to operate in an inert atmosphere m
order to avoid the well-known oxygen inhibition effect that becomes overwhelming

0025-116)</88/$03.00
2382 C. Decker, K. Moussa

under these conditions. With such slow techniques, it is therefore not possible to
analyse polymerizations that develop within 1 s or less, which is the usual time scale
for the UV hardening of most industrial UV-curable systems). Moreover, essentially
all of these methods proved inadequate for evaluating the precise amount of
unsaturation that remains in the UV-cured polymer.
IR spectroscopy alone permits both to follow quantitatively the polymerization by
monitoring the disappearance of the IR absorption characteristic of the polymer-
izable double bond and to determine at any moment the actual degree of conversion
and hence the residual unsaturation content. This analytical method has proved
extremely valuable for measuring the polymerization rates and quantum yields of
reactions that develop in the millisecond time scale4,). It still suffers the major
disadvantage of not being a real-time kinetic method, thus requiring tedious point-by-
point measurements. Furthermore, like the other methods, it integrates into the final
result the post-polymerization effect which may develop during the few seconds time
lapse between the end of the UV exposure and the actual measurement.
In this paper, we describe an original method, based on IR spectroscopy, which
combines the advantages of the various techniques and allows to follow quantitatively
and in real time the photopolymerization of multiacrylic monomers under intense
illumination and in the presence of air, and to determine precisely the amount of
residual unsaturation in the UV-irradiated material. Another interest of real-time
infra-red (RTIR) spectroscopy is that it permits to investigate accurately the
additional polymerization that occurs just after the end of the irradiation and which
was found to represent an important part of the final degree of polymerization in UV-
cured multiacrylates.

Experimental part
Materials
The photopolymerizable resin consisted of 3 main components: (i) a photoinitiator that
cleaves readily into radicals upon UV exposure, like a,a-dimethoxyphenylacetophenone
(Irgacure 651 from Qba-Geigy); (ii) an acrylate end-capped oligomer consisting of an aliphatic
polyurethane (Actilane 20 from SNPE) or of a polyester (Ebecryl830 from UCB) or of a deriva-
tive from the glycidyl ether of bisphenol Aa) (Ebecryl605 from UCB), incorrectly called epoxy-
acrylate; (iii) a reactive acrylic diluent, tripropylene glycol diacrylateb) (TPGDA from UCB),
hexanediol diacrylatec) (HDDA from SNPE), trimethylolpropane triacrylated) (TMPTA from
UCB), ethyldiethyleneglycol acrylatee) (EDGA from SNPE), or an oxazolidone monoacrylate
(Acticryl CL 959 from SNPE).
Typical formulations contained 5 % of photoinitiator and equal parts of the acrylic oligomer
and diluent. The resin was applied on an NaCl salt disc as a uniform layer of thickness in the
range 3 to 30 prn. The coated disc was placed into an infra-red spectrophotometer, the IR wave-
length being fixed at 810 cm- where the acrylic double bond exhibits a sharp and distinct ab-

a) Systematic name: 2,2-Bis[4-(2,3-epoxypropoxy)phenyl]propane.


b, Systematic name: diacrylate.
1,4,7-Trimethy1-3,6-dioxaoctamethylene
Systematic name: Hexamethylene diacrylate.
d, Systematic name: 2,2-Dimethylenebutyl triacrylate.
e, Systematic name: 3,6-Dioxaoctyl acrylate.
A new method for monitoring ultra-fast photopolymerizations . . . 2383

sorption. The detection of the IR signal was made by operating the spectrophotometer in the
absorbance mode.

Real-time infra-red (R TIR) spectrocopy


The sample was exposed for a short time to the radiation of a medium pressure mercury lamp
(Philips HPK-125 W), in the presence of air, since UV curing is performed under these cond -
tions in most industrial applications. Fig. 1 describes schematically the simple experimental set -
up devised in order to expose the sample at the same time to the polymerizing UV beam and the
analyzing IR beam. The UV irradiance at the sample position was measured by actionmetry and
was typically in the order of einstein . s - ' . cm-2 or 40 W . m-2. If desired, photopol)-
merizations could be carried out in a controlled atmosphere by flushing the air-tight spectrcm-
photometer chamber with an appropriate gas, like nitrogen or pure oxygen.

