Sei sulla pagina 1di 45

Prog. Polym. Sci., Vol. 14, 717-761, 1989 0079-6700/89 $0.00 + .

50
Printed in Great Britain. All fights reserved. © 1989 Pergamon Press plc

FREE-RADICAL POLYMERIZATION: INHIBITION


AND RETARDATION

FERENC TiJD(~*'** and TAMARAFt3LDES-BEREZSNICH**


*EbtviJs Lordnd University, Department of Chemical Technology, H-1088 Budapest,
Mftzeum krt. 6--8, Hungary
**Central Research Institute for Chemistry of the Hungarian Academy of Sciences,
11-1025 Budapest, Pusztaszeri ftt 59-67, Hungary

CONTENTS
1. Introduction 717
2. Inhibition kinetics of radical polymerization 719
2.1. The kinetics of polymerization in the presence of additives 719
2.2. Kinetics of inhibited polymerization 720
2.3. The kinetics of inhibited polymerization in the presence of side reactions 726
2.3.1. The pseudo-unimolecular reaction of the inhibitor 726
2.3.2. The pseudo-bimolecular side reaction of the inhibitor 729
2.3.3. Simultaneous uni- and bimolecular side reactions of the inhibitor 730
2.3.4. The inhibition kinetics in the case of "chain regeneration" 731
2.3.5. The kinetics of inhibition reactions in the case of secondary
retardation 733
2.3.6. Determination of the inhibition parameters in the case of retardation 735
2.4. Some further comments 738
3. The practice of inhibition 740
3.1. Determination of the rate of initiation 740
3.2. Investigation of molecular inhibitors 741
3.2.1. Vinyl monomers as inhibitors 741
3.2.2. Aromatic hydrocarbons 742
3.2.3. Quinones 743
3.2.4. Aromatic nitro compounds 745
3.2.5. Nitroso compounds and nitrones 746
3.2.6. Other inhibitors operating by an addition mechanism 749
3.2.7. Inhibition by transfer mechanism 750
3.3. Interpretation of the stoichiometric anomalies 753
References 757

I. I N T R O D U C T I O N

The inhibition of radical polymerization is very important from both theoreti-


cal and practical points of view.
In practice, inhibitors are used for decreasing the rate of polymerization or for
stopping or preventing polymerization.
Uncontrolled polymerization is undesirable because it: (i) may cause loss
of monomer, (ii) produces polymer not of the desired molecular mass and

717
718 F. T1DDOS and T. FOLDES-BEREZSNICH

mass-distribution, spoiling the quality of the product, and (iii) during storage or
other conditions of insufficient heat transfer, it may accelerate autocatalytically
and, in extreme cases, can lead to a thermal explosion.
In order to prevent undesired polymerization, different inhibitors are used.
According to their function, these can be classified as follows:
(a) storage inhibitors,
(b) inhibitors which can be used during processing steps (e.g. distillation),
(c) inhibitors used to stop a polymerization quickly (e.g. in the case of any
disturbance), the so-called "short-stop" inhibitors,
(d) inhibitors used in regulating the deliberate polymerization process itself
(e.g. partial or complete elimination of the gel-effect, regulation of the molecular
mass, etc.),
(e) inhibitors used to stop a polymerization when the desired conversion has
been reached to increase the stability of the polymer (e.g. antioxidants).
For the functions given by the points (a) to (c), the aim is to stop the
polymerization essentially completely. Hence inhibitors of high reactivity are
applied in a relatively small concentration for these purposes. By contrast, the
functions outlined in (d) and (e) are usually best served by the application of
moderately or slightly reactive inhibitors (retarders), in relatively high con-
centration. In the choice of inhibitors in each given case, further secondary
aspects should also be considered. These include the stability and possible side
reactions of the inhibitor itself, the temperature regime used, the colour and
reaction products of the inhibitor, its price, its toxicity, the possibilities of
suppression or elimination of its side reactions etc.
Due to the extensive research on kinetics and mechanism carried out in the last
half century, radical polymerization is today one of the best known chemical reac-
tions. The application of diverse monomers enables wide variation of the electronic
structure, polarity and steric conditions of the propagating radicals, i.e. the
factors determining their reactivity. Using different additives, both the rate and
degree of polymerization can be regulated. Therefore, the inhibited radical poly-
merization is an excellent model reaction for the study of the kinetics and mechan-
ism of radical reactions on the one hand, and for the investigation of the relationship
between structure and radical reactivity of the reacting molecules on the other hand.
The systematic investigations we started as early as the 1950s (and which we
continue even now with a certain intensity) to clarify the stoichiometry of
inhibition, led to the surprising recognition that the experimental value of the
stoichiometric coefficient (/~+xp)is, in the case of molecular inhibitors, generally
smaller and in some systems much smaller than the theoretical value (Ptheor)
derivable from the mechanism of inhibition. These widely observable stoich-
iometric anomalies could be interpreted only by the hot radical theory. ~ This
theory, in turn, is applicable in many more areas than simply describing kinetics
of inhibition, as it can give a general interpretation of the active role of reaction
heat in chain reactions.
FREE-RADICAL POLYMERIZATION 719

2. I N H I B I T I O N KINETICS OF RADICAL POLYMERIZATION

2.1. The kinetics of polymerization in the presence of additives


Under the usual conditions of radical polymerization, three elementary reac-
tions need to be considered. These are initiation, chain propagation and the
bimolecular termination of the polymer radicals. In the case of initiated poly-
merization, these reactions can be described by the following equations:
X ~ 2fR" (k,) (1)
R + M~ R (k2) (2)
and
R" + R" ~ polymer (k4) (3)

where X denotes the initiator, M is the monomer, f is the radical generation


efficiency and R" is the propagating radical. In our calculations, the concen-
trations are given by the corresponding small letters. If an additive Z or Z X
capable of participating in radical reactions is added to the system, then the
additive may react with the propagating radical. This reaction proceeds either
by an addition
R" + Z ~ R-- Z (ks) (4)
or by a transfer mechanism
R + Z- X~ R - X+ Z (ks). (5)
Equations (4) and (5) produce, independently of their mechanism, secondary
radicals having chemical and electronic structure, polarity and stereochemical
properties more or less different from those of the propagating radicals.
This means that they are different in reactivity. The additive changes the
macrokinetics of the system, depending on the rate of the reaction between the
intermediate radical and the monomer molecule, according to the following
equation:
R - Z" or Z" + M ~ R- Z- M" or Z- M" (k6). (6)
This reaction produces radicals in which the odd electron is localized on the
part of the molecule originating from the monomer unit. Thus, the reactivity of
the radicals produced by this reaction is practically equal to that of the primary
propagating radicals (according to Bagdasaryan's terminology, this is "chain
regeneration"). According to the rate of eq. (6), two cases can be distinguished
from each other:
(1) If the rate ofeq. (6) is higher than or commensurate with the rate of chain
propagation, the polymerization rate does not change considerably. Depending
whether the additive reacts according to pathways (4) or (5), we have either
copolymerization or a chain transfer reaction.
720 F. TODOS and T. FOLDES-BEREZSNICH

(2) If the rate of eq. (6) is lower than that of eq. (2), or eq. (6) does not
take place at all, the concentration of intermediate radicals strongly increases
and these radicals have two ways to react with the polymer radicals (cross
recombination):
R ~ Z" or Z" + R" ~ R - Z - R or Z - R (k'5) (7)
or

2 R - Z ' o R - Z - Z - R (8)
and
2Z" --+ Z - Z. (8')
It should be noted that the rate of the latter two reaction is, for chemical
reasons, usually much lower than that of the cross recombination (7) and,
consequently, cannot compete with it. That is why, in practice, the termination
reaction of type (8) need not be taken into account. (Naturally, both the (7) and
(8) type reactions can proceed by a disproportionation mechanism. This has,
however, no importance for the further treatment.) Since the intermediate radicals
can, irrespective of the actual mechanism, participate only in the termination
reaction, the additive decreases the concentration of chain carrier radicals. In
comparison, this effect leads to a decrease in the rate of polymerization, i.e. an
inhibition effect can be observed macroscopically in this case. It is suitable to
distinguish between two cases of inhibition.
(a) If reactions (4) and (5) proceed with a moderate rate (more precisely, if
ks~k2 < 10), the effect is retardation, and
(b) If the above ratio is reversed (i.e. ks/k2 > 10), true inhibition occurs. It can
be seen that the differences between the two types of inhibition are not quali-
tative but quantitative. It is worth making such a distinction, because in the
kinetic treatment of the problem, different kinetic simplifications can be made
for the two cases. 2'3'4
Naturally, a number of actual reaction systems cannot be classified into these
two limiting categories. Among the experimentally studied cases not only the
pure limiting cases can be found, but also many diverse transition cases. A
common transition case between chain transfer and inhibition is, for example,
the so-called degenerative chain transfer, and the transition between copoly-
merization and inhibition can also be observed experimentally.

2.2. Kinetics o f inhibited polymerization


The kinetics of inhibited polymerization will be briefly discussed (compare
Fig. 1). It can be seen that during the inhibition period the polymerization
proceeds at a much lower rate. As the inhibitor is consumed, the rate of the
polymerization gradually increases. After the inhibitor is completely consumed,
the polymerization rate reaches the rate of the non-inhibited process. The
inhibition period increases with increasing inhibitor concentration. Considering
FREE-RADICAL POLYMERIZATION 721

6 -

r~ 4
o

o
iOo :
400 800 1200
t(min)
Ill
1600
I
2000

FIG. I. Kinetic curves for the AlBN-initiated, styrene-inhibited polymerization of


vinyl acetate at 50°C.

the elementary processes given earlier, the changes in concentrations of the


monomer (m), inhibitor (z), polymer radical (r) and intermediate radical (y) can
be expressed by the following differential equations 3'4'5
dm
dt = k2rm (9)
dz
dt = ksrz (10)
dr
d-t = W1 - o~ksrz - k'sry - k 4 r 2 (ll)
dy
d-t = otksrz - k'sry (12)
where Wi is the rate of initiation (W~ = 2kf~xo = const.) and ~t is a probability
factor, the explanation of which follows later (see Section 3.3). By solving
eqs (10) and (12), we obtain the following expression: 4

Y = (k'5/k5
-
l) 1- \Zo/ )"
(13)
In the steady state ( d r / d t = O, d y / d t = 0), the following approximate equa-
tions are obtained:
~k5
= , z (14)
Y k5
and
rst
r = (15)
[~o + J(1 + ~)1
722 F. TODOS and T. FOLDES-BEREZSNICH

where

(16)

z
(17)
-- 4x0
~tk 5
(18)
x/2k, fk,"
It can be clearly seen that the dimensionless quantity tp determines unequivoc-
ally the relative concentration of chain-carrying radicals. Therefore, ~0 can be
regarded as the inhibition p a r a m e t e r of the inhibited process, which contains not
only the parameters determined by the structure of the inhibitor (ks,/~) but also
the relevant parameters of the polymerization process itself (k4, W~) and thus it
gives the complex kinetic characterization of the inhibited process.
Substituting eqs (15) and (17) into eq. (10) describing the inhibitor consump-
tion, we arrive at the following form:
do ~0
dt = tp + x/(1 + ~02) ksGt' (19)

the integration of which gives


F(~o) - q~ + x/(1 + q~2) _ log 1/~0[1 + ~/(1 + tp2)] = F(tp0) - ksrs, t.
(20)
This equation describes the variation of ~o (which contains the inhibitor con-
centration) in time, in the form of an inverse function. This is generally the case
when the Bodenstein principle is applied, which leads to separable differential
equations.
It should be noted that the first numerical calculation of the problem was
carried out by BagdasaryanY In integrated form, the equation was derived by
three groups of researchers, simultaneously and independently, in apparently
very different forms. The first version was given by Burnett s'9 in a complicated
form. The Burnett paper itself stated that "this equation is extremely cumber-
some and unlikely to prove of much experimental use". But the complicated
form of the Burnett equation was due mainly to the particular notations used.
Independently of Burnett, this equation was derived by us in the above given
form in 1954 and published in 1956. '0 A third solution was published by
Bamford et al. H in 1957, also in somewhat different form: as the independent
variable, they used the relative rate of the inhibited polymerization. It was later
shown that all three solutions were mathematically correct and equivalent.
In the form given here, eq. (20) is'very comfortable and it can be used both
for theoretical consideration or for practical calculations without any difficulty.
The study of the limiting cases leads to important theoretical conclusions. At
FREE-RADICAL POLYMERIZATION 723

sufficiently high q~ values, F(tp) has asymptotic behavior:


F(~0) _~ 2q~. (21)
Where the latter equation is valid, eq. (20) approaches the following approximate
form:
2~o = 2~00 - ksrs, t (22)
That means that in the validity interval of eq. (22) (q~ > 3), the inhibitor
concentration linearly decreases in time. Consequently there exists a moment in
time, t = t~, where q~ = 0, i.e. z = 0.
2q~0 #z0
t, = ksr,, = " ~ " (23)
The latter is the length of the inhibition period, i.e. the period of time required for
the complete consumption of the inhibitor. In this expression, # = 20t is the
stoichiometric coefficient of the inhibitor, i.e. the number of polymer radicals
deactivated by one inhibitor molecule.
From eq. (23), the length of the inhibition period is proportional to the initial
inhibitor concentration and to the stoichiometric coefficient and inversely
proportional to the rate of initiation. In other words, the inhibitor is consumed
in a zero order reaction, i.e. the rate of its consumption is independent of its
concentration. This result can be directly obtained for strong inhibition, i.e. with
the assumption
k4 r2 ~ 2otksrz. (24)
From eqs 10-12:

dz = W, = W_,zO = const. (25)


dt /~ #
Eq. (23) enables the determination of W~ if # is known and conversely, # can be
determined if W~is known. Therefore, this equation is one of the most important
relationships of inhibition kinetics. The zero order of the inhibitor consumption
can be proved either by the study of the kinetics of inhibitor consumption or by
the investigation of the function t~ = f(zo).
The kinetics of inhibition is, however, more complicated if ~o < 3, i.e. near
ti. The above calculation, owing to the applied simplification (24), can be
regarded only as a first approximation. Near t~ the condition ¢p > 3 does n o t
hold any more and the decrease of the inhibitor concentration becomes slower
in comparison to the rate given by eq. 25. Thus the bimolecular termination
according to eq. (3) becomes more important and finally predominant.
In the detailed treatment ~z of the problem, it can be shown that by the usual
experimental procedure, i.e. by the extrapolation of the linear part of the
function log molto = f ( t ) , only the apparent length of the inhibition period (t')
can be obtained (see Fig. 2). Between the apparent and real lengths of the
724 F. TODOS and T. FOLDES-BEREZSNICH

3
IssslSS
2

-1

-2

-3

80 90 100 110 120


t

FIG. 2. The courses of basic theoretical functions characterizing inhibited


polymerization.

inhibition period, however, the following equation is valid:


t,--- t'/F*(cp0) (26)
where F*(~Oo) is a correcting function.

