You are on page 1of 294

LIST OF ARTICLES ON THERMODYNAMICS OF POLYMERIZATION

BY HIDEO SAWADA PUBLISHED IN POLYMER REVIEWS

Sawada, Hideo (1969) 'Thermodynamics of Polymerization. I', Polymer Reviews, 3, 313 — 338
DOI: 10.1080/15583726908545926
URL: http://dx.doi.org/10.1080/15583726908545926

Sawada, Hideo (1969) 'Chapter 2. Heat of Polymerization', Polymer Reviews, 3, 339 — 356
DOI: 10.1080/15583726908545927
URL: http://dx.doi.org/10.1080/15583726908545927

Sawada, Hideo (1969) 'Chapter 3. Thermodynamics of Radical Polymerization', Polymer Reviews,


3, 357 — 386
DOI: 10.1080/15583726908545928
URL: http://dx.doi.org/10.1080/15583726908545928

Sawada, Hideo ((1969) 'Chapter 4. Thermodynamics of Polycondensation', Polymer Reviews, 3,


387 — 395
DOI: 10.1080/15583726908545929
URL: http://dx.doi.org/10.1080/15583726908545929

Sawada, Hideo (1970) 'Thermodynamics of Polymerization. II. Thermodynamics of Ring-


Opening Polymerization', Polymer Reviews, 5, 151 — 173
DOI: 10.1080/15583727008085366
URL: http://dx.doi.org/10.1080/15583727008085366

Sawada, Hideo (1972) 'Thermodynamics of Polymerization. III', Polymer Reviews, 7, 161 — 187
DOI: 10.1080/15321797208068162
URL: http://dx.doi.org/10.1080/15321797208068162

Sawada, Hideo (1972) 'Thermodynamics of Polymerization. IV. Thermodynamics of Equilibrium


Polymerization', Polymer Reviews, 8, 235 — 288
DOI: 10.1080/15321797208068172
URL: http://dx.doi.org/10.1080/15321797208068172

Sawada, Hideo (1974) 'Thermodynamics of Polymerization. V. Thermodynamics of


Copolymerization. Part I', Polymer Reviews, 10, 293 — 353
DOI: 10.1080/15321797408076101
URL: http://dx.doi.org/10.1080/15321797408076101

Sawada, Hideo (1974) 'Thermodynamics of Polymerization. VI. Thermodynamics of


Copolymerization. Part 2', Polymer Reviews, 11, 257 — 297
DOI: 10.1080/15583727408546025
URL: http://dx.doi.org/10.1080/15583727408546025
Table of Contents
Thermodynamics of Polymerization. I 5
Chapter 1. Introductory Survey 5
I. THE CEILING TEMPERATURE CONCEPT 6
A. Thermodynamic Approach 6
B. Kinetic Approach 7
C. Determination of Ceiling Temperatures 11
II. ENTROPY AND FREE ENERGY CHANGES OF
POLYMERIZATION 14
A. Entropy of Polymerization 14
B. Determination of Entropy of Polymerization 16
C. Free Energy Changes of Polymerization 22
REFERENCES 29
Chapter 2. Heat of Polymerization 31
I. GENERAL ASPECTS 31
A. Breaking a Multiple Bond 32
B. Resonance 32
C. Steric Strain 33
II. VARIATIONS IN HEATS OF POLYMERIZATION 33
A. Steric Strain in the Polymer 35
B. Conjugation and Hyperconjugation 38
C. Hydrogen Bond and Solvation 39
III. EMPIRICAL ESTIMATION OF HEAT OF POLYMERIZATION 39
REFERENCES 41
APPENDIX 42
Chapter 3. Thermodynamics of Radical Polymerization 49
I. GENERAL ASPECTS 50
A. Energetics of Radical Polymerization 50
B. Degree of Polymerization 51
C. Activation Energies of Elementary Reactions 53
II. GENERATION OF FREE RADICALS 53
III. PROPAGATION REACTION 58
A. The Polanyi Relation 58
B. Reactivity and Heat of Polymerization 60
C. Ceiling Temperature 63
IV. INTERACTION OF RADICALS 63
A. Combination and Disproportionation Reactions 63
B. Interaction of Small Hydrocarbon Radicals 64
C. Interaction of Large Hydrocarbon Radicals 65
D. Interaction of Some Large Radicals 66
V. FREE ENERGIES OF FORMATION OF POLYETHYLENE
AND POLYTETRAFLUOROETHYLENE 68
A. Free Energies of Polyethylene Synthesis 68
B. Free Energies of Polytetrafluoroethylene Synthesis 73
REFERENCES 77
Chapter 4. Thermodynamics of Polycondensation 79
I. GENERAL ASPECTS 79
II. DEGREE OF POLYMERIZATION 80
III. EQUILIBRIUM CONSTANT 81
IV. RING FORMATION IN POLYCONDENSATION 86
ACKNOWLEDGMENTS 87
References 87
Thermodynamics of Polymerization. II. Thermodynamics of
Ring-Opening Polymerization 88
I. GENERAL ASPECTS 88
II. HOMOCYCLIC COMPOUNDS 89
A. Angle Strain 89
B. Conformational and Transannular Strain 99
C. Steric Effect of Side Group 100
III. HETEROCYCLIC COMPOUNDS 102
A. Cyclic Ethers 103
B. Lactams 105
C. Lactones 107
D. Miscellaneous Heterocyclic Compounds 107
IV. SUMMARY 109
ACKNOWLEDGMENTS 109
References 109
Thermodynamics of Polymerization. III (Cationic Polymerization) 111
I. GENERAL ASPECTS 112
II. FORMATION OF CARBONIUM ION 114
A. Ionization Potential 114
B. Proton Affinity 115
C. Acidity 118
D. Free Energy Change of Formation of Carbonium Ion 119
E. Ions and Ion Pairs 121
F. Energetics of Salvation 122
III. INITIATION OF CATIONIC POLYMERIZATION 124
A. Energetic Consideration of Initiation Reaction by Halogen Acid 124
B. Catalytic Activity in Cationic Polymerization by Lewis Acids 126
IV. PROPAGATION OF CATIONIC POLYMERIZATION 127
A. Energetics 127
B. Heats of Reaction of Cations with Olefins 129
C. Activation Entropy Changes of Propagation 132
D. Thermodynamics of Formation of Zwitterions 134
V. CHAIN TRANSFER AND TERMINATION 136
Acknowledgments 136
References 136
Thermodynamics of Polymerization. IV. Thermodynamics of
Equilibrium Polymerization 138
I. POSSIBLE TYPES OF EQUILIBRIUM POLYMERIZATION 139
II. SOME CASE STUDIES OF EQUILIBRIUM POLYMERIZATION 146
A. Vinyl Polymerizations 147
B. Ring-Opening Polymerizations 150
C. Polymerization of Aldehydes 165
III. TRANSITION PHENOMENA IN EQUILIBRIUM
POLYMERIZATION 167
IV. MOLECULAR WEIGHT DISTRIBUTION 173
A. Equilibrium Polymerization 173
B. Living Polymerization 179
V, THERMODYNAMICS OF EQUILIBRIUM POLYMERIZATION 183
Acknowledgment 188
References 189
Thermodynamics of Polymerization. V. Thermodynamics of
Copolymerization. Part I 192
I. THE GENERAL THEORY OF BINARY COPOLYMERIZATION 193
A. Heat of Copolymerization 193
B. Entropy of Copolymerization 200
C. Equilibrium Sequence Distribution 203
D. Free Energy Change in Binary Copolymerization System 209
E. Equilibrium Monomer Concentration 217
F. Penultimate Unit Effect 222
II. DEGREE OF POLYMERIZATION AND COPOLYMER
COMPOSITION OF BINARY COPOLYMERIZATION SYSTEM 224
A. Degree of Polymerization 224
B. Copolymer Composition Equation 228
III. MULTICOMPONENT COPOLYMERIZATION 243
A. Heat of Terpolymerization 243
B. General Theory of Multicomponent Copolymerization 248
Acknowledgment 250
References 251
Thermodynamics of Polymerization. VI. Thermodynamics of
Copolymerization. Part 2 253
I. RADICAL COPOLYMERIZATION 254
A. Heat of Copolymerization 254
B. Ceiling Temperature 259
C. Q-e Scheme 263
D. Substituent Effect 265
E. Effect of Polymerization Temperature 268
F. Effect of Solvent 272
II. IONIC COPOLYMERIZATION 274
A. Energetic Characteristics 274
B. Reactivity 279
C. Effect of Polymerization Temperature 280
III. OTHER COPOLYMERIZATIONS 282
A. Ring-Opening Copolymerization 282
B. Miscellaneous Copolymerizations 284
Acknowledgment 290
References 290
Thermodynamics of Polymerization. I
HIDEO SAWADA
Central Research Laboratory
Daicel Ltd.
Tsunigaoka, Oi, Intmagwi
Saitama, Japan