Fig. 1. Experimental
set-up for kinetic study
Fssl-
speed
Resorde
1'
--l
Absorbance

time
,**I" film
NaCl dlrk

of photopolymeriza-
tions by real-time I R Spectrophotomstsr

infra-red spectroscopy L

As soon as the sample was exposed to UV radiation, the IR absorbance at 810 cm- droppcd
rapidly as a result of the reaction of the acrylic double bonds. This sharp decrease, which
accurately reflects the extent of the polymerization process, was monitored continuously and n
real time on the spectrophotometer recorder operating at maximum speed (5 c m . S K I ) . For
polymerizations that last for more than a few seconds, the IR signal was followed on a separate
recorder with scanning speeds in the range 0,l to 5 cm . s - ' . The full-scale response time of tlie
recorder was measured to be 0,l s, which allows a correct kinetic analysis of the photopoly-
merization of systems having half-cure times above 0,2 s.
Fig. 2 shows typical kinetic curves recorded for various acrylic resins. The degree of conver-
sion is directly related to the decrease of the IR absorbance, and can be calculated from the
equation:
(ASlO)O - @SlO)f looyo
Degree of conversion =
(*SlO)O

where (A810)o and (ASlo)frepresent the absorbance at 810 cm-' of the sample before and aft:r
UV exposure during time t. From these kinetic curves, one can readily evaluate how maxiy
acrylic bonds have polymerized, and thus determine both the actual rate of polymerization at a
given time:

(where [MIo is the original concentration of acrylate double bonds ( = 5 mol . I-')) and tlie
precise amount of residual unsaturation in the photocured polymer, residual unsaturation =
[ ( ~ s l o ) f ~ ( ~ s l ox
) o 1cm%.
l
The RTIR method is also suitable for following in real time the kinetics of polymerizatio is
that proceed more rapidly, in particular those induced by very intense UV exposure or by lasx
2384 C. Decker, K. Moussa

beams. The time scale is then shifted to the millisecond range. The only modification that must
be brought to the analysis device is to replace the recorder by a memory oscilloscope. The
polymerization kinetic curve will then appear directly on the screen, thus allowing both the rate
of polymerization and the induction period to be determined instantly. It should yet be
mentioned that in a routine IR spectrophotometer the response of the detector is in the order of
30 ms, thus making it possible to follow only polymerizations with half-cure times of 50 ms or
more; this is still within the range of most of the systems used in today's UV curing applications.
As compared to the other techniques of real-time investigation, RTIR spectroscopy thus permits
to push down by a 50-fold factor the lower limit of time detection.

Kinetics of the photopolymerization

The polymerization of the polyurethane-diacrylate investigated appears to proceed


rapidly upon UV exposure, even in the presence of air, leading to a highly crosslinked
and insoluble material within less than one second. Both the rate of polymerization
and the ultimate degree of conversion were found to be strongly dependent on the
type of monomer used as diluent (Fig. 2), the most efficient being the oxazolidone
monoacrylate newly developed by SNPE. The reaction kinetics follows a
characteristic S shape profile due to two major factors: (i) the initial induction period
is resulting from the well-known inhibition effect of O2 on those radical-induced
polymerizations; it disappeared completely for experiments carried out in vacuo or in
N,; (ii) the progressive slowing down observed at degrees of conversion above 30% is
the direct consequence of the network formation and the subsequent gelification
which reduces the segmental mobility of the growing polymer chains and of the
unreacted double bonds.

Rate of polymerization
From the slope of the RTIR kinetic curves it is easy to evaluate the rate of polymeri-
zation (R,) at any moment of the reaction (Tab. 1). As shown by Fig. 3, R , reaches

$100 -0 7 Fig. 2. Photopolymeriza-


.-
C 5 tion profiles recorded by
C E RTIR spectroscopy for a
- 0.2 00 polyurethane-diacr ylate
exposed to UV radiation in
0 the presence of a photoini-
- 0,L <
u
C