F*(cp0) = ~ ,( rp0 + ~ o - 1 + log ( 1 ,27,


where
~o = x/( 1 + tP02)• (28)
The inhibition parameter ~oo can be calculated from the initial relative rate of
the inhibited polymerization by the following equation:
¢0 = 1/2(W~' - W~I). (29)
The correlation between ti and t;, the F(W) function, the asymptotical
approximate functions of the latter and, finally, the log mo/m function can be
seen in Fig. 2 in the vicinity of tt. (The curves of Fig. 2 were calculated by the
following parameters: ¢90 = 10, ks/k2 -- 102 and ksrs, = 0.2. These values are
rather typical for inhibitors of medium activity.)
For weak inhibition, the length of inhibition period can be determined by the
following method) 3 Using the relationship
1
W~,/W = W~l = (30)
+ 4 o + ¢)
FREE-RADICAL POLYMERIZATION 725

0.5

0.1 0.2 0.3 0.4 0.5


log z°
z

FIG. 3. Determination of the ratio ks/k 2 by eq. (31) in the system styrene/AIBNat
50°C for the following inhibitors: (l) picryl chloride, (2) trinitroanisole, (3) trini-
trotoluene, (4) trinitroaniline.

which can be derived from eq. (15), q0 = q0(t) can be determined by the use of
W~j data obtained by differentiation of the log molto = f ( t ) plot. F r o m the ~o(t)
relationship, the F(q~) = g(t) function can be calculated. Then the best straight
line is fitted to the experimental points. The intersection point of this curve with
the time axis corresponds to the length of the inhibition period as can be seen
in Fig. 2.
The advantage of this method is that the gel effect which sometimes may take
place, cannot disturb the determination of ti.
The solution o f eqs (9) and (10) gives the following expression:

log m---2° = -k2


- log _z0 = _k2 log --.q~° (31)
m k5 z k5 tp
In this equation the value of Zo/Z can be expressed by the following approximate
formula:
Z0 ,.~ 1
z = (1 -- t/t,)" (32)
In the cases o f weaker inhibitors, when the accuracy of eq. (22) is insufficient,
the following, more accurate approximation can be used:
Zo 2q~o
-- ~- (33)
z • + 4(I + ~)
where
¢ = q~o{1 -- t/tt} - ¼q~o. (34)
The practical application of eq. 31 is illustrated by Fig. 3.
726 F. TODOS and T. F~JLDES-BEREZSNICH

Consequently, by the macrokinetic investigation of the polymerization, the


following values can be determined. From eq. (23), with the knowledge of the
rate of initiation, the stoichiometric coe~cient can be calculated. Conversely, if
we know the stoichiometry of the inhibition reaction, the rate of initiation can
be obtained. From eqs (31) and (32) (or (33)), the value of ks~k2, i.e. the relative
reactivity of the inhibitor, can be determined.

2.3. The kinetics of inhibited polymerization in the presence of side reactions


In practice, the kinetics of inhibition reactions often deviate more or less from
the simplest case treated above. These deviations suggest that other reactions,
in addition to those mentioned above, are involved in the inhibition mechanism.
This is not surprising. The active inhibitors are very reactive compounds.
Consequently, they can undergo other reactions too. These side reactions can be
divided into three groups:
(1) The inhibitor reacts directly with the monomer. This reaction is usually
first order in the monomer and first or second order in the inhibitor.
(2) The intermediate radical formed from the inhibitor in reaction (4) or (5)
has significant residual reactivity,and therefore can undergo a chain regener-
ation reaction according to eq. (6).
(3) The product formed in the inhibition reaction is not completely inert
toward radicals and is capable of undergoing further radical addition.
In these cases the kinetic analysis of the inhibition process becomes more
complex than for the simple inhibition. However, the differential equation
system can be solved applying the Bodenstein principle. Let us investigate these
cases in a more detailed fashion.

2.3.1. The pseudo-unimolecular reaction of the inhibitor - This case occurs when
stable free radicals are used as the inhibitor (e.g. in the system styrene/Banfield
radical/50°C 14:5) or aromatic nitroso compounds are used (e.g. in the system
MMA/aromatic nitroso compounds, 16'17:8 etc.)
(a) With the use of stable free radicals, the possibility of addition of this
radical to the double bond of the monomer should be taken into account. This
reaction is favored with very reactive monomers. The product is a propagating
radical:
Z" + M - - - , Z - M" - R" (k'). (35)
Consequently, a side reaction of initiation occurs in this case. The differential
equation system describing the concentrations of inhibitor and polymer radicals
is as follows:
dz
= kszr + k'mz ~- kszr + k'moz (36)
dt
dr
= 2k~fx + k ' m z - kszr ~- 2klfx + k ' m o z - kszr = 0. (37)
dt
FREE-RADICAL POLYMERIZATION 727

Considering that the consumption varies exponentially (x = x 0 exp ( - kt)),


upon combination of the above equations, we obtain
dz
= 2k~fxo et-ka) + 2k'moz. (38)
dt
(b) If inhibitors which are not themselves free radicals are used, the situation
is different from the above case in that the reaction between inhibitor and
monomer does not produce a propagating radical. In this case, the differential
equation system is as follows:
dz
= kszr + k'mz ~ kszr + k'moz (39)
dt
dr
= 2 k ~ f x - #kszr = 0. (40)
dt
By combination, we obtain

dz = 2klfxo e~-kt,) + k,moz. (41)


dt It
Equations (38) and (41) have the following common linear differential
equation form:
dz
= Ae -kt` + Bz. (42)
dt
The solution of the equation using initial conditions t = 0 and z = z0 is

( A ) ( A )e,_k,O" (43,
z = Zo + B - k------~ e ~ - B ° - B - - kl
In this case, the length of the inhibition period is

ti = go {1 + a Z:ot
~aal°g (44)

where
B - kl)
a = ~ x0. (45)

For case "a" (where the inhibitor is itselfa free radical) we get
2k'mo- kl
a = (46)
2klf
and for " b " (where the inhibitor is not a free radical) we get
#(k'mo - kl)
a = (47)
2klf
To illustrate the applicability of eq. (44), an example is shown in Fig. 4.
728 F. T O D O S and T. F O L D E S - B E R E Z S N I C H

300
a=O

a = 60 ,.,0
200

r-
~O ~
E

100 , . t "0~ ~

I I I
0 1 2 3
10 2 Zo/Xo

FIG. 4. The length of the inhibition period in the system methyl methacrylate/AIBN/
p-nitrosodimethylaniline at 50°C, plotted against the inhibitor concentration.

By the use o f numerical methods, it can be shown that eq. (44) can be
approximated by the following linear equation:

c\ -Xo- tJ = --
\ xo /
1 + 0.740 a z0_
Xo /
(48)

if zo/Xo < 10, with an error o f < 1%. Consequently, by plotting o f the experi-
mental data according to eq. (48), both constants o f eq. (44) can be directly

I I
2 4
ZO 10 2
XO

Fzc;. 5. Data for the system methyl methacrylate/AIBN/p-nitrosodimethylaniline at


50°C, plotted according to eq. (48).
FREE-RADICAL POLYMERIZATION 729

obtained. A linearized plot of the same experimental data is shown in Fig.


5.
It should be noted that in this case, k" can be directly determined by measur-
ing the rate of the reaction between the inhibitor and monomer, using a direct
polymerization method, t4-~7With the value of 2 k ~ f known, # can be determined
in two different ways for the case of " b " (inhibitor which is not itself a free
radical).

2.3.2. The pseudo-bimolecular side reaction o f the inhibitor - This is the case, for
example, when nitroso compounds are used as inhibitors in styrene polymeriz-
ation, tr-ls The differential equation system describing the concentrations of
inhibitor and polymer radicals is as follows:
dz
-- kszr + k"mz 2 ~- kszr + k"mo z2 (49)
dt
and
dr
d--t = 2 k l f x - #kszr ~ 2klfSc -- #kszr = 0 (50)

where, in order to eliminate the difficulties emerging in the integration, instead


of the actual initiator concentration, the logarithmic average of the initiator
concentration during the inhibition period is used, that is:
2klfYc = W1 = const. (51)
By combining these equations, one can derive a first order, homogeneous
quadratic differential equation:

dz = ~ + k,,moz2. (52)
dt #
The solution of the equation at initial conditions t = 0, z --- z0 is the following:

x/----
d arc tg ~/a" z = const - --# • t (53)
where
1
const - --7- arc tg ~/a" Zo (54)
~/a
and
k"#
a

From eq. (51) for the conditions t = t i and z = 0, the following expression
is obtained for the length of the inhibition period:

t, = (Wi. x/a) arc tg x/a • zo. (55)

To illustrate the applicability of eq. (55) an example is shown in Fig. 6.


730 F. TI~DOS and T. FOLDES-BEREZSNICH

6oo

/5>
200

0 1 2 3 4
102 Zo/'V~

Flo. 6. The length of the inhibition period in the system styrene/AIBN/p-


nitrosotoluene at 50°C, plotted against the inhibitor concentration.

From the above expression, an interesting conclusion can be drawn: the


inhibition period cannot be increased infinitely by increasing the inhibitor
concentration) 8 From eq. (53), the limiting value of the inhibition period is:

tioo = lim ti = ~ / ( ~ #J(zo). (56)


~]az 0 ~ oo z '~,/\ K' Wl ]

2.3.3. Simultaneous uni- and bimolecular side reactions o f the inhibitor - Both of
the above side reactions can occur simultaneously in some systems. This situ-
ation was observed experimentally in the case of styrene polymerization inhi-
bited by some of the nitroso compounds metioned above) 6-18 The differential
equation system describing the concentrations of the polymer radical and of the
inhibitor is the following:
dz
dt = kSzr + k'mz + k"mz 2 ~= kszr + k'mo z + k"mo z2 (57)

and
dr
d-~ = ~ - lzkszr = 0. (58)

Upon combination of both equations, we obtain the following first order


inhomogeneous quadratic differential equation:

dz = ~ + k'moz + k"moz 2 --- W~(1 + bz + az2). (59)


dt # #
The solution of the equation at initial conditions t = 0, z = z0 is the following:

I(z) = I(z0) - ~ t (60)


/t
FREE-RADICAL POLYMERIZATION 731

where the function I(z) can assume different forms depending on the actual
values of coefficients a and b:
(a) if the expression
A = 4a - b 2 (61)
is positive, i.e. A > 0, then
2 (2az+b)
I(z) = ~ arc tg x/A (62)

(b) if A < O, then


1 2az + b ~ x / ( - A )
I(z) = x / ( - A ) l°g 2az + b + x / ( - A ) (63)

(c) if A = 0, then
-2
I(z) - 2az + b" (64)

If k' = 0, that means b = 0, then eq. (62) reduces to eq. (53).


From eq. (60), considering the usual conditions (t = ti; z = 0), the length of
the inhibition period can be determined:

ti ~ g {I(z0) - I(O)}. (65)

It should be noted that in this case the length of inhibition period tends to a
limiting value with increasing inhibitor concentration, as in the preceding case.