Chapter 1. Introductory Survey


I. THE CEILING TEMPERATURE CONCEPT 314
;
A. Thermodynamic Approach 314
B. Kinetic Approach 315
C. Determination of Ceiling Temperatures 319
II. ENTROPY AND FREE ENERGY CHANGES OF POLYM-
ERIZATION 322
A. Entropy of Polymerization 322
B. Determination of Entropy of Polymerization 324
C. Free Energy Changes of Polymerization 330
REFERENCES 337

313
314 HIDEO SAWADA

I. THE CEILING TEMPERATURE CONCEPT


A. Thermodynamic Approach
The Gib'bs free energy of a system at temperature T is defined as
G = H - TS (1)
where H is the enthalpy and S the entropy of the system. The free
energy change for any polymerization will be, therefore,

AG = Gpoiymer ~ G m o n o m e r
=
"polymer ~ H m o n o m e r — T^Spoiyj^ej.— S monomer ) (2)
= A H p - T AS
When the polymer has a lower free energy than the initial mono-
mer, a polymerization can occur spontaneously, and the sign of AG
is negative. A positive sign for AG signifies, therefore, that the
polymerization is not spontaneous. When the system is in equilib-
rium at a certain critical temperature, there is no tendency for
polymerization, and, hence, AG = 0 [1-3]. This temperature is
known as the ceiling temperature. These three possible conditions
for free energy change of polymerization may be summarized as
follows:

monomer — polymer AG = - (spontaneous)


monomer — polymer AG = + (nonspontaneous)
monomer — polymer AG = 0 (equilibrium)
At the ceiling temperature, T c , AG is zero, so that
T c = AHp/ASp (3)
where AHp and ASp are the enthalpy and entropy changes per mono-
mer unit. When the polymer chains are long these quantities are
identical with the heat and entropy changes of polymerization. If the
standard state refers to unit concentration and the monomer behaves
ideally, AS = AS° + R In [M]; thus

T (4)
AS° + R In [M]
THERMODYNAMICS OF POLYMERIZATION. I 315

where AS0 is the entropy change accompanying polymerization at the


standard state when the concentration of monomer is unity. There-
fore, T c can be raised by increasing the concentration of monomer
when a solvent is present. Equation (4) emphasizes that T c is char-
acteristic of monomer-polymer equilibrium only and is quite inde-
pendent of the monomer or the nature of the active centers in the
system; for a given value of [M], the ceiling temperature should,
therefore, be the same whether the active centers are radicals or
ions.
Many polymers are stable even above the ceiling temperature
only because of the difficulty of initiating degradative centers on the
polymer molecule. In practice, terminated polymer appears stable
at temperatures above the ceiling temperature, being in a state of
metastable equilibrium. Therefore, the polymer cannot depolymerize
spontaneously but can do so under appropriate conditions. Catalyst
residues that are not removed during the purification of a polymer
may also cause depolymerization reaction.
There are four important possibilities of polymerization as follows:
a. In addition polymerization, AH and AS are usually both nega-
tive, and so AG becomes positive above the ceiling temperature of
the system. Thus, the high polymer cannot be formed above the ceil-
ing temperature.
b. If the polymerization is endothermic (AH > 0) and AS is greater
than zero, no polymer can exist below a floor temperature, above
which AG becomes negative. The phenomenon of floor temperature
is exhibited by the polymerization of S8 rings.
c. When AH is positive and AS negative, AG is always positive;
therefore, polymer cannot exist at any temperature.
d. When AH is negative and AS positive, AG is always negative;
therefore, polymer can exist at any temperature.
B. Kinetic Approach
It is interesting to consider the ceiling temperature phenomenon
from a kinetic point of view [1,3]. At ordinary temperatures the
rate constant for depolymerization is small. However, the activation
energy of this rate constant is quite high (10-26 kcal/mole) compared
to that for propagation, and at high temperature the depolymerization
can become important compared with the polymerization.
Let us consider the propagation reaction

p
AAAAR
k^ n*i
316 HIDEO SAWADA

The rate constants for propagation and depropagation reactions can


be expressed as
k p = A p exp(-E p /RT), kd = Ad exp(-E d /RT)
where Ap and Ad are the collision frequency factors which approxi-
mate to the entropy of activation, and E p and E d are the activation
energies for polymerization and depolymerization, respectively. If
the degree of polymerization is large, E p - E d = AHp. For long
chains, AHp is equal to the heat of the overall polymerization.
The rate of propagation is essentially the same as the overall
rate of disappearance of monomer, since the number of monomers
used in chain transfer and initiation must be small compared to that
used in propagation if the polymer chains are long.
Therefore,

^I (5)
where [M*] is the concentration of propagating species, and [M] is
the concentration of monomer.
Depolymerization may now be considered to be the reverse of
propagation, then

v
d = k dt M nl (6
Thus the overall rate of polymerization is

The degree of polymerization is given by the rate of polymerization


divided by the rate of termination, i.e.,

_ (k d -k p [M])[M n ] (8)

f([Mn])

where f([M*]) is a function of the number of active centers, M*,


present.
AHp is usually negative (the polymerization is exothermic), and
so E d is usually much larger than E p . Therefore, although k d may
be negligible compared with kp[M] at ordinary temperature, it will
increase more rapidly with increasing temperature.
At the ceiling temperature T c the rate of depropagation becomes
THERMODYNAMICS OF POLYMERIZATION.

equal to that of propagation, regardless of the variations of [MJ^] and


) with temperature,

kp[M*][M] =

which can also be written in the form:

Ap exp(-E p /RT c )[M] = Ad exp(-E d /RT c ) (9)

and, therefore,
Ep - E d AHp
T U0)
c - R In (Ap[M]/Ad) ~ R In (Ap[M]/Ad)

At this temperature, the extrapolated Rp vs T and DP vs T curves


will cut the temperature axis. This is illustrated in Fig. 1.1. At
temperatures that are not far below the ceiling temperature, only
polymers of low molecular weight form.
It seems to be impossible to predict the variations of Rp and DP
with temperature right up to T c from Eqs. (7) and (8). When kd ap-
proaches kp[Mx], DP becomes small and consumption of monomer
in the initiation process is no longer negligible. Then, k p and k d
may show a dependence on DP. Nevertheless the limiting slope of
dRp/dT as T approaches T c may be numerically so large that such
effects are of very minor importance and operate only over the last
fraction of a degree below T c . The limiting slope for the rate vs
temperature curve can be obtained by differentiating Eq. (7) with
respect to temperature:

dR p /dT = [M*](k p [M]E p /RT*-k d E d /RT 2 )

+ (k p [M]-k d )d[M*]/dT

and substituting kp[M] = kd when T = T c :

T lim c (dR p /dT) = kp[M][M*] (Ep - E d )/RT C ( n )

= k p [M][M*]AH p /RT c

Then kp[M][Mj![] is the rate that would have been observed at T c in


the absence of depropagation.
From transition state theory the frequency factor is given by
A = (kT/h) exp(AS*/R). Therefore,
318 HIDEO SAWADA

TEMPERATURE
Fig. 1.1. Expected shapes of rate vs temperature and DP vs temperature
graphs, without chain transfer to monomer.