0
tiator ( 5 % of Irgacure 651)
and various acrylic mono-
u)
n
mers. (1): Ethyldiethylene-
Q glycol monoacrylate
- 0.6 (EDGA), (2): EDGA +
TPGDA, (3): tripropylene-
glycol diacrylate (TPGDA),
- 0.8 (4):oxazolidone
monoacrylate (CL 959).
Light intensity: 4 . lo-'
2.0 einstein . s - ' . cm-,
Irradiation time in s
A new method for monitoring ultra-fast photopolymerizations . . . 2315

Tab. 1. Results of photopolymerization of a polyurethane-diacrylate.Wt. ratio Actilane 201


monomer = 50:50; Irgacure 651 content: 5 wt.-%, irradiance: 4 . einstein. s - ' . cm-' ( 1
einstein = 1 mol of photons)

Monomera)

Monoacrylate:
EDGA
CL 959
Diacrylate:
TPGDA
HDDA
Triacrylate:
TMPTA 73 13 - 35

a) EDGA: ethyldiethyleneglycol monoacrylate, CL 959: oxazolidone monoacrylate, TPGDP.:


tripropylene glycol diacrylate, HDDA: hexanediol diacrylate, TMPTA: trimethylolpropar e
triacrylate.
b, Maximum rate of polymerization.
Photosensitivity = energy needed for 50% conversion.
dl Residual unsaturation in tack-free UV-cured film.

its maximum value (= 8 mole 1-' . s-I for the most reactive system) at a degree o f
conversion of 25% for the 3 systems investigated, once O2 inhibition has been
overcome and gelification has not yet slowed down the polymerization rate. Highcr
R, values, up to lo3 mol 1-' * s - ' , were obtained with such multiacrylic monomei's
by simply increasing the intensity of the UV source4)or by using powerful .
It should be mentioned that, like for DSC, such R, profiles can be directly recorded
by taking the first derivative of the absorbance versus time kinetic curves. The overall
polymerization quantum yield, @,, was calculated from the ratio of R , to the

Fig. 3. Variation of the rate


of photopolymerization R , of
a polyurethane-diacrylate with 6-
the degree of conversion.
Influence of the monomer
used as diluent. (1):
Ethyldiethyleneglycol
monoacrylate (EDGA), (2):
tripropyleneglycol diacrylate
(TPGDA), (3): EDGA +
TPGDA, (4): oxazolidone
monoacrylate (CL 959). Light
intensity: 4 .
einstein . s - I . cm-2
OO 25 50 75 1c 0
Conversion in %
2386 C. Decker, K . Moussa

absorbed light intensity and found to be in the order of lo3 double bonds polymerized
per photon absorbed, which clearly demonstrates the efficiency of the chain reaction,
even in the presence of air.
In Tab. 1 we also report the values of the photosensitivity ( S ) defined as the
amount of absorbed energy required to polymerize half of the reactive double bonds.
This parameter is most valuable for practical applications since it integrates both the
rate of polymerization and the induction period, which can be of some importance, in
particular for monoacrylates (Fig. 2). S values appear to be quite low for these multi-
acrylate resins ( < 1 mJ.cm-2), thus reflecting their high reactivity under UV
exposure.
In nitrogen atmosphere the photopolymerization was found to proceed 10 times
faster than in air4),with a distinct R,-conversion profile (Fig. 3); high R, values are
reached from the very beginnng of the UV exposure, while a similar drop as in air is
observed at conversions above 30%. once gelification occurs extensively. From the
measured values of q3p (= lo4 molecules per photon) and of the initiation quantum
yieldzo)(= 0,l radicals per photon), one can estimate the kinetic chain length. Despite
the high rates of initation used radicals * 1-' * s-I), it was found to reach very
large values, each initiating radical polymerizing up to 100O00 acrylic double bonds.