2.3.4. The inhibition kinetics in the case o f "chain regeneration'" - If the radical
formed in the reaction between the polymer radical and the inhibitor has a
significant reactivity, it can participate in the chain regeneration reaction at a
lower or higher rate, depending on the reactivity of the monomer. The kinetics
of the process have been thoroughly investigated for the thermaP 9 and initiated 2°
polymerization of styrene. The kinetics of the process is briefly summarized as
follows. The differential equation system for this process is
dm
= k2rm + k6ym (66)
dt
dz
-d-~ = k5rz (67)

dr
d--t = W~ + k6ymo - ctksrz - k'sry = 0 (68)

dy
d--t = aksrz - k6moy - k'sry -- 0. (69)
732 F. TODOsand T. FOLDES-BEREZSNICH

From eq. (69), the concentration of intermediate radicals is


~tksrz
(70)
Y = k6mo + k'sr"
Two limiting cases of eq. (70) can be distinguished: (i) if
k'sr >> k6mo (71)
then

y = -- z (14)
k;
thus eq. (14) is returned. Consequently, we have here simple inhibition, without
side reactions. On the other hand, (ii) if
k'sr ~ k6mo (72)
then
ctksrz
Y = k6mo" (73)
This is identical to one of the basic copolymerization equations expressing the
equality of the rates of transformation of different radicals into each other
(equal rates of cross propagation). (In the treatment of copolymerizations,
usually other notations are used; further on ct = 1.) Thus, the general case (70)
is a transition between inhibition and copolymerization. Furthermore, if the
inhibitor undergoes a transfer reaction with the monomer according to eq. (5)
we have the transition between inhibition and chain transfer (according to
Bartlett's terminology, degradative chain transfer).
In our case, the differential equation of inhibitor consumption is
dz
(~){1 + x/(1 +c'zlYc)}. (74)
dt =
Integration of this equation with the initial conditions t = 0, z = z0 gives
2klfc"
I(z) = I(z0) - - t (75)
4/t
where

I(z) = {x/(1 + c ' z / X o ) } - (1 + l o g 2 ( x I(1 +c'z/xO)-c,z/xo 1)} (76)

and

Ct
4/on0ksk6
2klf " k'5 "
FREE-RADICAL POLYMERIZATION 733

8OO

6OO

e-
4OO

200

I I I t
0 0.1 0.2 0.3 0.4 0.5
Zo/X0

Fro. 7. The lengthof the inhibitionperiodin the systemstyrene/AIBN/trichloroquin-


one at 50°C, plotted against the inhibitorconcentration.

The length of the inhibition period is


4/~
t, = 2k~fc' I(z0). (77)

By a numerical investigation of the function I(z0) it can be shown that both


constants of eq. (77) can be determined directly by the following linear trans-
formation of the experimental data:

~ti,/
03 4c ) 78,
This equation can be used at the condition C'Zo/YC < 30. In Figs 7 and 8 practical
applications of eqs (77) and (78) respectively, are illustrated.

2.3.5. The kinetics o f inhibition reactions in the case o f secondary retardation -


It can often be observed that the rate of polymerization following the inhibition
period is somewhat smaller than the polymerization rate in the non-inhibited
process. In this case, the product of the reaction between the inhibitor and the
polymer radical is not quite inactive in radical reactions, though less reactive
than the original inhibitor itself.
Such kinetic behavior can be observed with the inhibition reactions of trini-
trobenzene derivatives 2~ (see Fig. 9) and of some aromatic hydrocarbons. 22
Depending on the reactivity of the product, two limiting cases can be
distinguished.
(a) The product has sufficient reactivity that itself can bring forth an inhi-
bition period (inhibition in two stages). This case was kinetically analyzed by
Bartlett and Kwart, 23'24 by using a differential method.
734 F. TODOS and T. FOLDES-BEREZSNICH

12

10

8
0

I I I I I
2 4 6 8 10 12
Zo/X o • 1 0 2

F1o. 8. Data for the system styrene/AIBN/2,6-dichloroquinone at 50°C, plotted


according to eq. (78).

(b) The product has only a slight reactivity, consequently only some decrease
in the steady rate takes place after the inhibition period.
We gave account earlier of our investigations concerning the (b)-type reac-
tion. 2~'22 For this case the kinetic analysis of the process is more complicated
than for simple inhibition. The differential equation system describing the
process, however, can be exactly solved in the steady state.
In a given case, the following implicit function exists between inhibitor
concentration and time:
F(0) = F(#0) - ksrs, t (79)

10

s / 2 4 5 6

o~
_9o
%

20O 1000
tlmin)

FIG. 9. Kinetic curves for the system styrene/AIBN/trinitrobenzene at 50°C.


FREE-RADICAL POLYMERIZATION 735

e-

0 1 2 3 4

Zo 102

FIG. 10. The apparent and real length of the inhibition period in the system styrene/
AIBN/trinitrobenzeneat 50°C.

where
F(0) = 0 + 0 + (0; log ( 0 - (0;)(0 + O)

- ~ ; l o g ( 0 -1 (0; ) { 1 + (0;0+~/(t~0)} (80)


and
o = (0 + (0; - (0' (81)
th' = ~/(1 + (0,2) (82)
0 = x/(1 + 02). (83)
Further notations are given in eq. (2). If (0; = 0, that is if no retardation takes
place, this equation is transformed into eq. (20). The correction function in
eq. (26) has in this case two parameters:

F*((0o,(0;) = 2(00 ((00 + 4 ) o ) - ((0; + ~b0) + (0; log ~o + •o}

If (0; = 0, the latter reduces into eq. (27). The practical use of eq. (84) is
illustrated by Fig. 10. The figure shows that the apparent length of the inhibition
period (t') is not a linear function of the variable Zo/2o. The calculated values of
t~, however, fall on a straight line.

2.3.6. Determination o f the inhibition parameters in the case o f retardation -


When retarders are used, the situation is somewhat different from that discussed
above. No inhibition period can be observed, therefore/~ cannot be determined
736 F. T~DOS and T. FOLDES-BEREZSNICH

from eq. 23, i.e. from the conversion-time curves. In this case, only the value of
fl can be determined from the polymerization rate data by eqs (29) and (18).
This constant, however, contains/~ and k5 in the form of a product.
For the compounds whose activity ranges between that of a retarder and that
o f a n inhibitor (1 < k s / k 2 < 20), the log mo/m = f ( t ) plot has some curvature.
Therefore the inhibition parameters can be determined for the latter with still
sufficient accuracy by the differential method treated in Section 2.2. However,
for the true retarders with still lower reactivity, this method cannot be used
either.
Nevertheless, a method can be developed by which the stoichiometric coef-
ficient can directly be determined from the known kinetics of inhibitor con-
sumption) 3 From the known value of g, the value of k 5 can also be calculated.
The definition of stoichiometric coefficient is:
number of deactivated radicals Ar
/~ = number of inhibitor molecules consumed = A---~" (85)
The retarder consumption must be measured directly by using any suitable
analytical method.
The next task is to determine the number of radicals deactivated by the
inhibitor in a given time interval (from 0 to t). Starting from the differential
equations of simple inhibition (eqs (9)-(12)), one can show that this value will
be given by the following integral:
Ar = 2 f~ k'sry" dt. (86)
However, from eqs (1 l) and (12) it follows that
2k'syr = W~ - k4r 2. (87)
After some minor algebraic transformations and integration we obtain
Ar = W~t(1 - W~,) (88)
and by the use of this equation,

/~ = Wit (1 2
- W~,). (89)
z0--z
This expression is very similar to eq. (23) for the length of the inhibition
period.
Concluding from the relatively low reactivity of retarders, the reactivity of
radicals formed from them is, in accordance with the general rules, somewhat
higher. Consequently, chain regeneration should be considered in the case of
retarders (eq. (6)). Kice 25 and Bagdasaryan 26 analyzed this process considering
bimolecular termination (eq. (8)) and they derived a three parameter, fourth
power equation.
As was mentioned in Section 2.1, the rate of reaction (8) is extremely low and
therefore it can usually be neglected. (Among the cases studied by us, this
FREE-RADICAL POLYMERIZATION 737

0.3
2

0.2 I

I I I I
0 0.2 0.4 0.6 0.8 1.0
Wr,,
FIG. I 1. Data for the system styrene/AIBN at 50°C, plotted according to eq. (94);
inhibitor: (1) 1,2,3,4-dibenzopyrene, (2) anthracene.

process never had to be considered.) On the other hand, if this reaction is


neglected, a simple two parameter equation is obtained, 27 the linearized form of
which is
W~I Zo 1
(P0 ~/x0 = f l W ~ l + y (90)
where the value of ~0 is calculated by eq. (29), and the intercept
k4k6 mo
-- k sk~ ~/x0 (9 l)
is proportional to the rate constant of chain regeneration. The practical use of
eq. (90) is illustrated by Fig. I I.
In this case there is another deviation - the stoichiometric coefficient cal-
culated by eq. (89) is not a true constant. It depends on the inhibitor concen-
tration ~z)- The real value of the stoichiometric coefficient is given by the
following equation:
# = #z(1 + fl~/W~,). (92)
In connection with this, the problem of the stoichiometric coefficient is worth
discussing generally. By minor rearrangement of eq. (23)

# = W, t~ (93)
z0
it is rather obvious that the real value of the stoichiometric coefficient can be
directly obtained from the experimental data by eq. (93) only when the ti/z o ratio
remains constant in the system investigated. This condition is fulfilled for simple
inhibition. If any side reactions occur, however, the proportionality between ti
738 F. T0DOS and T. FOLDES-BEREZSNICH

and z0 does not exist any more, as is clearly shown by eqs (44), (55), (65) and
(77). Thus, if a side reaction occurs, a concentration-dependent quantity (/~) is
obtained by calculation with eq. (93). For example, in the case of chain regener-
ation we get
4/~Xo . I(zo) (94)
~z ---- C' Z0

However, it can be pointed out that


lim /~ = lim (W~ti/Zo) = /z. (95)
z0~0 z0--*0

Researchers sometimes do not distinguish between these two quantities. This


results in erroneous data and theoretical confusion, as can be seen in Ref. 28.

2.4. Some further comments


As illustrated by the above-outlined kinetic analysis, the kinetics of inhibited
polymerization were clarified by the intensive research carried out in the 1950s
and 1960s. In the course of these investigations, the basic kinetic relationships
were determined which describe with sufficient accuracy the changes of the
concentrations of all components participating in the reactions. These relation-
ships were determined not only for simple inhibition, but for different side
reactions as well.
These side reactions complicate the kinetics of the process. However, they do
not cause theoretical difficulties. The differential equations describing the
process can be solved and the rates of side reactions can be measured by suitably
chosen methods. It should be noted that with the increase of the rates of side
reactions, the accuracy of the determined con ~tants gradually decreases.
The kinetic equation (13) describing the behavior of the intermediate radicals
R-Z" or Z" formed from an inhibitor was derived theoretically in our early
publicationJ 4 The correctness of this relationship could be verified later by
direct ESR measurements.29This method of ESR study of intermediate radicals
was extended by Janzen 3° in 1968 to other (non-polymerization) radical reac-
tions and it was given the name "spin trapping".
Nevertheless, in the last decade some papers were published3~-36dealing with
the kinetic analysis of inhibited polymerization. These publications (which,
except for Ref. 36, rarely refer to earlier results) generally do not ( and theoretic-
ally cannot) contain important new results: one kinetic problem can have only
one (and in this case yet unknown) solution. Occasional differences may be caused
only by the different (and generally rather badly chosen) notations of the
authors.
In addition to the analytical studies treated above, there is also a known
numerical study carried out by computer. 37'28 The authors integrate the dif-
ferential equation system of inhibited polymerization by the Runge-Kutta
FREE-RADICAL POLYMERIZATION 739

method. In order to diminish the computing time (in 1969), the authors assume
steady radical concentrations and apparently do not realize that for such cases
the analytical solutions are known. And what is even worse, the authors start
from an incomplete differential equation system (two terms are missing from the
equation of dy/dt). Thus, although their calculations do reflect something from
the inhibited polymerization, they are quantitatively defective and the con-
clusions drawn are erroneous.
In the question of applicability of the steady state, i.e. the Bodenstein principle,
uncertainty can often be found even among people dealing with reaction kinetics.
A practically useful criterion of this was given by Semenov in a paper written
in 1943 and rarely cited recently,38 in which the Bodenstein principle was
generalized. According to the generalized Bodenstein-Semenov principle, the
concentration (Rj) of the jth intermediate product of a chain reaction can be
considered steady state, i.e. the condition
drj
- - ~- 0 (96)
dt-
can be applied when the inequality
t > zj (97)
is fulfilled at the reaction time (t). In eq. (97) zj denotes the steady state lifetime
of R~. According to the definition of the theory of chain reactions, 39 the steady
state lifetime of propagating radicals in the polymerization is

r rs, (98)

In the presence of inhibitor, the steady state lifetime (%) of the radicals is

_ r = (rs,~ 1 _ z~, . (99)


~ w, \ w,/¢p + ,/(1 + ~o2) - q, + .,//(1 + ~o2)

It is obvious that, if inhibitor is present (q~ > 0), then


z= < "L, (100)
i.e. the polymerization becomes steady state sooner in the presence of inhibitor
than without it. Since, under the usual reaction conditions, %, = 0.1-10 sec, and
the region of usual inhibition times is t i = 1-20 hr, the inhibited polymerization
can be regarded as steady state practically from the moment t = 0. Thus, dis-
regarding the steady state approximation means a completely unjustified mathe-
matical complication which cannot give any physically rational result.
Therefore, the use of a computer seems to be reasonable only for the final
improvement of the numerical values of inhibition constants determined by
analytical solutions.
740 F. TODOS and T. FOLDES-BEREZSNICH