X = lim
TT

AS0 = AS* - ASd = - R In A d /A p (12)


Substituting Eq. (12) into Eq. (10), we find

c (4)
AS3 + R In [M]
Polymerization of monomer and depolymerization of the polymer
may take place under various conditions. These are indicated in
Table 1.1 for polymerization (for depolymerization, numerical values
have the opposite sign). When the polymer is partially or wholly
crystalline this may be denoted by c'. The superscript (°) for the
THERMODYNAMICS OF POLYMERIZATION. I 319

Table 1.1
Thermodynamic Definitions for Polymerization-Depolymerization Equilibria

Notation State of monomer State of polymer

gg Gas Gas (usually hypothetical)


gc Gas Condensed (liquid or
amorphous solid)
1c Liquid Condensed
Is Liquid Solution in monomer
ss Solution Solution
sc Solution Condensed

thermodynamic quantities indicates that the standard state is speci-


fied (e.g., ASgS refers to the standard state of 1 mole/liter of mono-
mer). If a mixture of different monomers is the starting material,
the standard state is unit concentration for each monomer.
In general, a pure liquid monomer that gives an insoluble polymer
will have a single well-defined ceiling temperature, given by
T c = A H I C / A S I C . A pure liquid monomer that gives a soluble poly-
mer will have a series of ceiling temperatures corresponding to dif-
ferent percentage conversions of monomer to polymer. The ceiling
temperatures for a number of monomers in bulk are given in
Table 1.2. The ceiling temperatures at equilibria involving mono-
mers in solution have been extensively investigated and the results
are summarized by Ivin [4].

C. Determination of Ceiling Temperatures


The ceiling temperature for a monomer in its standard state of
unit concentration is given by the term, T c = AHp/ASp. Thus, if the
ceiling temperature cannot be measured directly, it may be estimated
from AHp and ASp. Experimentally, the ceiling temperature may be
estimated from plots of the rate of formation of polymer (or average
molecular weight of the high polymers) against T [2]. As the ceiling
temperature is approached, the rate decreases steeply. Extrapola-
tion to zero rate gives, therefore, the required ceiling temperature.
The accuracy of the method depends on the steepness of the descend-
ing line and usually it is good to about 2-3°C.
Comparing Eqs. (4) and (10), we may write
ASp = R In (A p /A d ) + R In [M] = ASp + R In [M]
320 HIDEO SAWADA

Table 1.2
Ceiling Temperatures of Pure Liquid or Gaseous Monomers in Bulk
Mole fractione
Standard T
c (monomer at
Monomer states °C equilibrium) Ref.
a
Acetaldehyde Is -31 1 g
Is -3& 1 g
n-Butyraldehyde Is -16 1 h
Chloral gc 96 760 mm Hg f i
c
Ethylene gg 407 760 mm Hg f j
Formaldehyde gc 126 760 mm Hg f k
gc 133 760 mm Hg f 1
Methacrylic acid,
ethyl ester gc 173 760 mm Hg f m
Methacrylic acid,
methyl ester gc 164 760 mm Hg f n
Methacrylonitrile 177 o
Propionaldehyde Is -31a 1 g
Is -39b 1 g
Propylene gg 300C 760 mm Hg f j
d P
Selenium Is 83 1
Styrene gg 235= 760 mm Hg f j
gc 275C : 760 mm Hgf j
Styrene, a-methyl Is 61 760 mm Hg f q
Sulfur . Is 159 d 1 r
Sulfur trioxide lc 30. 4 1 s
Tetrahydrofuran Is 70 ± 5 1 t
Is 80 ± 3 1 u
Trioxane gc 36= 1 • V

a
Atactic polymer.
b
Isotactic polymer.
c
Calculated value.
d
Floor temperature.
e
Based on total monomer units.
1
Equilibrium pressure.
g
A. M. North and D. Richardson, Polymer, 6, 333 (1965).
h
O. Vogl, J. Macromol. Set., Al, 243 (1967).
'W. K. Busfield and E. Whalley, Trans. Faraday Soc,59,679 (1963).
••Reference [2].
k
F. S. Dainton, K. J. Ivin, and D. A. G. Walmsley, Trans. Faraday Soc,
55, 61 (1959).
THERMODYNAMICS OF POLYMERIZATION. I 321

Table 1.2 (continued)


'Reference [37].
m
R. E. Cook and K. J. Ivin, Trans. Faraday Soc, 53, 1132 (1957).
n
K. J. Ivin, Trans. Faraday Soc, 51, 1273 (1955).
°S. Bywater, Canadian J. Chem., 35, 552 (1957).
PA. Eisenberg and A. V. Tobolsky, J. Polymer Set., 46, 19 (1960).
q j . G. Kilroe and K. E. Weale, J. Chem. Soc, 1960, 3849.
r
F . Fairbrother, G. Gee, and G. T. Merrall, J. Polymer Set., 16, 459
(1955).
S
D. C. Abercromby, R. A. Hyne, and R. F. Tiley, J. Chem. Soc, 1963,
5832.
'C. E. H. Bawn, R. M. Bell, and A. Ledwidth, Polymer, 6, 95 (1965).
"Reference [42].
v
Reference [18].

Therefore, an equilibrium concentration of monomer at a given tem-


perature is given by

Rewriting Eq. (4), we obtain

ln[M]=^P-^!p (13)
eq
RTC R

Equation (13) gives the equilibrium monomer concentration as a


function of the ceiling temperature. Thus, if the monomer is vola-
tile, T c may be estimated from measurements of the equilibrium
concentration (pressure) of the volatile species over the polymer at
different temperatures.
Since both AHp and ASp are functions of temperature, the correct
method of evaluating T c is to plot AHp and T ASp against tempera-
ture and to identify the ceiling temperature with the points of inter-
section of these curves. The enthalpy and entropy of polymerization
are generally not very dependent on temperature if phase changes do
not take place. Thus, it is usually accurate enough to operate with
values given for a definite temperature, e.g., 25°C. Since both AHp
and ASp change in the same direction with temperature, their ratio
changes only slightly.
322 HIDEO SAWADA