Residual unsaturation (7)

One of the unique advantages of the IR technique is to permit an instant and


precise evaluation of the amount of unreacted acrylic double bonds which remain in
the polymer network. Its value is highly dependent on the monomer functionality.
The polyurethane-diacrylate film, UV-cured up to the tack-free point, was found to
contain 35% residual unsaturation when using a triacrylate diluent, as compared to
10 to 15% with a diacrylate and only 2% with a monoacrylate (Tab. 1).
It should be emphasized that RTIR spectroscopy proved also very valuable to
determine precisely both the rate of polymerization and the amount of residual
unsaturation in acrylic resins where the initial double bond content was unknown.
Indeed, from the measured value of the acrylate molar absorptivity at 810 cm-' ( E =
73 1 * mol-' * cm-'), one can determine through the IR spectrum the acrylate concen-
tration at any exposure time, and thus evaluate both R, and 7 .
An additional advantage of RTIR spectroscopy is that the coating obtained after
UV exposure and IR analysis can be further examined by various techniques for
determination of hardness, gloss, solvent resistance, adhesion, scratching, and
abrasion resistance, etc. It should be noticed that these types of measurements cannot
be carried out on the samples analysed by some of the other methods of kinetic
investigation like dilatometry, calorimetry, or laser nephelometry and interferometry.

Effect of various parameters on the polymerization kinetics


RTIR spectroscopy has revealed as a powerful technique to evaluate rapidly and
quantitatively the efficiency of new photoinitiators, the reactivity of functionalized
A new method for monitoring ultra-fast photopolymerizations . . . 23K7

monomers and oligomers as well as to study the effect on R, of such important


factors as the thickness of the resin film, the light intensity, and the concentration of
02.

Influence of the photoinitiator


The most important constituent in a UV curable formulation is certainly the photo-
initiator, since even the most reactive acrylate monomers hardly polymerize when
exposed in pure form to UV light. It also governs the depth of penetration by radia-
tion, and thus the depth of cure, which is directly related to the concentration clf
photoinitiator. RTIR spectroscopy revealed as an excellent tool for evaluating
precisely, within a few seconds, the initiation efficiency of a given compound. As an
example, Fig. 4 shows the kinetic curves recorded upon UV exposure of a polyw-
ethane-diacrylate resin containing different types of photoinitiators. It permits an
accurate comparison of the performance of these compounds with regard to the trce
rate of polymerization R,, the photosensitivity S,and the amount of residual unsatw
ration r which are all determined within a single experiment.

.-C
C
Fig. 4. Influence of the 0

photoinitiator ( 5 % ) on the
>
photopolymerization
profiles of a polyurethane-
diacrylate + Acticryl CL
959 resin. (1): Irgacure
184; (2): Irgacure 651; (3):
Quantacure PDO; (4):
Irgacure 500; ( 5 ) :
Benzophenone +
N-methyldiethanolamine
(N-methyl-2,2 '-iminodi-
ethanol)

Irradiation time in 5

Influence of the functionalized oligorner


As shown previously (Fig. 1 and Tab. l ) , the type of monomer used as reacthe
diluent affects greatly both R, and r. The chemical structure of the functionalized
oligomer has a less pronounced effect on R,, but it acts mostly on the ultimate degree
of conversion, i.e. on the residual unsaturation content in the cured polymer. This s
clearly apparent from Fig. 5 which shows the polymerization curves recorded for a
polyester-acrylate and a polyurethane-acrylate exposed to UV light, the reactive
diluent and photoinitiator common to these two systems being the Acticryl CL 959
and the Irgacure 651, respectively. The polyurethane-acrylate yields an elastomeric
network where the active species retain sufficient mobility to make react almost all the
2388 C. Decker, K . Moussa

100
.-C
c
0
.-
Ll 75
>
C
0
Fig. 5. Photopolymerization
V profiles recorded by RTIR
spectroscopy for various resin
50
formulations (1): Polyur-
ethane-acrylate + Acticryl CL
959, (2): polyester-acrylate +
25 Acticryl CL 959, (3):
polyurethane-acrylate +
HDDA, (4): epoxy-acrylate +
HDDA. Photoinitiator:
Irgacure 651 (5%)
0.5 1.0 1.5 1

Irradiation time in s

acrylic double bonds. By contrast, in the polyester-acrylate the crosslinking polymeri-


zation leads to a stiff and vitreous polymer network where mobility restrictions make
the reaction stop at about 70% conversion. A similar behaviour was observed with
epoxy-acrylates, the slowing down being even more pronounced when a diacrylate
monomer was employed as reactive diluent (Fig. 5). The use of a difunctional
monomer was also shown to drastically reduce the ultimate degree of polymerization
in polyurethane systems (Fig. 5 ) .