3. T H E P R A C T I C E O F I N H I B I T I O N

3.1. Determination of the rate of initiation


The above kinetic analysis indicates that one of the central problems of
inhibition kinetics is the determination of the rate of initiation. For this purpose
only those inhibitors can be used which participate with only one electron in the
reaction with the polymer radical. This enables one to assume that the inhibition
reaction consists of only one elementary step (e.g. recombination). Such
inhibitors are transition metal ions II (e.g. Fe 3+) and the stable free radicals.
Transition metal ions can be used, due to their limited solubility in organic
solvents, only in highly polar media. The most simple and accurate way to
determine the rate of initiation is by using stable free radicals. The stable free
radicals used for this purpose must fulfill the following requirements:
(1) They must be sufficiently stable under the reaction conditions used.
(2) They must not undergo any reaction except for that with the propagating
polymer radicals.
(3) The product of the reaction between the inhibitor and polymer radical
must not have any influence on the further course of the polymerization.
In the last years it was proved that in liquid phase polymerization the values
of kl and f a r e usually dependent on the composition of the system. The systems
investigated so far can be classified into two basic types:
(i) The rate constant of initiation in the mixture is additively composed of the rate
constants valid in the pure components, e.g. St/Bz (Bz = benzene), ~St/Do (Do =
dioxane),~ EA/St/Bz (EA = ethyl acrylate), 4°St/CCI4,4zMMA/DMF, 42St/MMA.43
Systems in which the rate constant o f initiation is constant regardless of
composition can be considered as limiting cases of type I, e.g. EA/Bz,44
AN/DMF, 45'42 MA/DMF, 46 BA/Bz,47 and MMA/Br-Bz. 48
(ii) The rate constant of initiation is not additively composed of the values
obtained in the pure components. The function 2knf = f ( z ) u (where Xu is the
molecular fraction of monomer in the mixture) can have different features:
(a) minimum curve, e.g. DEM/St (DEM = diethyl maleate), 49AN/BPA/DMF5°
(the total concentration of monomers was constant; in these experiments only
the ratio of monomers was varied in the mixture),
(b) maximum curve, e.g. St/DMF, 51 AN/MMA/DMF, s2
(c) saturation curve, e.g. AN/MA/DMF, s3 EA/DMF. 44
Although in some cases the solvent dependence of the rate constant of
initiation can be interpreted (e.g. by selective solvation44), the character of this
dependence usually cannot be "predicted". Therefore, the previously usual
practice of substituting "plausible" suppositions, instead of direct measurements
of the rate of initiation, cannot be accepted, because this procedure can cause
serious errors. The correct method is to determine the value of 2knfexperimentally
for the whole composition range and then calculate for each system with the
effective (actual) 2 k t f value belonging to the given concentrations.
FREE-RADICAL POLYMERIZATION 741

For the determination of the rate of initiation, the Banfield radicaP 4 has been
primarily used. Several aliphatic nitroxyl radicals were investigated as well. 5~
However, the deviations observed by the authors from the expected stoichiometry
(/~ = 1), are in contradiction to general experience and they can probably be con-
sidered due to experimental error. The picryl group of DPPH often causes strong
secondary retardation. 15'19 Therefore the 1,1-diphenyl-2-(2,6-dinitrophenyl)-
hydrazyl (DPDH) radical, which has the same stability, is more suitable than
DPPH for kinetic inhibition investigation. 15 For the study of polar monomers,
verdazyl derivatives or the Koelsch radical could be successfully used in some
cases. 43 Pseudo-unimolecular side reactions can often be observed in practice.
However, the ti = f(z0) relationship can be described by eq. (44), the par-
ameters of which can be directly determined by the use of eq. (48).

3.2. Investigation o f molecular inhibitors


If the electron structure of the inhibitor contains an even number of electrons,
i.e. it is a diamagnetic molecule, then, in radical reactions, it can furnish
diamagnetic molecules unable to undergo further radical reactions only if it
reacts with two free radicals containing unpaired electrons. This requires that
the inhibitor molecule be transformed to inert end products in at least two
(maybe four or six) consecutive elementary reactions. Therefore, the mechanism
of inhibition can be sometimes rather complex.
3.2.1. Vinyl monomers as inhibitors - The polymerization of a mixture of two
monomers can generally be considered as copolymerization. If, however, one of
the radicals formed in the reaction is much less reactive (e.g. owing to delocaliz-
ation) than the propagating radical, the process becomes inhibition from a
kinetic point of view. Such an example was presented in Fig. 1: styrene inhibits
the polymerization of vinyl acetate at a very low concentration, and ti depends
linearly on the styrene concentration, according to eq. (23). (The kinetic
criterion between copolymerization and inhibition is defined by eqs (70)-(73).)
The relative reactivities (ks~k2) and stoichiometric coefficients (#exp), as well as
the relative methyl affinities of the styrene derivatives investigated, are listed in
Table 1.
Owing to the release of considerable delocalization energy in the formation
of the resultant resonance-stabilized radical, the elementary inhibition reaction
is very strongly exothermic, AH s = - 1 7 5 k J m o l -l.ls From the temperature
dependence of the relative reactivity (between 20 and 60°C) in the system vinyl
acetate/AIBN/1,1-diphenylethylene, the following Arrhenius parameters can be
determined:
E5 -- Ez = -- 12.64kJmol -j (-3.02Kcalmo1-1) (101)
and
log A s / A ~ = -0.492. (102)
742 F. TODOSand T. FOLDES-BEREZSNICH

TABLE1. Stoichiometric coefficients and relative reactivities of monomers as inhibitors in vinyl


acetate polymerization(vinyl acetate/AIBN/substituted styrene at 50°C)

Inhibitor Stoichiometric Relative Localization Relative


coefficient, reactivity energy, methyl
gcxp of inhibitor (L), fl units affinity at
ks/k 2 65°C

Styrene 0.065 40.8 1.704 23.3


p-Cl-Styrene 0.095 52.8 - -
m-C1-Styrene 0.078 55.3 - -
p-CH3-Styrene 0.095 39.2 - -
m-CH3-Styrene 0.090 38.9 - -
or-Methyl Styrene 0.625 14.7 1.703 27.2
1,1-Diphenylethylene 1.77 33.9 1.514 46.8
1,2-Diphenylethylene 2.73* 2.158 2.80
1-Vinyl naphthalene 0.725 30.4 1.643 23.9
2-Vinyl naphthalene 0.183 48.5 1.684 -
9-Vinyl anthracene 0.975 38.0 1.405 12.9

*The compound acting as a retarder, only the #ks/k 2 product can be determined; this is listed in
the table.

3.2.2. A r o m a t i c h y d r o c a r b o n s - The higher aromatic ring systems are favored


model compounds for the investigation of radical reactivity, since these com-
pounds can easily be studied by quantum chemical methods.
Only a few of the above compounds exert any measurable retarder effect on
styrene polymerization. Usually chain regeneration can be observed. An inhibi-
tion period is observed only for tetracene, where the experimental data can be
described by the equations of simple inhibition. Since t~ is a strictly linear
function of z0, no chain regeneration takes place is this system. 59 The values of
inhibition constants are k s / k 2 = 156 (at 50°C) and # = 0.337 (30-60°C).
In the polymerization of methyl methacrylate, however, even tetracene exerts
only a retarder effect and a slight amount of chain regeneration can be observed.
In vinyl acetate polymerization the aromatic hydrocarbons were found to be
considerably stronger inhibitors. No chain regeneration took place, in accordance
with the much lower reactivity of the monomer toward radicals. The reactivity of
linearly annelated aromatic hydrocarbons as retarders or inhibitors increases
rapidly with the number of aromatic rings. 6° Angular annelation, however,
decreases the reactivity compared to the parent compound. 61
Since in primary radical attack a n-electron should be "torn off" the aromatic
n-electron system, reactivity should correlate theoretically with the atomic
localization energy (L). In fact, from their investigations of the addition of CCI~
to aromatic hydrocarbons, K o o y m a n and Farenhorst 62 proved in 1953 that log
(rate constant) varies linearly with atomic localization energy. In the course of
extensive studies made by Szwarc and coworkers on the addition of methyl
radicals 63 the energy of singlet-triplet excitation (Esx) was employed to
FREE-RADICAL POLYMERIZATION 743

characterize reactivity. The quantities L and Esx are not in a strict theoretical
relationship with one another, but a relatively close correlation can be empirically
demonstrated between them. Szwarc finally considered localization energies 64
too. Later on, Koutecky and coworkers showed all reactivity indexes of alter-
nating hydrocarbons to be in mutual correlation with each other. 65
The ks~k2 reactivities determined in vinyl acetate polymerization show strong
correlation with atomic localization energies. The relationship is, however, not
completely linear. 6~
The values for the experimental stoichiometric coefficients (~oxp)are consider-
ably smaller than those deduced theoretically (/~theor)"The quotient #exp/#th~or
may be considered approximately constant for a given monomer (0.21 for
styrene and 0.50 for vinyl acetate). 6° For explanation of such stoichiometric
anomalies, see Section 3.3.

3.2.3. Q u i n o n e s - Quinones are the most extensively studied inhibitors of


radical polymerization.
Szwarc, 66 investigating methyl affinities of numerous benzo- and naphtho-
quinones, has come to the conclusion that methyl radicals attack the C=C
double bond of quinones. Other very reactive radicals (e.g. phenyl radicals)
are likely to react by a similar mechanism. Less reactive radicals (having a
delocalized electron structure) however, as reported by Waters e t al. 67 and
Bevington e t al., 68 attack the oxygen atom of the quinone, resulting in mono-
and (mostly) diethers of hydroquinone as end products:

R" + O O = R--'--O O" ,, R ~ O OR. (103)


Investigations with labelled compounds indicated, however, that nz = 1.7-
1.85 quinone molecules were incorporated into an average polymer molecule
formed during the inhibition period# s That is, the actual mechanism of the
reaction is more complicated. Nevertheless, it is clear that in the elementary
inhibition reaction the quinonoid n-electron structure will be transformed to a
benzenoid, i.e. to an aromatic structure (contrary to the reaction taking place
with aromatic hydrocarbons, where the aromatic character of one or more rings
is lost).
According to Breitenbach's investigations,69 in the polymerization of styrene
and methyl methacrylate,7° log (reactivity) is a linear function of the redox
potential of quinones (E°), while Bartlett et al. 71 claim that this relationship is
non-linear.
Our very extensive studies suggested that, considering their reactivity in
styrene polymerization, the quinones can be divided into three groups:
(1) Benzoquinone and its non-halogen-substituted derivatives.72-74
(2) Halogen-substituted derivatives of benzoquinone)°
(3) Quinones with condensed ring systems.74
744 F. TI[IDOSand T. F~)LDES-BEREZSNICH

TABLE2. Stoichiometric coefficients (/z), relative reactivities (ks~k2) and redox potentials (E°) of
diverse quinone derivatives in the radical polymerization of styrene initiated by AIBN at 50°C

ks/k2 k~/k2* E ° (V)


exp. calc.

1. 1,4-Benzoquinone(BQ) 1.26 520 630 0.711


2. 2-Me-BQ !.48 227 264 0.653
3. 2,3-di-Me-BQ 0.70 110 83 0.588
4. 2,5-di-Me-BQ 1.05 89 95 0.604
5. 2,6-di-Me-BQ 0.89 154 136 0.607
6. Tri-Me-BQ 0.65 25 28 0.529
7. Tetra-Me-BQ 0.5 4.4 4.3 0.475
8. 2,5-Me,i-Pr-BQ 1.18 56 61 0.597
9. 2,5-di-i-Pr-BQ 1.56 32 31 0.595
10. 2,6-di-i-Pr-BQ 1.79 95 112 0.590
11. 2-t-Bu-BQ 1.62 235 170 0.636
12. 2-5-di-t-Bu-BQ 1.36 4.3 4.3 0.558
13. 2,6-di-t-Bu-BQ 1.86 61 55 0.506
14. 2-OMe-BQ 1.08 193 207 0.642
15. 2,5-di-OMe-BQ 1.00 15 16 0.476
16. 2,6-di-OMe-BE 0.91 50 67 0.530
17. 2,6-3i-Br-BQ 0.60 440 418 0.744

18. 1,4-Naphthoquinone 0.78 50 47 0.484


19. 3,8-Pyrenequinone 1.65 367 225 0.666
20. 4,4'-Diphenoquinone 0.73 3000 2700 0.954

21. 1,2-Naphthoquinone 1.15 605 688 0.576


22. 1,2-Anthraquinone 1.49 248 186 0.489
23. 1,2-Phenanthrenequinone 0.90 2200 2126 0.651
24. 9,10-Phenanthrenequinone 1.61 168 142 0.471
25. 5,6-Chrysenequinone 1.97 90 129 0.465

*k*5/k2 values of p-quinones (I-20) were calculated by the equation log ks/
k2 = 3.744E° - 0.1405 and of ortho-quinones (21-25) by the equation log ks~
k2 = 6.534E° - 0.9261. In case of substituted 1,4-benzoquinones, the sterie effect of the
substituent was taken into account by the techniques given in T, L. Sim~indiand F. Tiid6s, Eur.
Polym. J. 21, 865 (1985).