II. ENTROPY AND FREE ENERGY CHANGES OF


POLYMERIZATION
A. Entropy of Polymerization
The entropy of a system is a measure of the statistical probability
or degree of disorder of that system. The depolymerization of one
polymer molecule to many monomer or oligomer molecules is ac-
companied by an increase in translational entropy since it is a dis-
sociative process, whereas polymerization results in a decrease in
translational entropy. The growth of oligomers should probably lead
to a smaller change in the entropy of the system than the growth of
analogous high polymers.
The entropy values of polymerization for most monomers are
quite similar. In vinyl polymerizations, for instance, the entropy of
polymerization is within the limits of 25 to 30 eu for a wide variety
of monomers, although the monomers have very different ceiling
temperatures. The ceiling temperature should reflect the binding
forces in a polymer, since depolymerization introduces disorder
into the polymer, and a high binding energy between monomer units
would tend to oppose the introduction of disorder. The higher the
ceiling temperature, the stronger the binding forces between mono-
mer units in the polymer. Thus,

ASp = AH p /T c = 25-30 cal/°K/mole (14)

An empirical relation between the energy of vaporization and the


boiling temperature is known as Trouton's rule. Equation (14) is
analogous to Trouton's rule.
Steric hindrance in the polymer, which markedly affects AHp,
has comparatively little effect on ASp. Steric hindrance may de-
crease the rotational entropy of the polymer chain by increasing its
rigidity and, hence, it could increase the value of -AS p . The loss of
internal rotational entropy in the polymer as a result of steric hind-
rance seem to outweigh the gain of internal vibrational entropy. How-
ever, the observed changes in entropy of polymerization are rela-
tively insignificant, e.g., - A S p for styrene polymerization to a solid
polymer amounted to ~ 25 eu, whereas the corresponding value for
a-methyl styrene was found to be ~ 26 eu. This is, indeed, a small
change when compared with a 10 kcal/mole decrease in the heat of
polymerization. Therefore, side-group steric hindrance serves to
lower the ceiling temperature further with increasing steric repul-
sion of the side groups.
THERMODYNAMICS OF POLYMERIZATION. I 323

Monomer resonance implies higher bond energies and relatively


large vibration frequencies of the bond within the monomer. For
example, a bond which is increased in strength from 80 to 88 kcal
and in vibration frequency from 1000 to 1100 cm' 1 will change its
entropy contribution at room temperature from 0.094 to 0.065 cal/
deg mole. The total effect of monomer resonance will always be a
reduction of the internal entropy of the monomer. Therefore, rela-
tively small entropy changes of conjugated vinyl monomers and
styrene are probably due to the decreased total internal entropies
of these monomers. The entropy contribution to the free energy
change at 25°C will vary only within 7.4 to 9.0 kcal/mole. Struc-
tural influences on AG, thus, operate mainly through the enthalpy
terms.
Dainton and Ivin [2] analyzed AS|g in terms of the component
entropy changes, translational ASj., external rotational AS r , vibra-
tional ASV, and internal rotational ASj r , in order to try to under-
stand why the entropy is an approximately additive property, lead-
ing to approximately constant values of ASgg. Values of St, S r , and
S v for the monomer can be calculated from standard formulas,
since its molecular weight, moment of inertia, and vibrational fre-
quencies are known. Sj r is found by difference (S^ + S r + Sv) and
the experimental (third-law) value for Sg (monomer). For the poly-
mer, it is readily shown from standard formulas that S v + Sj r » St
+ S r regardless of molecular shape, so Sj + S r may be neglected.
Thus, S v + Si r is the standard entropy of the gaseous polymer, which
must be found semiempirically. The numerical values for ethylene,
isobutene, and styrene are summarized in Table 1.3.
These values show that on polymerization the loss of external
rotational entropy nearly balances the gain in vibrational and in-
Table 1.3
Analysis of AS g g for Vinyl Polymerization 3 (cal/deg/mole)

,, Polymer
Monomer .
Substance M St Sr Sv Sir S| unit,
Si g =Sv + Sir -ASgg

Ethylene 28. 05 35 .9 15.9 0.6 0 52.4 18 .4 34.0


Isobutene 56. 10 38 .0 23.1 - 9.1 70.2 29 .2 41.5
Styrene 104. 14 39 .8 27.9 10.1 4.7 82.5 47 .0 35.5

"Values from F. S. Dainton and K. J. Ivin, Quart. Rev., 12, 81 (1958).


324 HIDEO SAWADA

ternal rotational entropy, so that -AS|g has a value quite close to


the monomer's translational entropy; this is fairly insensitive to the
molecular weight or structure of the monomer.
Joshi and Zwolinski [5] have observed that the variation of the
polymerization entropy bears some relationship to the molar volume
change of polymerization and may be empirically fitted to a linear
relationship, approximately given by the equation AS = 25 + \ R In
(Vi/V-j), where Vx and V2 are the unit volumes of monomer and poly-
mer, respectively.
Entropy of stereoregularity is discussed later.

B. Determination of Entropy of Polymerization


The entropy change can be calculated from the measurement of
equilibrium monomer concentration, from the entropies of monomer
and polymer, from the ratio of the kinetic frequency factors Ap and
Aj, and from empirical and statistical calculations [1-3].
1. Measurement of Equilibrium Monomer Concentration. Equa-
tion (13) gives the equilibrium monomer concentration as a function
of the ceiling temperature. Instead of saying that a monomer at con-
centration [Mx] has a ceiling temperature T c , it will frequently be
convenient to reverse the viewpoint and say that at temperature T the
monomer concentration in equilibrium with the long-chain polymer
is [ M j e , where t = T c and [ M j e = [Mj = exp {(AH°- T AS°)/RT}.
Measurements of equilibrium monomer concentrations at two
different temperatures, at least, permit evaluation of AH and AS
from the slope and intercept, respectively, of the plot of In [M] vs
l / T as related by Eq. (13).
2. Enthalpy Change and One Equilibrium Temperature. This
method is based on the relation AS = AH/T, and it is only limited
by the accuracy of the measurements of heats of polymerization and
ceiling temperatures.
3. Entropies of Monomer and Polymer. Standard entropies of
monomers can be determined from specific and latent heat data, and
the standard entropies of some gaseous monomers could be derived
from spectroscopic data.
The estimation of entropies from specific heat data is based on
the third law of thermodynamics which states that the entropy of
perfect crystal is zero at 0°K. Since polymer molecules have a re-
sidual entropy at 0°K, polymer molecules cannot achieve perfect
crystallinity at the absolute zero.
Temperley [6] points out that the four most relevant considera-
tions regarding this problem are that: (a) a polymer is a mixture of
chains of many different lengths; (b) since crystallization i n a polymer
THERMODYNAMICS OF POLYMERIZATION. I 325