Effect of film thickness


Fig. 6 shows the polymerization curves obtained by varying the thickness of the
resin film coated onto the NaCl plate, for a polyurethane-acrylate (47,5%) + Acticryl
CL 959 (47.5%) + Irgacure 651 (5%) system (formulation l), irradiated in the
presence of air. As the film thickness decreased to a few pm, the polymerization
efficiency dropped concomittantly. Such a drastic effect was not observed when

$75-
.-
C

C
.-
z50- Fig. 6. Effect of the film
>
thickness on the
0

ws,
u photopolymerization profiles
recorded by RTIR
25 - spectroscopy for a
polyurethane-acrylate
(formulation 1); I =
'
1,2 . IO-* einstein . s - . cm-2
OO 0.5 1.0 1.5 2.0
Irradiation time in s
A new method for monitoring ultra-fast photopolymerizations ... 23119

operating in a nitrogen-saturated atmosphere. It must therefore result from a moi'e


pronounced 0, inhibition in thin films due to a faster replacement of the oxygen
consumed by atmospheric 0, that diffuses through the surface2'). The observed
variation of R, with the film thickness (Fig. 7) appeous to be strongly dependent on
the light-intensity (I), sharpening the profile as Z increases, because of a shorter
exposure time during which O2 diffusion occurs.

q
I
-
a 3
E
\
_ - _ - - - -

Air
- N2

en

Fig. 7. Dependence of the


rate of polymerization R , on
the film thickness; (-):
air; ( - - -): N, -
30
Film thickness in pm

It should be mentioned here that one of the few limitations of RTIR spectroscopy is
that the sample must be less than 50 pm thick in order to avoid the saturation of the
IR absorbance. Fortunately, this happens to fall well within the thickness range of
essentially all the photocurable systems used today in the coating industry.

Effect of light intensity


Photopolymerizations of the polyurethane-acrylate (formulation 1) were also
performed at various light intensities (Fig. 8). It appeared that the usual kinetic law,
R, - Pt5, is no more obeyed with those multifunctional monomers and that lip
100 I
I

Fig. 8. Influence of the light


intensity on the
photopolymerization profiles
of a polyurethane-acrylate 25
(formulation 1). Light
intensity: (1): 1 , 2 . lo-'; (2):
1,7. lo-'; (3): 3 . lo-'; (4):
4 . 10-8 einstein.s-' .ern-'
0.5 1.0 1.5 20
Irradiation time in s
2390 C. Decker, K. Moussa

follows a close to first-order relationship with the light intensity (Fig. 9). This can
been explained on the basis of a unimolecular termination process which competes
with the usual bimolecular interaction of polymer radicals. The most likely first-order
termination appears to be a radical occlusion10~2z): as polymerization of these multi-
acrylates proceeds, a tight network is building up, thus sharply decreasing the chain
mobility, so that the polymer radicals will finally stop growing and remain trapped in
the matrix. ESR investigationsz3)and post-effect experiments) at moderate tempera-
ture ( Q 80 C) have recently confirmed the presence of long-living radicals in UV-
cured multiacrylates.

Fig. 9. Dependence of the rate of


polymerization R , on the light
intensity for a polyurethane-acrylate
(formulation 1 ) exposed to UV
radiation in the presence of air

108.Light intensity in
einstein.s-.cm-*

Post-polymerization
One of the basic questions that arises when studying fast-proceeding photopoly-
merizations is whether the reaction takes place only during the short exposure or also
afterwards in the dark. The RTIR spectroscopy proved well suited to address this
problem since it permits to follow continuously the post-effect after the light has been
switched off. For the polyurethane-acrylate (formulation l), polymerization was
found to continue to develop essentially during the first 2 s of the dark period
(Fig. lo), no further polymerization being observed 5 s after the end of the irradia-
tion. This result explains and quantifies an interesting observation that we made while
investigating the laser-induced polymerization of multiacrylic monomers in the
presence of airz4).A single laser shot, 8 ns wide, was shown to be able to polymerize
up to 70% of the reactive double bonds, thus implying that the polymerization was
taking place mainly in the dark, after the laser exposure. This result has recently been
corroborated by the remarkable work of Hoyle et aL2) on the effect of the pulse repe-
tition rate in the laser-initiated polymerization of acrylic monomers. Post-polymeri-
zation was found to develop extensively during a few seconds in the dark, in good
agreement with our results, leading to high-molecular-weight polymers. By contrast,
when a dark period of only 25 ms was imposed between two successive laser pulses,
A new method for monitoring ultra-fast photopolymerizations . . . 23'U