Earlier investigations72'73 supported Breitenbach's o b s e r v a t i o n that the value


o f log ks/k2 is basically d e t e r m i n e d by E °. Later investigations o f derivatives
c o n t a i n i n g b u l k y substituents at the 2,5- a n d 2,6- positions, however, showed
that the steric h i n d r a n c e caused by the substituents decreased the reactivity in
some cases by as m u c h as I-1.5 orders of magnitude. Simultaneous consideration
o f both o f these factors enables the reactivity to be described very accurately (see
T a b l e 2).
The halogen substituted derivatives form charge transfer complexes with
styrene. 2° The complex f o r m a t i o n considerably increases the reactivity. The
observed effect could be interpreted as due to the increase o f the t r a n s m i s s i o n
coefficient (x) o f the e l e m e n t a r y i n h i b i t i o n reaction. 75 (The m a j o r i t y o f these
reactions belong to the class o f so-called n o n - a d i a b a t i c reactions from a q u a n -
t u m chemical p o i n t o f view.)
FREE-RADICAL POLYMERIZATION 745

The reactivities of multi-ring quinones are influenced not only by redox


potential but also by the position of the C = O groups. The o-quinones are
definitely more reactive than the corresponding para derivatives, as can be seen
for the naphthoquinones in Table 2.
In vinyl acetate polymerization, 76the majority of quinones exhibit such a high
reactivity that the value of ks/k2 cannot be measured by the usual methods,
except for some quinones with E ° -,, 0.5V. Also the inhibition mechanism
seems to be more complicated. In some cases the attack of radicals along the
C = C double bond cannot be excluded either.
The stoichiometric coefficients are surprisingly low: #exp = 0.16-1.5, depend-
ing on the structure of the quinone. The data for duroquinone (ks/k2 = 83,
#exp = 0.72) sufficiently agree with Bartlett's measurements. 23 (It should be
noted that Bartlett's interpretation of the value of #e~p as equal to 1 cannot be
maintained any more.)
The stoichiometric coefficient values are anomalous in styrene polymerization
also 72-76
Yassin and his coworkers published several papers on the study of the
inhibition mechanism of the quinones. In their first publication 77 they argued
that substitution of the polymer chains into the quinone nucleus was the
predominant mechanism. In a later publication, 78 however, they suggested a
mechanism based on electron transfer from the polymeric radicals to the quin-
ones. At the same time, it would seem that the relative monomer and inhibitor
concentrations used (Zo/Xo = 0.67) drastically differed from the normal con-
ditions of kinetic measurement. Also, the method did not assure oxygen-free
conditions. (Chloranil can be used in high concentration as a dehydrogenating
agent for preparative purposes.) Thus, it is not surprising that the results of
these experiments were in contradiction to those of both Bevington's investi-
gations 68 on the incorporation of inhibitor and the ESR study 79 of intermediate
radicals.

3.2.4. Aromatic nitro compounds - The first kinetic inhibition investigations with
nitro compounds were carried out as early as the 1940s. 8°'81 More thorough
kinetic studies, however, were performed later by Bartlett and his coworkersY "24"82
They found that in the polymerization of vinyl acetate and of styrene, the
propagating radicals attack at the oxygen atom of the nitro group. Further
studies 21'83revealed that electron-accepting substituents increase the reactivity of
the nitro group and that the effect can be well described by the Hammett
equation (for styrene, Q = + 1.3). In styrene polymerization inhibited by sub-
stituted trinitrobenzenes the first nitro group can deactivate from 2.73 to 3.78,
propagating radicals. In vinyl acetate polymerization inhibited by substituted
mononitrobenzenes s4 the value of #expwas found to vary between 1.33 (p-F) and
2.85 (p-OH). Concluding from these data, the mechanism of inhibition has to
consist of four consecutive elementary reactions:
746 F. T1DDOS and T. F O L D E S - B E R E Z S N I C H

R" + NO2 -* R-O-N -O" ~ N=O + R - O - R (104)


I I I
Ar Ar Ar
and
R" + N=O ~ R - N - O _E.. R - N - O R . (105)
I I I
Ar Ar Ar
The existence of the reaction series (105) was directly verified by Waters et aL g5
The corresponding hydroxylamine derivative is produced in nearly quantitative
yield by the reaction of nitrosobenzene and the 2-cyano-2-propyl radical.
The polymerization of vinyl acetate proved especially suitable for the investi-
gation of the substituent effect in radical reactions, since no side reaction takes
place and the parameters can be calculated by the equations of simple inhibition
(eqs (23) and (31)). We investigated altogether 34 substituentss4 in both o- and
p-positions. The value of ks/kz varied between 3.82 (p-NH2) and 165.3
(p-SOzCl). Based on these results, we could prove that the representation of
Hammett constants according to the Taft equations
ap = 2o l + trR (106)
and
(7"m = O"1 -]- 0~0"R (107)
where al and trR are inductive and mesomeric components of the substituent
constants, can be used with high accuracy also for the description of radical
reactions.
In the polymerization of the considerably more reactive monomer styrene, the
mono- and dinitrobenzene derivatives exert only a retarder effect. This retar-
dation can be described quantitatively by eq. (90).27The k~ constants determined
this way, however, show an opposite relationship to the Hammett constants. A
O0 value of - 0 . 9 was found, in full agreement with the theory of radical
reactivity. As for aliphatic nitro compounds, a weak retarder effect of 2-fluoro-
2,2-dinitroethanol was observed in the polymerization of methyl methacrylate,s6
The kinetic data can be quantitatively described by eq. (90).

3.2.5. Nitroso compounds and nitrones - Although Foord s° had already observed
that nitrosobenzene was a strong inhibitor of styrene polymerization and the
inclination to radical addition of the strongly unsaturated N=O was definitely
proved by preparative studies,85systematic investigation of the aromatic nitroso
compounds was started only in the 1960s. Based on their effect on the inhibited
polymerization of styrene and of methyl methacrylate, the nitroso compounds
tested have to be divided into two groups:
(I) aromatic nitroso compounds,
(2) N-nitroso compounds.
FREE-RADICAL POLYMERIZATION 747

According to Ingold'ss7 preparative studies, aromatic nitroso compounds


readily react with the C=C double bonds of some vinyl monomers. This
side reaction can be observed with styrene too, but it is especially rapid with
methyl methacrylate. This direct reaction is pseudo-unimolecular for every
nitroso compound with methyl methacrylate. In styrene, however, it is unimol-
ecular only at very low concentrations. At higher concentrations it is pseudo-
bimolecular with respect to the nitroso compound, except for p-nitroso-aniline
and its N-substituted derivatives, where no side reaction occurs at all. Is It can
be proved photometrically that nitroso compounds are present in dissociated
form in methyl methacrylate, while in the apolar styrene they are mostly
dimerized, except for the derivatives of p-nitroso-aniline, which cannot dimer-
ize. Consequently, it is obvious that the "monomeric" form of the nitroso
compound reacts with methyl methacrylate while its cis-dimer reacts with
styrene:~6
o
Ar O .~
~N. J Ar~N~CH2
II • - I I ---A'--N----C"--CH2--N--A, (108)
Ar/ "~0 CH--~ 1 O" • O°
O
For 6-nitroso-m-cresol, p-nitrosodimethylaniline and p-nitrosodiphenylamine,
this side reaction could not be observed in the polymerization of methyl acryl-
ate 88 and of acrylonitrile.89 The ti = f(zo) relationships were found to be fully
linear.
The C-nitroso compounds have a reactivity so high that it cannot, ~7"~8"9°or
hardly can be measured88'89 dilatometrically. Only 6-nitroso-m-cresol has a
moderate reactivity both in the polymerization of acrylonitrile (ks/k2 = 255 at
50°C) and of methyl acrylate (55.5 at 50°C). This inhibitor reacts perhaps in its
tautomeric form, i.e. as toluquinone monoxime. Owing to their high reactivity,
aromatic nitroso compounds are par excellence suitable for practical inhibition
purposes. In the choice of an inhibitor, however, the above side reactions should
be considered.
Based on preparative studies,85 the expected value of ~h~or for nitroso com-
pounds is 2, in accordance with eq. (105). Experiments generally give a lower
value. The/~oxp values determined for the system styrene/AIBN are listed in
Table 3.
From Table 3 it can be seen that, independently of the specific substituents on
the amino group, the/~e~p values for different p-nitroso-aniline derivatives are
constant within the limits of experimental error (___5*/0). The substituents on the
aromatic ring, however, strongly affect the/~e~p values. For the p-Cl derivative,
for example, values as low as 1/6 of the theoretical can be measured. The value
of #~xpwas found to be independent of the temperature for both styrene and
methyl methacrylate.
748 F. TODOS and T. FOLDES-BEREZSNICH

TABLE 3. Stoichiometric coefficients for nitroso compounds in the polymerization of styrene at 50°C

No. Q, or the name of the T, °C tt


compound, resp.

1 p-NH 2 50 1.66
2 p-NHCH 3 50 1.60
3 p'-N(CH3) 2 50 1.67
4 p-N(CH3)2 60 1.54
5 p-N(C2Hs)2 50 1.56
6 p-NHPh 50 1.48
7 p-NPh2 50 1.55

8 -H 50 0.83
9 p-CH3 50 1.68
10 p-C2H5 50 1.27
11 o-C2H 5 50 1.70
12 p-OH 50 1.60
13 p-OCH 3 50 1.01
14 p-Cl 50 O.33

15 1-Nitroso-2-naphthol 50 1.39
16 2-Nitroso- 1-naphthol 50 1.26

17 N-Nit rosodiphenylamine 50 3.82


18 N-Nitrosodiphenylamine 60 3.58
19 N-Nitrosodiphenylamine 70 2.70

The value of/-~expfor N-nitrosodiphenylamine~'9~ considerably differs in nature


from those of C-nitroso compounds. For this compound, #expdepends on the
temperature, and its reactivity, in contrast to all other inhibitors, increases with
the temperature.* The observed differences can be interpreted by taking into
account the fact that the N - N bond is relatively weak. This means that at higher
temperatures the inhibitor may dissociate to the radicals Ph2N" and NO'. By
addition of a propagating radical, nitric oxide is transformed into a nitroso
compound which is then capable of reacting with two additional radicals. Thus,
~theor of these compounds is 4.
Aromatic aldonitrones were found 92 to be effective inhibitors of styrene
polymerization in studies which used m- and p-substituted c-aryl-N-phenyl-
nitrones. Primary radical addition results in a stable nitroxyl radical, just as in
the reaction ofnitroso compounds. For all investigated aldonitrones, the length
of the inhibition period is a non-linear function of the initial concentration of
the inhibitor. This means that an inhibitor-consuming side reaction must be

*Of course, the numbering of curves in Fig. 2 of Ref. 91 is bad, should be commuted!
FREE-RADICAL POLYMERIZATION 749

taken into account. According to the investigations of Huisgen and his coworkers,
a dipolar cyclo-addition producing an isoxazoline ring takes place. The reactivities
show a significant substituent effect92 obeying the Hammett-Taft equation with
Q = +0.86.
Due to the extreme stability of the nitroxyl radical, the radical reactions of
C-nitroso compounds and nitrones are very convenient for ESR studies. 1:6'29'94
Owing to the very high k5 values, the concentration of the intermediate radicals
(see eq. (13)) can, in these reactions, often reach 10-6-10 -5, in some cases even
10-4moldm -3. The ESR spectra, supplying direct evidence of the nature and
structure of intermediate radicals, help to determine unequivocally the mech-
anism of inhibition.