is never complete, at least some of the polymer is present in the


amorphous or disordered state—the communal entropy problem;
(c) because of "freedom at the joints," any one chain may assume a
wide variety of configurations in space; and (d) in a copolymer the
number of ways in which the two sets of monomer can be linked to
form a chain is infinitely great.
Temperley [6] concluded that for a hydrocarbon-like polymer the
contribution to S from (a) and (b) is no more than 3- 5 times k per
polymer chain and the contribution from (c) is no more than k/5 per
link; for a rubberlike polymer the entropy associated with chain con-
figurations is no more than k per monomer unit. The general validity
of these conclusions is borne out by studies on natural rubber and
isotactic and atactic polypropylene, which indicate that the residual
entropy of these polymers is less than 1 cal/deg/mole.
Data have become available on the entropies of the following poly-
mers by the specific heat method: polythene [7,8], isotactic and
atactic polypropylene [9], isotactic polybut-1-ene [9,10], polyiso-
butene [11], isotactic and atactic polystyrene [9], cis- and trans-1,4-
polybutadiene [12], polyisoprene [13], polyvinyl alcohol [14], poly-
(vinylidene chloride) [15], polymethyl methacrylate [16], polyoxy-
methylene [17,18], polytetrafluoroethylene [19], poly-3,3-bis(chloro-
methyl) oxacyclobutane [20], propene, but-1-ene and hex-1-ene poly-
sulfones [21], polycarbonate [22], ethylene-propylene copolymers [23],
and poly(-4-methyl-l-pentene) [24]. These results are summarized
in Table 1.4.
•The entropy of polymerization can be obtained by subtracting the
entropy of the monomer from that of the polymer. All available re-
sults obtained by this method are summarized in Table 1.5.
Second-law entropies of polymerization can be obtained from mea-
surements of equilibrium pressure (or concentration) or ceiling tem-
perature as discussed in Sec. II. B.I. The only systems for which
both second- and third-law entropies of polymerization are available
are shown in Table 1.6. To obtain a direct comparison between a
second- and third-law entropy of polymerization, it is essential that
specific heat measurements on a given polymer should be combined
with monomer-polymer equilibrium studies on the same polymer.
From such data it would be possible to estimate the residual entropy
of the polymer. However, in no cases were the second- and third-
law entropies of polymerization evaluated for the same monomer-
polymer system. Thus, although the small difference between the
two values for styrene may be due to the possession of a residual
entropy by the polymer, it may also result from different tacticities
or crystallinities of the polymers used in the two sets of measure-
ments [30].
326 HIDEO SAWADA

Table 1.4
Entropies of Linear Polymers (cal/deg/mole, 25°C)
Polymer % Crystallinity Ref.

Polyethylene 0 14.8 [9]


58 12.4
79 11.7 m
19]
100 11.0 [9]
Isotactic polypropylene 0 19.4 [9]
48 17.2 [9]
100 14.8 [9]
Atactic polypropylene 0 19.6 [9]
16 18.9 [9]
100 15.0 [9]
Isotactic polybut-1-ene 0 27.6 [9]
44 24.6 [9]
100 20.7 [9]
Isotactic polystyrene 0 32.0 [9]
43 31.4 [9]
100 30.7 [9]
Atactic polystyrene 0 32.2 [9]
100 30.9 [9]
Polyisobutene 0 22.9 [11]
cis-1,4- Polybutadiene - 27.5 [12]
trans-1,4- Polybutadiene - 23.2a [12]
Polyisoprene - 30.6 [13]
Polymethyl methacrylate - 34.2 [16]
Polyoxymethylene Highly 10.3 [17,18]
crystalline 10.6
Polytetrafluoroethylene Highly 24.1b [19]
crystalline
Poly-3,3-bis(chloro-
methyl) oxacyclobutane - 45.4 [20]
Propene polysulfone - 32.1 [21]
But-1-ene polysulfone - 37.2 [21]
Hex-1-ene polysulfone - 52.8 [21]
Poly(4,4'-dioxydiphenyl-
2,2-propane carbonate) — 75.3 [22]
Ethylene-propylene copoly-
mer (containing 31 mole
% propylene) - 15.1 [23]
Isotactic poly(4-methyl-
1-pentene) 65 36.2 [24]
Polyvinylidene chloride — 17.0c • [15]
a
At 317°K.
b
At 280°K.
c
At 200°K.
THERMODYNAMICS OF POLYMERIZATION. I 327

Table 1.5
Entropies of Polymerization (cal/deg/mole, 25°C)

Standard
Monomer states -AS Ref.

Ethylene gc 41.5 17]


Ethylene gc 37.7 [V]
Propylene gc 49.0 [9]
Propylene gc 44.4 [9]
Propylene gc 45.7 [25]
But-1-ene gc 52.3 [9]
But-1-ene gc 45.4 [9]
Isobutene lc 28.8 [11]
Styrene lc 24.9 [3]
Styrene lc 25.2 [9]
Styrene • lc 26.7 [26]
Styrene lc 26.5 [9]
Buta-l,3-diene to the
cis-l,4-polymer lc 21.2 [3]
Buta-l,3-diene to the
• cis-l,4-polymer lc 20.1 [12]
Isoprene lc 24.2 [3]
Methyl methacrylate lc 9.6a [27]
Formaldehyde gc 41.8 [17]
Trioxane gc 37.2 [28]
Tetrafluoroethylene lc 26.81> [3]
3,3-Bis (chloromethyl) lc 19.9 [20]
Oxacyclobutane
Vinylidene chloride gc 21.2= [15]
4- Methyl- 1-pentene gc 51.7 [24]

"At 210°K.
b
Atl98°K.
= At 200°K.

4. Ratio of the Frequency Factors of the Propagation and Deprop-


agation Reactions. By comparing Eqs. (4) and (10), Dainton and Ivin
[2] state that the entropy of polymerization is given by

AS° = R In (Ap/Ad) (15)

where Ap and A^ are the frequency factors for the propagation and
depolymerization reactions, respectively. Assuming that A^ is 1013
sec' 1 , i.e., the reaction is a normal unimolecular one, they have com-
328 HIDEO SAWADA

Table 1.6
Comparison of Entropy Data for Polymerization 8 (cal/deg/mole, 25°C)

Standard Second-law Third-law


Monomer states entropies entropies Ref.

Styrene lo 25.0 25.8 [3,9,26]


Formaldehyde gc 41.8 30.7 [17,29]

"Values from Ref. [30].

puted the change in entropy in passing fro"m the liquid monomer to a


molar solution of the polymer in the monomer. For vinyl acetate
this figure is-26.2 cal/deg mole (Ap = 1.65 x 108 mole"1 sec"1)
and for styrene, -27.9 cal/deg mole (Ap = 1.0 x 106 mole"1 sec"1).
5. Empirical Estimation. The methods available for evaluating
AS|g empirically are in principle the same as those for the estima-
tion of AH|g. Thus, Jessup [31] has utilized a relation connecting
AGf for the formation of the mono-olefins and the number of carbon
atoms to calculate AG™ for the polymerization of ethylene. The
values of AS|g may then be obtained in the usual way. The calculated
entropies of polymerization show better agreement with the experi-
mentally found values than do the predicted heats of polymerization,
because the entropy change in polymerization is not as sensitive to
monomer structure.
6. Statistical Calculations. The factors that determine the magni-
tude of AS|g may be appreciated by considering the translational,
rotational, and vibrational contributions separately [2].
For the dimerization, the translational contribution ASgg(t) is
given by

AS gg(t) = - | J | R In 298.1-2.298 J + | R In (2m) - | R In (m)


= -11/967 - 3.4305 log (m) cal/deg mole (16)
where m is the molecular weight of the monomer. The external rota-
tional contribution AS| g ( r ) is given by
(T T T U/S n-l/2
AS
ee(r) = - 5 - 8 0 + 2 - 2 8 7 log UAiB
*c^ - 4.575 \og-?— (17)

where a denotes symmetry number, IA moment of inertia in atomic


THERMODYNAMICS OF POLYMERIZATION. I 329

mass A2 units, and the subscripts 1 and 2 refer to monomer and dimer,
respectively. The vibrational contribution AS° (r) is