Fig. 10. Kinetics of the post-


polymerization in the dark
after UV exposure of a
polyuret hane-acrylate
(formulation 1) in the presence
of air. ( 0 ) :UV irradiation,
( - - -): dark polymerization,
(A): UV irradiation + 2 s
post-effect
1
Irradiation time in !i

the polymerization hardly occurred, with formation of very low-molecular-weight


compounds.
Such an important post-effect stresses the limitations of the conventional IR analy-
sis and the DSC or dilatometry measurements which, with their time delays of a few
seconds, give access only to the total UV + dark polymerization. By contrast, the
RTIR analysis makes it possible to differentiate easily those two processes and
evaluate precisely their relative importance. Fig. 11 shows that in the polyurethane --
monoacrylate resin the dark polymerization develops intensely in the early stages c f
the UV curing, accounting for up to 60% of the total polymer formed, while its cor -
tribution decreases steadily at conversions above 30% when gelification and vitrifics -
tion occur. As expected, this post-effect was slightly less pronounced when using a
diacrylate like TPGDA as the reactive diluent, which generates a highly crosslinke
polymer network with less chain mobility (Fig. 11).
The brief but still important post-polymerization observed in these UV-cure i
acrylic networks is likely to result primarily from the further growing of the polymer
radicals, since no new chains can be started in the absence of light, except for somc

Fig. 1 1. Relative importance of the


post-polymerization as a function of
the degree of conversion for a
polyurethane-acrylate irradiated in the
presence of air. Monomer (50%):
oxazolidone monoacrylate (CL 959)
( ) and tripropyleneglycol diacrylate
(TPGDA) (A)

Conversion in %
2392 C. Decker, K. Moussa

transfer reactions. In the dark, the polymer chains will stop growing because of the
usual bimolecular interactions of radicals, the trapping of radicals in the polymer
matrix, and their scavenging by oxygen. Further experiments are in progress to
evaluate the importance of those processes, in particular the latter, which is expected
to play a predominant role owing to the high reactivity of O2 toward alkyl radicals.
RTIR spectroscopy revealed also appropriate to study the kinetics of photocationic
polymerizations, and in particular the dark reaction, which develops here much more
effectively than in radically induced polymerizations, because of the low reactivity of
oxygen toward the propagating ionic species. The fast photopolymerization and post-
cure of epoxy and vinyl monomer could thus be followed in real-time by monitoring
the decrease of the IR absorbance at 790 and 1635 cm-' , respectively.