3.2.6. Other inhibitors operating by an addition mechanism - The early poly-


merization literature95 reported phenylacetylene to be a strong inhibitor of
styrene polymerization. It was claimed to decrease the rate of polymerization
and the molecular mass of the polymer formed when present in a concentration
above 0.01%. Later studies%'97do not support these early claims. Phenylacetyl-
ene decreases the rate of styrene polymerization by not more than 30% even in
a concentration of 0.5 mol dm -3, which is quite unusual for inhibitors. In the
polymerization of methyl methacrylate, the retarder effect is stronger. The
MMA polymerization rate decreases by a factor of 4-6 at the same inhibitor
concentration, depending on the substituent present in the p-position.
The effect of some 6-substituted dibenzofulvenes on the polymerization of
methyl methacrylate was studied by Kice,98 who found a retarding effect. In
addition, chain regeneration could be detected. We studied the inhibitory effect
of four dibenzofulvene derivatives99 on the polymerization of vinyl acetate,
which gives propagating radicals of much higher reactivity. The values of #e~p
(1.98-2.14) were in accordance with the expected/~or. Consequently, in these
systems, it is apparent that a simple two-step inhibition takes place. The only
exception is the/~xp value of 4.12 obtained for the 6-phenyl derivative. So it must
be assumed that the primary product of the inhibition produces two stable
radicals by dissociation, which deactivate two other propagating radicals. The
assumed mechanisms were proved by ESR study of the structures of inter-
mediate radicals. The reactivity decreases with increasing van der Waals radius
of the substituent(s) in position 6.
The inhibiting or retarding effects of aromatic azo compounds have been
observed for radical polymerization of several monomers, e.g. styrene, l°°'l°t
isoprene, ~°2 vinyl acetate, ~°3'1°4vinyl chloride ~°4 and methyl methacrylate)°4A°5
A study~°6was made on the inhibition by some 3,3'- and 4,4'-disubstituted azo-
benzenes of the polymerization of vinylacetate initiated by azo-isobutyronitrile
at 50°C. The inhibitory effects of these substances can be attributed to their
ability to engage in radical addition giving a less reactive hydrazyl type radical.
The mechanism of the inhibition has been established by ESR and kinetic
750 F. TODOS and T. FOLDES-BEREZSNICH

(stoichiometric) measurements. The value of ks Ik2 was determined for 9 sub-


stituents. The radical reactivity of the aromatic azo group was decreased by
electron-donating substituents and increased by electron acceptors. The sub-
stituent effect can be well interpreted by the Hammett equation, p = + 0.53.
The Pcxpvalues (1.17-2.17) are, having considered the limits of experimental
error ( _ 5-8%), very close to the theoretical.
A publication ~°7 reports the inhibiting effect of triphenyl formazan and its
derivatives in methyl methacrylate polymerization. The method used (DSC)
is, however, not suitable for the determination of precise kinetic data under
oxygen-free conditions.
Sugiyama et aL report that 1,1,4,4-tetraphenyl-2-tetrazene acts as a retarder
of the polymerization of methyl methacrylate.
It has been known since Staudinger and his coworkers' early investigations~°9
that molecular oxygen induces an inhibition period in the polymerization of
some monomers. In the polymerization Of 1,1-diphenylethylene1°9and of methyl
methacrylate,It° they could isolate the alternating copolymer formed. Investi-
gations carried out in the 1950shI-112furnished additional evidence. The kinetics
of the process were studied.114 Although oxygen is, due to its singlet ground state,
very reactive towards radicals, it cannot be used as a storage inhibitor because
of the limited amount present in a closed container which is nearly full. In
addition, at elevated temperatures, the polymeric peroxide formed decomposes,
giving radicals capable of initiating polymerization.

3.2.7. Inhibition by transfer mechanism - If the reactivity of the Z" radicals


formed in reaction (5) is low enough, they can undergo further reaction only by
cross termination (7) and the process should be regarded as inhibition (see
Section 2.1). The low reactivity of the Z" radical can also be expressed in terms
of the low dissociation energy of the Z - X bond (D[Z-X]). This condition is
fulfilled by rather many bonds. Of these, we are going to deal with the radical
exchange reactions of C-H, O-H and N-H bonds, since for these systematic
investigations in the literature can be found. If X=H, the relevant compounds
are generally known as antioxidants because in the oxidation chain reaction
they can react with the chain-carrying peroxyl radicals at a sufficient rate:

RO~ + H - Z ~ ROOH + Z" (109)

but, owing to the low reactivity of Z', they do not undergo a chain reaction, or
if so, only very slowly.

Z" + H - R ~ Z - H + R " (110)

Correspondingly, the retardation or inhibition of the oxidation reaction can be


observed in their presence, depending on the rate of the reaction (109). For a
very recent discussion of this topic, see Ref. 115.
FREE-RADICAL POLYMERIZATION 751

3.2.7.1. Effect ofphenois - It is known that the inhibition period brought about
by oxygen dissolved in styrene is lengthened substantially by the presence of
phenols ~16whereas in oxygen-free systems the rate of polymerization is unaffected
by, for example, hydroquinone. For a long time, phenols were believed to react
only with peroxy radicals 1~7and not with carbon radicals. This phenomenon was
explained by the transformation of phenols into quinones in the presence of
oxygen, followed by inhibition of radical polymerization by these powerful
inhibitors. 11s Most of these investigations were unfortunately performed with
styrene. The polystyrene radical is relatively unreactive and phenols of more
polar character are only slightly soluble in styrene. These two factors together
caused the effects of simple phenols in oxygen-free systems to be only at the limit
of detection.
Developments in polymerization techniques led to observation of the small
retarding effects of phenols in the absence of oxygenH9"~2°but the phenomenon
was attributed to chain transfer. The effect of phenols on the polymerization of
vinyl acetate has been studied by Bird and Russell ~2~who calculated the transfer
constants for some substituted phenols from degree of polymerization, as well
as from velocity data. 122
Very systematic investigations are reported in the paper, 123which contains the
inhibition kinetic parameters of 34 substituted phenol derivatives in the poly-
merization of vinyl acetate at 50°C. The reactivity data gave the best correlation
with the electrophilic substituent constants (a +) according to the Hammett
equation, where the spread of the data did not exceed +_30%:
logcfl = 0 . 3 4 - 1.52tr+ (lll)
where fl is the constant which can be calculated by eqs (17), 08) and (29), and
c is the reciprocal value of the number of equivalent reaction centers. 123 More
considerable negative deviation can be observed only for the tertiary butyl
substituents, which can be interpreted as steric hindrance. Also, the tr + values
o f - F , -Cl and -Br need minor correction.
The rate determining step in the reactions between phenols and polyvinyl
acetate radicals is hydrogen transfer from the phenols. The experimental evi-
dence for this claim includes (1) the very slight reactivity of the O-alkyl and
O-acyl derivatives and (2) the kinetic isotope effect. ~24 Since a rapid isotopic
exchange occurs between deuterated phenols and traces of moisture present in
the medium, the isotope effect is measurable only by saturation of the monomer
with heavy water. Electron-repelling substituents in the phenols increase the
kinetic isotope effect.
The kinetic isotope effect is great, k5/k5
H D = 4.8 (phenol) and 15.5 (tetramethyl
phenol). The term #expwas near to theoretical (#exp = 1.5 + 0.4).

3.2.7.2. Effect o f aromatic amines - The polymerization of vinyl acetate was


inhibited or retarded by substituted aniline derivatives. The reactivities of the 23
752 F. TODOS and T. FOLDES-BEREZSNICH

compounds investigated125vary within a range of two orders of magnitude. The


N,N-disubstituted aniline derivatives (which do not contain an N-H bond) can
be regarded as practically inert. In contrast with the behavior of phenols, even
small ortho substituents (-CI, -CH 3, etc.) considerably decrease the reactivity of
aniline derivatives. The N-heterocyclic compounds containing an N-H bond
also belong to this group. For example, the reactivity of phenothiazine can be
compared to that of the most active nitroso compounds in the polymerization
of methyl acrylate.8s The stoichiometry is normal ~exp = 1.93) but a slight
amount of chain regeneration can be observed at 50°C.
The reactivity of p- and m-substituents is described with sufficient accuracy by
the Hammett equation with an electrophilic substituent constant:
log k5/1~55 = - 0.6a + . (112)
The homolytic splitting of a C-H bond in polymer chemistry is normally the
rate determining step of a chain transfer reaction. However, in the case of some
hydrocarbons of special structure, true inhibition takes place. Among the 14
compounds studied, ~26the most reactive are 9,10-dihydroanthracene (fl = 350)
and 5,12-dihydrotetracene (fl = 432) in the polymerization of vinyl acetate at
50°C.
In the study of the temperature dependence of transfer reactions, the interest-
ing observation can be made that the activation energy (E) and pre-exponential
factor (A) of some reactions are not independent from each other but they fulfill
the following linear correlations: 127
logA = logA 0 + ~,E (113)
where A0 and V are constants characteristic of the type of reaction. According
to the investigations, for these reactions
1
> 2.3R----T" (114)

In this case the change in the pe-exponential factor is predominant. Therefore,


in contrast with the usual situation, the value of the rate constant increases with
increasing activation energy (over compensation).
By the use of inhibitors with antioxidant character, an interesting synergism
can be observed in the presence of oxygen: the inhibition effect is much stronger
than without oxygen. Recently, several authors dealt with the explanation and
kinetic treatment of this phenomenon, ~28-13° using phenolic inhibitors in the
polymerization of styrene, methyl methacrylate and acrylic acid. Although
antioxidants are weak inhibitors of radical polymerization, they are widely
applied in the industry, despite the fact that theit stabilization effect ceases in a
closed container after the oxygen is consumed.
Due to the limited size of our survey, we are unable to publish the inhibition
kinetic parameters of all inhibitors studied. In its time, Ulbricht's tabulation TM
FREE-RADICAL POLYMERIZATION 753

was considered a practically complete compilation of k5 data and a great


number of #-values can be found in Ref. 1.

3.3. Interpretation of the stoichiometric anomalies


The main problem we are faced with when using the inhibition method for the
determination of the rate of initiation is the uncertainty in the stoichiometry of
the inhibition reaction.
It seems to be a general rule that the experimental stoichiometric coefficient
of inhibitors which are not themselves free radicals is smaller than the theoreti-
cal one. This would, however, mean that the experimental determination of its
value cannot be disregarded since the assumption of a "theoretical" (strictly
speaking, speculative) value for the stoichiometric coefficient is liable to cause
an error of as much as two orders of magnitude in connection with the rate of
initiation. This explains the contradictory information obtained by using the
inhibition method and the skepticism concerning its practicability. It is an
unquestionable fact that only the inhibitors reacting with one electron (i.e. the
metal ions with varying valencies and the stable free radicals) can be regarded
as "reliable" and even these have to be employed with considerable caution.
Several attempts were made by the author to account for the observed
stoichiometric anomalies, by assuming various (more or less trivial) reaction
schemes. In the course of systematic control-examinations, however, these
reaction schemes, based mainly on parallel side reactions, proved, one by one,
to be contradictory, so that a new interpretation of the observed phenomena
had to be found.
In the reactivity of the inhibitors certain regularities were observed which
have led to the conclusion that, independently of the differences in the chemical
nature of the inhibitors, the inhibition reaction is a single elementary chemical
act and not the sum of different parallel reactions. Consequently, the above
considerations concerning the stoichiometric coefficient can be brought into
connection only with the subsequent reactions of the intermediate radical
formed by the addition of the inhibitor molecule to the growing radical.
If the inhibition reaction consists of a single elementary reaction, then the
radical formed in the process ought to have a uniform chemical structure also.
Data on the stoichiometric coefficient do, however, show this radical to exist in
two forms:
(a) In a reactive form, readily reacting with the double bond of another
monomer molecule whereby a chain-carrying radical of the type R" is again
formed in chain regeneration.
(b) In a non-reactive form, no longer able to react with the monomer; there-
after this form reacts without activation energy with an additional polymeric
radical and is transformed into an inert end product (inhibition).
754 F. TODOS and T. FOLDES-BEREZSNICH

Since no chemical distinction can be made between the two forms of the
radical formed as an intermediate product, one has to assume that the difference
is of a physical or, more exactly, of an energetic nature. Only the energy
liberated during the inhibition reaction can be held responsible for such a
difference. Since the reaction is of the addition type, the energy liberated cannot
be transformed into translational energy but has to appear in the form of
potential (vibrational) energy. The whole scheme of reaction can thus be given
by the aid of the following reactions:
R" + Z --, R - Z * (ks) (115)
where R" denotes the macroradical, Z the inhibitor molecule and the asterisk the
excess energy; the reactive intermediate-product radical ( " h o t " radical) may
enter into chain regeneration:
R - Z * + M--* R - Z - M " ( = R ' ) (k*) (116)
or, if colliding "by mischance" with a monomer molecule, it may lose its excess
energy:
R-Z* + M ~ R-Z + M (k*) (117)
where also its reactivity will decrease ("cold" radical). The next reaction will
thus be as follows:
R-Z" + R" --, R - Z - R (k~). (7)
It is to be noted that in principle the possibility that the reaction
R - Z " + M --', R" (kr) (6)
takes place, cannot be excluded either. This would cause the regeneration of the
cold radical. However, as both in the case of retardation (2.3.6.) and in that o f
inhibition (2.3.4.), such a reaction brings forth characteristic deviations from the
macrokinetic laws of the simple inhibition, it can be readily detected. The
regeneration of the cold radical, except in a few special cases, is so slow that it
can be neglected without causing any considerable error. The slowness of the
reaction resides with its high activation energy; according to our measurements,
in the reaction (semiquinone radical + styrene), the value of the activation
energy, E6, amounts to 9.4 + 2.5Kcalmol-~. ~9
During the reactions 115 and 116 of the previous reaction scheme, one
inhibitor molecule is consumed, but the number of macroradicals does not
change. Hence the stoichiometric coefficient for this reaction pathway is zero. In
the consecutive series 115, 117 and 7, one inhibitor molecule and two chain-
carrying radicals are consumed. Here the value of the stoichiometric coefficient
will be two. The probability (~) that the hot radicals will "cool down" obviously
depends on the rates of reactions (116) and (117) relative to each other:

=
kt
k~' + k~*" (118)
FREE-RADICAL POLYMERIZATION 755

When this equation is used in the case of a two-step inhibition, the value of the
stoichiometric coefficient is
# = 2~. (119)
From the above-mentioned reaction scheme, the experimentally determined
inequality is easily obtained:
0 < /~exp < /~heor.
In the light of this statement, the experimental fact that the stoichiometric
coefficient does not depend on temperature will also seem to be self-evident. The
rate constants in expression (6) then belong to the reactions of the hot radical
which dispose of large amounts of energy. Since these reactions do not require
any external activation, the quotient of these constants should obviously be
independent of the temperature.
Bevington's work is evidence supporting the statement that the surplus energy
of the reacting molecule can influence the rate of reaction in the liquid phase.
To control further this assumption in a direct way, we conducted experiments
in which vibrationally excited radicals were produced by UV irradiation of
AIBN. ~33The primary hot radicals formed were scavenged by various nitroso
compounds ("spin trapping") and transformed into more stable radicals, the
concentrations of which could be measured by the ESR method. A comparison
of the steady concentrations of intermediate radicals formed under the effect of
thermal and hot radicals (the ratio [ R - Z * ] / [ R - Z ' ] reaches 10-102) permits the
determination of the rate and probability of relaxation of the hot radicals. It was
found that for the complete relaxation of the hot cyanopropyl radicals in
benzene about 2 • 1 0 6 collisions are needed. ~33
If the value of the stoichiometric coefficient really depends on the physical
process of energy transfer, then it should be possible to detect it by changing the
rate of the vibrational deactivation. One may add to the system a chemically
inert solvent which is unable to react with any of the reaction components under
the conditions of the reaction. The concentration of the monomer and with it
also the rate of reaction 116, will decrease. By contrast, the rate of the cooling-
down process will change only to a slight extent or not at all. In such cases, an
additional deactivation process has to be taken into consideration, in which the
solvent (S) is the energy-transferring partner:
R-Z* + S --, R-Z" + S (k~*). (120)
Accordingly, in the case of polymerizations taking place in bulk and in solution
respectively, the values of the stoichiometric coefficients must differ from each
other. In this case the probability that the cooling-down will occur can be given
by the following equation:
k*r*m + k'8*r*s k* + (k'a*s/m)
ot = k*r*m + k*r*m + k~*r*s = k* + k* + (k's*s/m)" (121)
756 F. T1DDOS and T. Ft~LDES-BEREZSNICH

1000[ 12 2.5

800 1 0.3 0

6o0
C
E
"" 400

200

0 10 20 30 40 50
ZO/,~"102

FIG. 12. The effect of dilution on the length of the inhibition period. System: styrene/
AIBN/benzene/tetrabromoquinone/50°C. The approximate value of dilution (s/m) is
indicated for each straight line.

It is readily conceivable that the value of ct is higher in solution than in bulk


polymerization and furthermore that the limiting value of tr tends towards unity
with increasing dilution:
lim 0t = 1 (122)
$ / m ~ oo

i.e. on dilution the value of the stoichiometric coefficient tends towards the
theoretical value:

lim /~ = /4heor" (123)


$ / m ~ oo

The validity of relationships (121-123) can be checked experimentally. In


these investigations the following inhibitors have been examined: ~tetrabromo-
quinone (bromanil, BA) and tetrachloroquinone, tetracene (Tc), 1,3,5-trini-
trobenzene, picric acid and N,N-dimethyl-p-nitrosoaniline (NDMA) in the
polymerization of St. We have also looked at NDMA and [,-nitrosodiphenyl-
amine in the polymerization of AN and MA. 134
In Fig. 12 the results of a complete series of measurements are shown. They
are, by the way, rather characteristic. Different s/m parameter values belong to
the individual curves (s/m = 0 means, by definition, bulk polymerization). The
system examined was styrene-tetrabromoquinone in benzene solvent at 50°C,
using azobisisobutyronitrile (AIBN) as initiator.
As can be seen, the slopes (#/2kLf, cf. eq. (23)) of the curves in Fig. 12 rapidly
increase with increasing s/m values. If the values of/~ are calculated from the
slope and the data are plotted against the molar fraction (XM) of the monomer,
then curves of the type shown in Fig. 13 are generally obtained.
FREE-RADICAL POLYMERIZATION 757

2.0

1.s ~"t~qqDN ~ e

/J. 1.0 ,q

0.5

I I I I
0 0.2 0.4 0.6 0.8 1.0
XM

FIG. 13. The stoichiometric coefficient plotted against the mole fraction of monomer
in the monomer-solvent mixture. System: styrene/AIBN/benzene/50°C. Inhibitor:
O - tetracene, • - tetrabromoquinone.

One of the two curves shows the change of the stoichiometric coefficient of Tc.
the other that of BA. It can be readily seen that in the case ;~u --* 0 the
stoichiometric coefficients of both inhibitors tend towards the theoretical value
of two. Consequently, the values of the stoichiometric coefficient deduced from
chemical considerations are valid only in the case of infinite dilution. The two
full curves have been calculated according to eq. (121). Consequently, the above-
mentioned reaction scheme is supported by the experimental data not only
qualitatively but also quantitatively.
It is evident that elementary inhibition reactions cannot essentially differ from
other exothermic reactions leading to the formation of free radicals. If, however,
in such elementary reactions the reaction heat liberated has such a decisive role,
it is hardly conceivable that in the case of analogous elementary reactions
similar consequences should not be reckoned with. It can thus be stated as a
general rule that every exothermic reaction in which an active center is formed,
leads to the formation of a hot particle. It is easy to concede that the chemical
character of the active center (free radical, carbonium ion or carbanion, respec-
tively) is indifferent. It seems probable that this effect may play an especially
important role in processes where the mechanism involves constant repetition
of these reaction steps, such as the propagation step in chain reactions.

REFERENCES
1. F. Tf3IXSs, Acta chirn, hung. 43, 97 and 44, 403 (1965).
2. N.I. SMIRNOV,F. TOPOS and V. FOR,sT, Fiirst, Trudy, leningr, tekhnol. Inst. 44, 60 (1958).
3. F. T0t)Os and N. I. SMmNOV, Acta chim. hung. 15, 389 (1958).
4. F. Tf3txSs and N. I. SMIRNOV,Acta chim. hung. 15, 401 (1958).
5. F. TOPOs, Magy. Tudom. Akad. Ki~zp. k~m. kut. Int~z. Kfzl. 2 51 (1959).
758 F. T~DOS and T. FOLDES-BEREZSNICH

6. H.S. BAGDASARYAN)Theory of radical polymerization, Izd-vo A.N.U.S.S.R. Moscow


(1959).
7. H.S. BAGDASARYAN,Zh.fiz. Khim. 20, 1415 (1953).
8. G. M. BURNETr, Mechanism of Polymer Reactions, Interscience, New York (1954).
9. G . M . BURI~TT and P. R. E. J. COWLEY,Trans. Faraday Soc. 49, 1480 (1953).
10. F. TOPOS, Kinetics of inhibited thermal polymerization of styrene, Thesis, LTI im.Len-
sovieta, Leningrad (1956).
11. C.H. BA~ORD, A. D. JENKINS and R. JOHNSTON, Proc. Roy. Soc. A239, 214 (1957);
C. H. BAMFORD,W. G. BARB,A. D. JENKINSand P. F. ONYON, The Kinetics of Vinyl Poly-
merization by Radical Mechanisms, Butterworths, London (1958).
12. F. TODOS,and N. I. S~lmNOV,Acta chim. hung. 54, 255 (1967).
13. F. TOvOs, T. BEREZSNICHand B. TURCS.~NYI,Vf~sokomolek. Soedin. 4, 1584 (1962).
14. F. T/)D(SS, T. FSLDES-BEREZSNICHand M. AZORI, IUPAC International Symposium on
Macromolecular Chemistry Proceedings, Vol. III B-4 Wiesbaden (1959), Chimia i Technologia
Polymerov. I, 78.
15. F. TODOS,T. F6LDES-BEREZSNICHand M. AZOR[, Acta chim. hung. 24, 91 (1960).
16. I. KENDE,L. SOMEG[and F. TODOS,Kinetics and Mechanism ofPolyreactions, VoL 3, p. 97,
Publishing House of the Hung. Acad. of Sci., Budapest (1969).
17. I. KENDE, L. SOMEG1and F. TODOS, Fur. Polym. J. 8, 1281 (1972).
18. F. TODOS, Some Problems of the Kinetics of Radical Polymerization, Thesis, LTI im.Len-
sovieta, Leningrad (1964).
19. F. T(~DOSand N. I. SMIRNOV,Acta chim. hung. 15, 409 (1958); F. T(R~3s, T. SIIVL~NDIand
M. AZORX, Vf~sokomolek. Soedin. 4, 1431 (1962).
20. F. T/~DOS,T. SIM~NDIand M. AZORI, Vf,sokomolek. Soedin. 4, 1431 (1962).
21. F. TOI~8~S,I. KENDEand M. AZORI,J. Polym. Sci. AI, 1353 (1963); J. Polym. Sci. AI, 1369
(1963).
22. T. F()LDES-BEREZSNICHand F. TOD6S, Fur. Polym. J. 2, 219 (1966).
23. P.D. BARTLETTand H. KWART, J. Am. chem. Soc. 72, 1051 (1950).
24. P.D. BARTLETTand H. KWART, J. Am. chem. Soc. 74, 3969 (1952).
25. L J. KICE, J. Am. chem. Soc. 76, 6274 (1954).
26. H.S. BAGDASARYAN,Zh. fiz. Khim. 32, 2614 (1958).
27. F. TOD6S, I. KENDEand M. AZOR1, Vysokomolek. Soedin. 4, 1262 (1962).
28. L. ZIKMUND,Colin Czech. chem. Commun. 34, 1543 (1969).
29. F. TODOS,I. KENDE,T. BEREZSNICH,S. SOLODOVNIKOVand V. VOEVODSKU,Kinet. Catal. 6,
203 (1965).
30. E.G. JANZENand B. J. BLACKBURN,J. Am. chem. Soc. 90, 5909 (1968).
31. M.P. BEREZIN,L. I. MACHONINAand G. V. KOROLYOV,Vf,sokomolek. Soedin. Ser. A. 17,
1043 (1975).
32. N . N . TVOROGOVand A. G. KONDIC~,TJEVA,V~sokomolek. Soedin. Ser. A. 20, 230 (1978).
33. T.M. TURECHANOV,T. A. KUZMINAand L. B. IRISKINA,Dokl. Akad. Nauk SSSR 2621, 370
(1982).
34. B.R. SMIRNOV, Vysokomolek. Soedin. Set. A. 24, 787 (1982).
35. B.R. SmRNOV, Vysokomolek. Soedin. Ser. A. 24, 877 (1982).
36. L G. SOROVTSEV,Vysokomolek. $oedin. Set. A. 20, 451 (1986).
37. M. KUalN and L. ZIKMUND,Colin Czech. chem. Commun. 34, 1254 (1969).
38. N . N . SE~NOV, Zh.fiz. Khim. 17, 187 (1943).
39. N . N . SE~NOV, Chemical Kinetics and Chain Reactions, Clarendon Press, Oxford (1935).
40. A. F~F~RVi~RI,T. FSLDES-BE~zSNICHand F. TOPOS, J. macromolec. $ci. Chem. AI6, 993
(1981).
41. T. F6LD~-BF.REZSNICH,M. SZ~ZTAY, E. BOROS-GYEvland F. TODOS, J. macromolec. Sci.
Chem. A16, 977 (1981).
42. J. SZAFKOand K. MANCZYK,Makromolek. Chem. Rapid Commun. 1, 153 (1980).
43. M. ARNOLD,T. TAPLICK, M. RATZSCHand R. H o v e , Acta polym. 31, (1980) 75.
44. •. FEH~RV./~Ri,T. F6LDES-BEREZSNICHand F. TODOS,J. macromolec. Sci. Chem. A14, 1071
(1980).
45. I. CZAJUK, T. Ft3LD~-BEREZSNICH, S. SzAr~cs and F. TODOS, Fur. Polym. J. 14, 1059
(1978).
FREE-RADICAL POLYMERIZATION 759