(18)

where a = 4.827 x 10' 3 cm, v denotes a vibration frequency, and the


summations are taken over all vibrations of the two molecules con-
cerned.
Owing to the lack of values of the moment of inertia, symmetry
numbers and vibration frequencies of the commonly used monomers
and their dimers, expression (16), (17), and (18), do not seem to be
useful; Moreover, all internal rotational contributions to the entro-
pies of the monomer and dimer have been included as vibrations in
[18]. Nevertheless, analysis of AS|g for the dimerization in this
way has advantages. Thus, a consideration of the vibration fre-
quencies assigned to aliphatic hydrocarbons by Pitzer [32] imme-
diately indicates that the total vibrational, including internal rota-
tional contributions, of most monomers or dimers at room tempera-
ture is only about 10-15% of the total entropy. Also there will be
little change in external rotational entropy on dimerization for those
systems in which a! =<r2 = 1 and

In this case the main loss of entropy is due to the replacement of the
translational entropy of the monomer by one-half the translational
entropy of the dimer.
For polymerization, Dainton and Ivin [2] obtain

ASgg(r) + ASgg(v)

• (25.98 + 6.861 log m) + —^— log n

2
-287 (I A I B Ic)n +
(I A IBIC>I

+ ? £ (An) - R E (AX) (19)


330 HIDEO SAWADA

where RA is the contribution to the entropy of one vibration and y


and z denote the number of such terms appropriate to polymer and
monomer, respectively; it will be noted that

When n is very large AS° = -S° of the monomer plus the sum of the
vibrational and internal rotational entropy of the monomer unit in the
polymer. In ethylene, AS| g = -34.1. whereas S | for the monomer
= 52.5 cal/deg mole, and we, therefore, conclude that this latter
quantity is of the order of 18 cal/deg mole.
The same problem has been treated by Evans and Baxendale [33].
Their treatment leads to a value for the frequency factors of propa-
gation and termination given by

log A p = 1 - log (fm / f p ) - log w + log kT/h (20)

log At = 1 - log wr + log kT/h (21)

and the ratio of the two frequency factors is, thus, found to be

A p /A t = fp/fm (22)

The entropy of polymerization ASJS is given by

ASJS = R{1 - log i/o- - log (f m /f p )} (23)

where v is the coordination number of the space lattice, cr is the


symmetry number of the polymer molecules, and fm and fp are the
respective partition functions of the free monomer units and of the
monomer segments in the polymer. This expression refers to bulk
polymerization and a standard state of 1 mole/liter. The same dif-
ficulty arises, namely, that the vibration frequencies of the monomer
units in the polymer chain are unknown, and, therefore, fp and ASjs
cannot be evaluated.
C. Free Energy Changes of Polymerization
Since
AG° = A H 0 - T AS D

it is easy to calculate AG° at a given temperature if AH" and AS° are


known. The results of such calculations for all cases in which the
THERMODYNAMICS OF POLYMERIZATION. I 331

necessary data are available are shown in Table 1.7. We get for
AG° of a polymerization reaction at equilibrium,

AG° = - R T In K a (24)

where K a is the thermodynamic equilibrium constant of polymerization.


Under the equilibrium conditions the rates of forward and reverse
reactions occurring at all active centers M*, M* M , etc., may be
equated:

p =kdf) [M*] (25)

For a high-molecular-weight polymer

£ tMn] = S [Mn]
n n+1
and, hence,

(26
• ^ = K a =[M e r >

where [M] e is the equilibrium monomer concentration.


If polymer solutions were ideal, the equilibrium condition may be
established at the ceiling temperature T c . Thus,

AG° = - R T C In K a = RTC In [M]e


= AH°-T C AS° < 27 >

Alternatively, this may be written

AH
T P
• c AS^+ R l n [ M ] e
and
[M] e = exp(AG 0 /RT c ) (28)

Equation (29),known as the van't Hoff reaction isobar,defines the


temperature coefficient of K a in terms of the heat of reaction AH° and
the temperature T:
3 l n K a _ AH°
3T RT2 (29)
332 HIDEO SAWADA

Table 1.7
Heats, Entropies, Free Energy (changes , and Equilibrium Constants
for Polymerization at 25°C

Standard
Monomer state -AH" -AS' -AG» Ka Ref.

Tetrafluoro-
ethylene 1c 37 26.8 29 3.0 x IO22 [34]
Ethylene gg 22.2 34.0 12.1 - [35]
lc 25.9 41.5 13.5 — [7,36]
Vinyl acetate lc 21.2 26.2 13.4 7.1 x 1010 [34]
Propylene lc 19.5 27.8 2.0 x io 9 [34]
lc 24.9 49.0 10.3 - [9,36]
gg 20.7 39.9 8.8 - [35]
Butadiene lc 17.6 20.5 11.5 2.9 x 109 [34]
Butene-1 gg 19.1 39.8 7.2 - [35]
lc 19.0 26.8 - 1.3 x io 9 [34]
lc 20.7* 29.8 a 12.8 a - [10]
Isoprene lc 17.9 24.2 10.7 7.5 x 108 [34]
Styrene lc 16.7 25.0 9.2 5.0 x io 7 [34]
gg 18.1 35.5 7.5 - [35]
Methyl
methacrylate lc 13.2 28.0 4.8 3.5 x 104 [34]
lc 13.8 9.6 11.8 — [27]
Ethyl
methacrylate lc 13.8 29.7 4.9 [34]
Isobutylene lc 12.9 28.8 4.3 5.0 x 104 [34]
lc 17.2 3.9 [36]
a-Methyl styrene gg 8.1 35 -2.3 [35]
lc 8.4 24.8 1.0 4.1 x 10 [34]
Vinylidene
'chloride lc 14.4 b 21.17 b 10.2 b [15]
Ac enaphthy lene gg 20.4 45 7.0 [35]
Formaldehyde gc 17.2 43.8 4.1 [37]
gc 13.2 41.8 0.7 [17]
gc 13.2 41.8 0.9 [36]
Tetrahydrofuran lc 5.3 18 0.0 1.0 x 10 [34]
Trioxane gc 3.83 12.39 0.14 [18]
a
At 265°K.
b
At 200'K.
THERMODYNAMICS OF POLYMERIZATION. I 333

For polymerization reactions, when K a = [M]"1, Eq. (29) may be


written

9T RT2
Polymer solutions are, however, rarely ideal, and equations more
complex than Eq. (28) are usually needed to derive heats and entro-
pies of polymerization from ceiling temperatures. Bywater [38] and
Small [39] have given expressions for nonideal solutions.
Considering the polymerization of 1 mole of monomer to 1 base
mole of polymer, the change per mole in free energy is given by

AG = AG° + RT In (a p /a m ) • (31)

where a p is the activity per base mole of the polymer and a m that of
the monomer; AG° is the standard free energy change with the com-
pounds at unit activity. This corresponds to AGic of Dainton's nota-
tion if the activities of monomer and polymer in the pure liquid and
solid states, respectively, are taken as unity. If the standard state
is defined as a 1 M solution, the relation becomes

AG = AGSS + RT In (a p /a m ) + RT In (a m /a p ) (32)

where a p and a m are the activities in the standard 1 M solution and


AGSS is the change in free energy with reactants and products in
1 M solution.
Since polymer solutions do not behave ideally, it is necessary to
modify the above equations by relating activities to concentrations
using the Flory-Huggins-type equations. In order to make the equa-
tionsusable, it is necessary to use simplifying assumptions. Thus,
for this purpose, it is assumed that the ju has the same value for the
monomer-polymer and solvent-polymer interactions and the heat of
mixing of solvent with monomer will be neglected.
With these assumptions, the equations relating activity and volume
fraction 6 become formally equivalent to the simple Flory-Huggins
equations [40]:

In a m = In <£m + (1 - l/x)(£p + M4>P (33)

In a p = (l/x) In <f>p + (l/x - 1)(1 - 6p) + M(1 - <f>p)2 (34)

where x is the ratio of the molar volumes of polymer and solvent,


but $ m + $ p is no longer equal to unity as in the simple case. Graph-
334 HIDEO SAWADA

ical plots of these functions in the concentration ranges of a large


excess of solvent show that these equations reduce with sufficient
accuracy to

In a m = In 0 m ; In a p = /i - 1 (35)

Equations (35) are a good approximation for x > 100 and give a rea-
sonable approximation for x values as low as 10. Equations (31) and
(32) become, on substitution of Eq. (35),

AG = AGlc + RT In Ox - 1) - RT In 0 m (36)
AG = AGSS + RT In (l/[M]) (37)

where [M] is the molar concentration of monomer.