Conclusion
The real-time infra-red (RTIR) spectroscopy, presented here for the first time,
appears as a remarkable technique for following quantitatively and continuously
light-induced polymerizations that develop within a fraction of a second. It has been
applied to investigate the cure kinetics of the highly reactive multiacrylic resins which
are now commonly used in the coating industry. One of the unique advantages of this
method is that conversion versus time curves can be directly recorded, thus allowing
to determine instantly, at any moment, both the true rate of polymerization and the
amount of residual unsaturation in the cured polymer. The effect of the various con-
stituents of the resin on the photoreactivity can be rapidly evaluated by this technique
which appears as an ideal tool for assessing the efficiency of new photoinitiators and
monomers. Owing to its short response time, RTIR spectroscopy can be used to
monitor in real time polymerizations carried out under intense illumination, and thus
to operate in the presence of air under similar conditions as in industrial UV curing
operations. This analytical method proved also well suited to study the dark poly-
merization that occurs just after the end of the UV exposure and which was found to
represent a significant part of the overall process.
The distinct characteristics of RTIR spectroscopy become clearly apparent when
one compares the performance of the various kinetic methods commonly used today
to study photopolymerizations (Tab. 2); it turns out to be the only one that fulfils all
the different requirements. Besides its accuracy, reliability, and sensitivity, this
method of real-time investigation offers many advantages over the other techniques,
in particular, its versatility, easiness to use, immediate reading, and low cost, since it
requires no special equipment, apart from a routine IR spectrophotometer. It can be
applied to any fast-reacting polymerizable system, provided it exhibits a distinct IR
absorption specific of the reactive double bond, which is most often the case.
Furthermore, RTIR spectroscopy is not restricted to UV initiation since its field of
application can be extended to other types of radiation, like lasers, microwaves,
plasma, or electron beams, which are now increasingly used in various industrial
sectors for inducing instant polymerization processes.
Tab. 2. Performance analysis of various methods used for kinetic investigation of light-induced polymerizations 5'
09
E
*
Characteristics of Real-time Response time High intensity Rate ( R , ) Unsaturation Post-polyme- Properties
analytical methods analysis in s operation evaluation measurements rization evaluation
monitoring
-?E
IR s p e c t r o ~ c o p y ~ ~ ~ ) no Yes
13C NMR spectroscopy5) no yes
Photoacoustic spectroscopy6*') no no
Calorimetry ( D S C ) ~lo)~ Yes no
Dilatometry". I*) yes no
IR r a d i ~ m e t r y ' ~ ) Yes no
Laser-nephelometry 14p ' Yes no
Laser-interferometryj6) Yes no
RTIR spectroscopy yes yes

t4
w
B
2394 C. Decker, K. Moussa

) C. Decker, J. Coat. Technol. 59 (11751).97 (1987)


2, C. G. Roffey, Photopolyrnerziation of Surface Coatings, Wiley Interscience, Chichester
1982
) G. L. Collins, J. R. Constanza, J. Coat. Technol. 51 (n648), 57 (1979)
4, C. Decker, T. Bendaikha, Eur. Polym. J. 20, 753 (1984)
) V. Thalacker, T. Boettcher, Radcure Conference, Basel 1985
6, R. D. Small, J. A. Ors, B. S. Royce, ACS Symp. Ser. 242, 325 (1984)
) M. S. Salim, R. Cundall, A. Davies, Y. Dandikar, M. Slifkin, Radcure Conference, Basel
1985, FC 85-422
*) G. R. Tryson, A. R. Schultz, J. Polym. Sci., Polym. Chem. Ed. 17, 2059 (1979)
9, D. D. Le, J. Radiat. Curing 12 (2), 2 (1985)
lo) J. G. Kloosterboer, G. Lijten, Polymer 28, 1149 (1987)
D. R. Pemberton, A. F. Johnson, Polymer 25, 529 (1984)
12) A. K. Davies, R. B. Cundall, N. J. Bate, L. A. Simpson, J. Radiat. Curing 14 (2), 22 (1987)
3 G. B. Tanny, A. Lubelsky, 2. Rav-Noy, E. Shchon, Radcure Conference, Basel 1985, FC
85-440
14) C. Decker, M. Fizet, Makromol. Chem., Rapid, Commun. 1, 637 (1 980)
Is) M. Fizet, C. Decker, J. Faure, Eur. Polym. J. 21, 427 (1985)
16) C. Decker, in Radiation Curing of Polymers, ed. by D. Randell, Royal Society
Chemistry, London 1987, vol. 64, p. 16
17) C. Decker, J. Appl. Polym. Sci. 28, 97 (1983)
18) C. Decker, ACS Symp. Ser. 266, 207 (1984)
19) C. Decker, J. Coat. Technol. 56 (n713), 29 (1984)
m, A. Merlin, J. P. Fouassier, J. Chim. Phys. Phys.-Chim. Biol. 78, 267 (1981)
21) C. Decker, A. Jenkins, Macromolecules 18, 1241 (1985)
22) C. Decker, K. Moussa, J. Appl. Polym. Sci. 34, 1603 (1987)
23) C. Decker, K. Moussa, J. Polym. Sci., Polym. Chem. Ed. 25, 739 (1987)
24) C. Decker, J. Polym. Sci., Polym. Chem. Ed. 21, 2451 (1983)
2s) C. Hoyle, M. Trapp, C. Chang, Polym. Muter. Sci. Eng. 57, 579 (1987)

Вам также может понравиться