46. I. CZAJLIK, T. F6LOES-B~ZSNICH, F. Tlh~Ss, and I~. V~R~r.S, Eur. Polym. J. 17, 131
(1981).
47. G. KASZAS,T. FOLDES-BEREZSNICHand F. TOpOs, Fur. Polym. J. 19, 469 0983).
48. Gv. FENYVESl,A. FEH~VAm, T. F6LDES-BEREZSNICHand F. TOPOS, J. macromolec. Sci.
Chem. A12, 203 (1985).
49. Zs. LASZL6-HEDVIG,L. SOMF_.G!and F. TODd, Kinetics and Mechanism of Polyreactions,
Vol. 3, p. 87, Publishing House of the Hung. Acad. of So., Budapest (1969).
50. A. MILLER and J. SZAFKO,J. Polym. Sci., Polym. Chem. Edn 18, 1177 (1980).
51. T. F6LDES-BEREZSNICH, M. SZESZTAY, E. BOROS-GVEVI and F. TOD6S, J. Polym. Sci.,
Polym. Chem. Edn 18, 1223 (1980).
52. J. SZAFKOand K. MANCZYK,Makromolec. Chem. 179, 2719 (1978).
53. I. CZA~LXK,T. FOLDES-BEREZSNlCH,F. TODOSand ~. V~RTES,J. macromolec. Sci. Chem. 14,
1243 (1980).
54. F . H . BANFIELDand J. KENYON,J. chem. Soc. 1612 (1926).
55. A.V. TRUBNIKOV, M. D. GOLDFF~N, A. D. STEPUKHOVICHand E. A. RAFIKOV, Vysoko-
molek. Soedin. Ser. B. 18, 419 (1976); A. V. TRUBN1KOV,M, D. GOLDFEIN,N. V. KOZHEV-
NIKOV,E. A. RAFXKOV,A. D. STEPUKHOVICHand V. I. TOMASHCHUK,Vy,sokomolek. Soedin.
Set. ,4.70, 2448 (1978).
56. A. RF.HAK,B. TURCSANVland F. TODOS, Kinetics and Mechanism of Polyreactions, Vol. 3,
p. 91, Publishing House of the Hung. Acad. of Sci., Budapest (1969).
57. M. KINOSHITAand Y. MINON, Makromolek. Chem. 124, 211 (1969).
58. F. TOPOS, and T. FOLDES-BEREZSNICH,J. macromolec. Sci. Chem. 1, 523 (1966).
59. T. FOLDES-BEREZSNlCHand F. TODOS, Acta chim. Hang. 42 149 (1964).
60. T. FOLDF.S-BF.P,EZSNICHand F. TODOS, Eur. Polym. 3". 2, 219 (1966).
61. F. TOPOS and T. F6LDES-BEREZSNICH,Fur. Polym. J. 2, 229 (1966).
62. E.C. KOOYMANand E. FARENHORST,Trans. Faraday Soc. 49, 58 (1953).
63. M. SZWARC,J. Polym. Sci. 16, 367 0955).
64. J.H. BINKSand M. SZWARC,J. chem. Phys. 30, 1494 0959).
65. J. KOUTECZKY,R. ZAHRADNIKand J. CIZEK, Trans. Faraday Soc. 5"7, 169 (1961).
66. A. REMB^UMand M. SZWARC,J. Am. chem. Soc. 77, 4468 0955); R. P. BUCKLEY,A. REM-
BAUMand M. SZWARC,J. chem. Soc. 3442 (1958).
67. A. BICKEL and W. A. WATERS, J. chem. Soc. 1764 (1950); B. A. GINGRAS and W.A.
WATERS,J. chem. Soc. 1920 (1954).
68. J.C. BEVINGTON, N . A . GHA~rEM and H . W . MELVILLE, J. chem. Soc. 2822 (1955);
J. C. BEVINGTON,N. A. GHANEMand H. W. MELVILLE,Trans. Faraday Soc. 51,946 (1955);
J. C. Bevington and H. W. MELVILLE,Usp. Khim. 25, 1436 (1956); J . C. BEVINGTO~,Radi-
cal Polymerization, Academic Press, London and New York (1961).
69. J.W. BItr~ITENnACHand H. Ba£1TESB^CH,Z. phys. Chem. Ser. `4. 190, 361 (1942).
70. J.W. BREITENnACHand A. FALL¥, Mh. Chem. 84, 391 (1953); J. W. B~ITENB^CH, Z. Elek-
trochem. 60, 286 (1956).
71. P.D. Bartlett, G. S. Hammond and H. KWART, Discuss. Faraday Soc. 2, 342 (1942).
72. F. TOtX)s and T. L. SXMANOI,Vysokomolek. Soedin. 4, 1271 (1962).
73. F. TOpOs and T. L. SIMkNDX,Vysokomolek. Soedin. 4, 1425 (1962).
74. T.L. SIMANDIand F. Tfn~Ss, Fur. Polym. J. 21, 865 (1985); T. L. StM~NDI, F. TODOSand
B. TURCSk~ntl, J. Polym. ScL C16, 4607 (1969).
75. F. TOD~, $. LADIKand B. TuRcskNYI, Eur. Polym. J. 6, 1321 (1970).
76. T.L. SIM~NDI, F. TOD6S and B. MOHOS,Kinetics and Mechanism of Polyreactions, Vol. 3,
p. 129, Publishing House of the Hung. Acad. of Sci., Budapest (1969).
77. A.A. YASSINand A. M. EL-I~LDY, Eur. Polym. J. 9, 657 (1973).
78. A . A . YASSINand N. A. RIZK, Eur. Polym. J. 13, 441 (1977); Polymer 19, 57 (1978); J.
Polym. Sci., Polym. Chem. Edn 16, 1475 (1978).
79. T.L. SXMANOl,A. ROC~NnAtr~ and F. TOD~, Eur. Polym. J. lg, 67 (1982); 19, 427
(1983).
80. S.G. FooRo, J. chem. Soc. 48 (1940).
81. G.V. Scl~rl.Z, Makromolek. Chem. 1, 94 (1947).
82. G.S. HAMMONDand P. D. B A R ~ , J. Polym. Sci. 5, 617 (1950).
83. Z.A. SUqITSUqAand H. S. B^ODASAgVAN,Zh. fiz. Khim. 32, 2614 (1958); 32, 2663 (1958).
760 F. TI~DOS and T. F()LDES-BEREZSNICH

84. T. FOLDES-BEREZSNICH, F. Ti~DOS, and S. SZAK.~CS, Kinetics and Mechanism of Polyreac-


tions, Vol. 3, p. 103, Publishing House of the Hung. Acad. of Sci., Budapest (1969). Fur.
Polym. J. 8, 1237 (1972); Eur. Polym. J. 8, (1972) 1247.
85. B.A. G1NGRASand W. A. WATERS, J. chem. Soc. 1920 (1954).
86. Z.A. KARAPETYAN,G. V. ORESHKOand B. R. SMIRNOV,V~sokomolek. Soedin. Ser. A. 28,
305 (1986).
87. C.K. INGOLD and S. D. WEAVER,J. chem. Soc. 125, 1456 (1924).
88. I.T.~NCZOS, F. TOD6S and T. F6LDES-BEaEZSNICH, Eur. Polym. J. 18,295 (1982).
89. I. TANCZOS, A. REHAK and F. T~D6S, Eur. Polym. J. 18, 487 (1982).
90. I. KENDE, T. L. SIMAND1and F. TOD6S, Kinetics and Mechanism of Polyreactions, Vol. 3,
p. 109, Publishing House of the Hung. Acad. of Sci., Budapest (1969).
91. I. KENDE, T. L. SIMANDIand F. TOPOS, Polym. Bull. 3, 325 (1980).
92. S. SZAKACS,T. F6LDES-BEREZSNICH,F. T~3DOSand L. J6KAY, Eur. Folym. J. 15, 295 (1979).
93. R. HUISGEN, R. GRASHEY, H. HAUCK and H. SAIDL, Chem. Bet. 101, 2548 (1968).
94. I. KENDE, F. TOD6S and L. Sf3MEGI, Acta chim. hung. 54, 315 (1967); T. L. SIMANDI,
F. TOD6S and I. KENDE, .4cta chim. hung. 68, 75 (1971); T. L. SIMANDI,I. MAYER, I. KENDE
and F. TOD6S, Magy. k~m. Foly. 78, 254 (1972).
95. A.L. WARD and U. J. ROaERTS, Monomery, Uzd. Inostrannoy Literatury, Moscow (1953).
96. K. HIGASHIU~ and M. OIWA, J. Polym. Sci. A-l, 6, 1857 (1968).
97. I. GV6NGYHALMI,A. MATriX, F. Ti)DOS and T. T6TH, Magy. kdm. Foly. 87, 424 (1981).
98. J.L. KITE, J. Am. chem, Soc. 80, 348 (1958).
99. T. F6LD~-BEREZSNICH, F. Tt~DOS, L. JtSKAYand B. MOHOS, Bul. Inst. politeh, lasi 16, (20),
221 (1970).
I00. L.W. BREITENBACH, H. Pm~USSLER and H. KARLINGER, Monatsh. 80, 150 (1949); O. F.
OLAJ, J. W. BRErrENnACHand I. HOrFREITER, Monatschefte 89, 997 (1967).
101. D. BRAUN and G. ARCACHE, Makromolek. Chem. 148, 119 (1971).
102. G. SALOMON, B. B. S. T, BOONSTRY,O. VAN DER MEER and A. J. UL~E, J. Polym. Sci. 4,
(1949) 203.
103. T. MOTAYAMAand S. OKAMURA, Chemy high Polym. 8, 475 (1951).
104. O.F. OLAJ, J. W. BREm~NaACHand I. H o a r i e R , Makromolek. Chem. 110, 72 (1967).
105. M. MAGAT and R. BO~ME, Compt. rend. 232, 1657 (1951).
106. S. SZAK/~CS,L. J6KAY,T. F6LDES-BEPa'ZSNICHand F. TOPOS, Eur. Polym. J. 14, 181 (1978).
107. L.G. SUROVTSEV,V. V. KYSELEV, V. V. DVORTSOV and N. I. KH1ZHNYAK, Vysokomolek.
Soedin. Ser. B. 20, 765 (1978).
108. K. SUGIYAMA,T. ODA and T. MAESmMA, Makromolek. Chem. 183, 2445 (1982).
109. H. STAUDINGER, Chem. Ber. 58, 1075 (1925); H. STAUD1NGER, K. DYCKERrtOrr,
H. W. KLEWr.S and L. RUZI~KA, Chem. Ber. 58, 1079 (1925); H. STAUDINGER and
L. LAUTENSCHL~,GER,Liebigs Annln Chem. 488, 1 (1931).
110. C.E. BORNES, R. M. ELOFSONand G. D. JONES, J. Am. chem. Soc. 72, 210 (1950).
111. F . R . MAvoandA. A. MILLER, J. Am. chem. Soc. 78,(1956) 1017;78, 1023(1956);80,2493
(1958); F. R. MAYO, A. A. MILLER and G. A. RUSS~L, J. Am. chem. Soc. 80, 2500 (1958).
112. K.C. SMELTZand E. DYER, J. Am. chem. Soc. 74, 623 (1952); E. DYER, S. C. BROWN and
R. W. MEt)EROS81, 4243 (1959).
113. G.V. SCHULZ and G. HENRIO, Makromolek. Chem. 19, 437 (1956).
114. M. MUNZER, Angew. Makromolek. Chem. 11, 15 (1970).
115. F. TODOS, Zs. FODOR and M. IRING, The Kinetics and Mechanism of Inhibited Autoxi-
dation. In: Oxidation and Organic Materials (J. POSI'~IL and P. O. KLEMCHUK, Eds) CRC
Press Inc., New York, in press.
116. J. BREITENBACH,Bet. 71, 1438 (1938).
117. C. WALLINGand E. R. BRIGGS,J. Am. chem. Soc. 68, 1141 (1946).
118. B.A. DOLGOPLOSK and D. SH. KOROTKINA, Zh. obshch. Khim. 27, 2226 (1957); B.A.
DOLC_,OPLOSKand G. A. PARFENOVA,Zh. obshch. Khim. 27, 3083 (1957).
119. M.P. GOI)SAV, G. A. HARPELL and K. E. RUSSELL, J. Polym. Sci. 57, 641 (1962).
120. Y. MINOURA, N. YASIMOTOand T. Isml, Makromolek. Chem. 71, 159 (1964).
121. R.A. BIRD and K. E. RUSSELL, Can. J. Chem. 43, 2123 (1965).
122. A.D. JENKINS, Trans. Faraday Soc. 54, 1895 (1958).
123. M. SIMONVl, F. Ti3DOS and J. POSPI~IL, Eur. Polym. J. 3, 101 (1967).
FREE-RADICAL POLYMERIZATION 761

124. M. SIMONYI,F. TOD6S, S. HOLLYand J. POSPIglL,Eur. Polym. J. 3, 559 (1967); Kinetics and
Mechanism ofPolyreactions, Vol. 3, p. 125, Publishing House of the Hung. Acad. of Sci.,
Budapest (1969).
125. M. SIMONYIand F. TC'DOS,Kinetics and Mechanism ofPolyreactions, Vol. 3, p. 119, Publish-
ing House of the Hung. Acad. of Sci., Budapest (1969).
126. M. SIMONYI,F. TOPOS and L. KovAcs, Kinetics and Mechanism of Polyreactions, Vol. 3,
p. 115, Publishing House of the Hung. Acad. of Sci., Budapest (1969).
127. F. TOD6S, Kinetics and Mechanism of Polyreactions, Plenary and Main Lectures, p. 485,
Publishing House of the Hung. Acad. of Sci., Budapest (1971).
128. N . N . TvogoGov, V~sokomolek. Soedin. 17, 1461 (1975).
129. J.J. KURLAND,J. Polym. Sci., Polym. Chem. Edn 18, 1139 (1980).
130. SHow-AN CHEN and Ln-CI-IYNANTSAI, Makromolek. Chem. 187, 653 (1986).
131. J. ULBRICHT,Inhibitors and inhibition constants in free radical polymerization, In: Polymer
Handbook, lind Edn (J. BRANDRUPand E. H. IMMERGUT,Eds) p. II-53, Wiley, New York
(1975).
132. F. TODOS, Makromolek. Chem. 79, 8 (1964).
133. F. Tf~DOS, J. Polym. Sci., Part C., 16, 3461 (1968).
134. I.T.~NCZOS, T. FOLDES-BEREZSNICHand F. TODOS, Eur. Polym. J. 19, 225 (1983).

Potrebbero piacerti anche