At equilibrium, AG = 0 and concentrations become equilibrium
concentrations, hence,

-AG l c /RT = (M - 1) - In <& (38)

/ = ln

Equation (39) is identical with Eq. (28), when polymer solutions are
ideal.
The two AG/RT values in the case of polymerization of methyl
methacrylate have been plotted against reciprocal temperature in
Fig. 1.2. From the slopes, values of AHjc = -13.4 kcal/mole and
AHSS =-12.9 kcal/mole are obtained [38].
Next the heats of mixing of monomer, solvent, and polymer are
taken into account [41]. Three /i parameters are retained: /xSp,
Mmp> a n ^ Msm c o r r e s P ° n d i n g to the pairs solvent-polymer, mono-
mer-polymer, and solvent-monomer.
At equilibrium,
-AG/RT = ln a p - ln a
m
where ap and a m are activities of polymer and monomer, respectively.
The activities refer to liquid monomer and solid polymer as their
respective standard states:

-AG l c /RT = (fj sp - Msm)0s ~ 1 ~ l n <Pm + Mmp(<£m ~ 4>p)


(40)
THERMODYNAMICS OF POLYMERIZATION. I 335

26

IO'/T
Fig. 1.2. Plots of AG/RT values vs reciprocal temperature. Upper line,
AGic; lower line, AGSS [38].

where the <p terms are volume fractions at equilibrium of solvent


(s), monomer (m), and polymer (p).
The molar free energies for mixing of monomer (AG^j) and of
polymer (AG0) with sufficient solvent to form a molar solution are

- cbp'1 In In

AG*/RT = - 0 s ) " 1 In 4
-AG S S /RT = In {4>hUm) + (M s p - Msm)(0s - 4>%) + Mmp^m" <Pp)
where 4> is the volume fraction at equilibrium and <p* that in a 1 M
solution.
336 HIDEO SAWADA

When fiSp = /ism and if either Mmp = 0 or <£m and <£p tend to
zero, Eq. (41) reduces t o - A G s s / R T = In ($m/(f>m)- If 4>m/<Pm ~
[M]e, the laws of ideal solution apply, i.e., AG SS /RT = -In [M]e.
For experiments carried out in a highly dilute solution, Eq. (41)
is given by the simplified equation,
+
-AG S S /RT = In ( 0 m / 0 m ) + (Msp- MsmK<£s ~ 4>s> Mmp^m (42)
Deviations from ideal behavior arise from three causes: (1) 0 m is
not proportional to the monomer concentration in which case
$m/$rh * [MJe! (2) the interaction of solvent with monomer is
different from that with polymer, i.e., Msm * f-'sp' $) the monomer-
polymer interaction is not negligible, i.e., jimp ^ 0-
The inequality jnSm ^ f*sp accounts for the dependence of [M]e on
the nature of the solvent. The effect of solvent is not eliminated even
at an extremely low equilibrium monomer concentration. The last
term, Mmp(0m~ 0p)> i s unimportant if </>m and $ p are small. How-
ever, it becomes significant in concentrated polymer solution.
Ivin and Leonard [42] considered the effect of a soluble polymer
in a liquid monomer on the free energy of polymerization. Consider
a homogeneous equilibrium mixture containing a volume fraction <£m
of monomer and 0 p of polymer of degree of polymerization n. The
free energy of polymerization in an equilibrium mixture is zero and
may be expressed as the sum of three terms [-AGj, the free energy
change for the removal of 1 mole of liquid monomer from the mix-
ture; AGic, the free energy of polymerization of 1 mole of liquid
monomer to 1 base-mole (l/n_moles) of amorphous (liquid or non-
glassy solid) polymer; and AG2, the free energy change for the addi-
tion of 1 base-mole of polymer to the mixture]:

-AG^ + AGlc + AG~ = 0


By inserting the appropriate expressions for AG^ and AG2 from the
Flory-Huggins expression, we obtain

AGlc = RT{ln ^ m - (In <£p)/n + 1 - l/n + x(<f>p ~ <£m)} (43)


where x is the polymer-monomer interaction parameter. If n is suf-
ficiently large and $ p is not too small, this can be reduced to
AGlc = RT{ln 0 m + 1 + x(0p - 0 m )} (44)
Equation (44) is applicable to a system such as liquid tetrahydrofuran-

J
THERMODYNAMICS OF POLYMERIZATION. I 337^

dissolved polytetrahydrofuran. Taking the polymer-monomer inter-


action parameter as 0.3, Ivin and Leonard [42] obtained AHjc =
-3.0 kcal/mole and ASjc = -9.8 cal/deg/mole in the case of tetra-
hydrofuran.

References
[1] F. S. Dainton and K. J. Ivin, Nature, 162, 705 (1948).
. [2] F. S. Dainton and K. J. Ivin, Trans. Faraday Soc, 46, 331 (1950).
[3] F. S. Dainton and K. J. Ivin, Quart. Rev., 12, 61 (1958).
[4] K. J. Ivin, in Polymer Handbook (J. Brandrup and E. M. Immergut,
eds.), Wiley-Interscience, New York, 1966, pp. 11-363.
[5] R. M. Joshi and B. J. Zwolinski, in Vinyl Polymerization (G. E. Ham,
ed.), Vol. 1, Dekker, New York, 1967, p. 494.
[6] H. N. V. Temperley, J. Res.Natl. Bur. Std., 56, 55 (1956).
[7] F. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia, Polymer,
3, 277 (1962).
[8] R. W. Warfield and M. C. Petree, Makromol. Chem., 51, 113 (1962).
[9] ' F. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia, Polymer,
3, 286 (1962).
[10] R. W. Warrield and M. C. Petree, J.Polymer Set., A-2, 5, 791 (1967).
[11] G. T. Furukawa and M. L. Reilly, J. Res. Natl. Bur. Std., 56, 285 (1956).
[12] F. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia, Polymer,
3, 297 (1967).
[13] N. Bekkedahl and H. Matheson, J. Res. Natl. Bur. Std., 15, 503 (1935).
[14] R. \V. Warfield and R. Brown, Kolloid-Z., 185, 63 (1962).
[15] R. W. Warfield and M. C. Petree,,/. Polymer Set., 4, 532 (1966).
[16] T. P. Melia, Polymer, 3, 317 (1962).
[17] F. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia, Polymer,
3, 263 (1962).
[18] T. P. Melia, D. Bailey, and A. Tyson, J.Appl. Chem., 17, 15 (1967).
[19] G. T. Furukawa, R. E. McCoskey, and G. J. King, J. Res. Natl. Bur.
Std.,49, 273 (1952).
[20] F. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia, Polymer,
3, 271 (1962).
[21] F. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia, Polymer,
3, 310 (1962).
[22] F. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia, Polymer,
3, 316 (1962).
[23] T. P. Melia, G. A. Clegg, and A. Tyson, Makromol. Chem., 112, 84
(1968).
[24] T. P. Melia.and A. Tyson, Makromol. Chem., 109, 87 (1967).
[25] D. R. Gee and T. P. Melia, Makromol. Chem., 116, 122 (1968).
[26] R. W. Warfield and M. C. Petree, J. Polymer Sci., 55, 497 (1961).
[27] R. W. Warfield and M. C. Petree, J . Polymer Sci., Al, 1701 (1963).
338 HIDEO SAWADA

[28] G. A. Clegg, T. P. Melia, and A. Tyson, Polymer, 9, 75 (1968).


[29] F. S. Dainton, K. G. Ivin, and D. A. G. Walmsley, Trans. Faraday Soc.
56, 1784 (1960).
[30] T. P. Melia, J.Appl. Chem., 14, 461 (1964).
[31] R. S. Jessup, J. Chem.Phys., 16, 661 (1948).
[32] K. S. Pitzer, J. Chem.Phys., 5, 469 (1937); 8, 711 (1940).
[33] J. H. Baxendale and A. G. Evans, Trans. Faraday Soc, 43, 210 (1947).
[34] R. M. Joshi and B. J. Zwolinski, in Vinyl Polymerization (G. E. Ham,
ed.), Vol. 1, Dekker, New York, 1967, Chap. 8.
[35] R. M. Joshi and B. J. Zwolinski, Macromolecules, 1, 25 (1968).
[36] G. S. Parks and H. P. Mosher, J. Polymer Sci., Al, 1979 (1963).
[37] T. P. Melia, Polymer, 7, 640 (1966).
[38] S. Bywater, Trans. Faraday Soc, 51, 1267 (1955).
[39] D. A. Small, Trans. Faraday Soc, 49, 441 (1953).
[40] P. J. Flory, Principles of Polymer Chemistry, Cornell Univ. Press,
Ithaca, N. Y., 1953, Chap. 12.
[41] S. Bywater, Makromol. Chem.,52, 120 (1962).
[42] K. J. Ivin and J. Leonard, Polymer, 6, 621 (1965).
Chapter 2. Heat of Polymerization

I. GENERAL ASPECTS 339


A. Breaking a Multiple Bond 340
B. Resonance 340
C. Steric Strain 341
II. VARIATIONS IN HEATS OF POLYMERIZATION 341
A. Steric Strain in the Polymer 343
B. Conjugation and Hyperconjugation 346
C. Hydrogen Bond and Solvation 347
III. EMPIRICAL ESTIMATION OF HEAT OF POLYMER-
IZATION 347
REFERENCES 349
APPENDIX 350

I. GENERAL ASPECTS
The heats of polymerization for a number of monomers are shown
in the Appendix. Since AHp = AE - P AV, if the volume change can
be neglected, the heat of polymerization becomes equivalent to the
change in internal energy of the molecule. The internal energy may
be reduced in a number of ways, for example, by a release of steric
strain or by a loss of IT- electron energy.
Here, we consider three factors: (a) breaking a multiple bond,
(b) resonance, and (c) opening a ring under strain.

339
340 HIDEO SAWADA

A. Breaking a Multiple Bond


Polymerization is often accompanied by the opening of a double
bond C=C and the formation of a single bond C—C. A double bond
has a higher energy than a single bond, and conversion of a ;r-bonded
• molecule to a a -bonded framework results in an overall decrease in
bond energy. Therefore, the heats of polymerization of vinyl poly-
mers are about 20 kcal, which is the normal difference between the
average bond energy of a C=C bond (146 kcal) in a monomer and the
sum of the two C—C single bonds (83 x 2 = 166 kcal) formed in the
polymer.
If no other factors predominate, therefore, polymerization of alde-
hyde or ketones to polyoxymethylenes should result in an increase in
enthalpy (Table 2.1) and becomes thermodynamically impossible be-
cause of the obvious negative entropy involved in any association
process. However, it may be seen that the predicted values of AHgg,
based on the more refined Cox scheme [1] for the compounds men-
tioned above, are definitely about 5 kcal negative.
Table 2.1
Decrease in Bond Energy on Polymerization

Predicted heat
Bond energies, of polymerization,
kcal/mole kcal/mole

C=C (145.8) to -C-C-(82.6) -20


C=O (176) to -C—O—(85.5) +5
C=N (147) to -C—N-(72.8) +1.4
C=N (212.6) to -C=N-(147) -7.2
C=S (128) to -C-S-(65) -2
S=O (104) to -S-O-(55.5) -7

B. Resonance
It should be pointed out that increased delocalization or resonance
stabilization will lower the internal energy and the enthalpy of a
molecule. For certain polymerizations, resonance may be entirely
responsible for the free energy change in polymerization. If, there-
fore, the delocalization per repeating unit is different in the mono-
mers and polymers, then those species with the greater resonance
stability will be favored in the equilibrium if no other influences
predominate.
THERMODYNAMICS OF POLYMERIZATION. I 341

If the polymer is appreciably stabilized by delocalization and the


monomer is not, then AHp will be negative, and the position of the
final equilibrium will strongly favor the polymer rather than the
'monomer. The bond energies of the C=N, C=N, and C—N groups
are simple multiples of each other (in the ratio 3:2:1) so that the
opening of these multiple bonds in an addition polymerization will
not result in a lower enthalpy and, hence, addition polymerization
will not occur. However, the polymerization products -4c(R)=N4-
are stabilized to a greater or lesser extent by delocalization, and
— AH|g should be 10 kcal/mole. This phenomenon is termed reso-
nance-induced polymerization [2].
The hypothetical head-to-head addition of cyclohexene, according
to the equation

2n I I *- -{(CH2)4—CH=CH—CH=CH-(CH2)44h

results in resonance in the polymer which was not present in the


monomer. If the resonance energy of this system is similar to
butadiene, 3.5 kcal/-fCH=CH—CH=CH-}- unit, then-AHg g of the
polymerization would be 1.8 kcal/mole of cyclohexene. Although the
ceiling temperature of such a polymerization will be low, there
should be some temperature below which the polymerization would
be thermodynamically possible. This polymerization is also an
example of resonance-induced polymerization.
C. Steric Strain
Since steric strain must raise the internal energy, interactions of
this type will raise the enthalpy of the ring monomer relative to the
linear polymer where the substituent groups become widely spaced.
The heat of polymerization, therefore, provides a direct measure of
the strain energy in the monomer ring [3-5].
A more detailed discussion is given in the next section.

II. VARIATIONS IN HEATS OF POLYMERIZATION


Variations in the heats of polymerization of various monomers
arise mainly from the following causes: (a) steric strain in the
polymer; (b) differences in stabilization energy in monomer and
polymer due to conjugation or hyperconjugation; and (c) hydrogen
bond and solvation. Some of the factors that affect the heat of polym-
erization are summarized in an internal energy versus reaction co-
ordinate diagram in Fig. 2.1.
CO
to

>
O
CONJUGATION
UJ A HYPERCONJUSATION
z H Y D R O G E N BOND
UJ SOLVATION
RESONANCE
DELOCAUZATION
z
DC STERIC STRAIN
UJ

MONOMER
STERIC HINDRANCE

POLYMER

REACTION COORDINATE m

Fig. 2.1 Internal energy vs reaction coordinate diagram illustrating the factors that affect the heat of polymerization. o
THERMODYNAMICS OF POLYMERIZATION. I 343

A. Steric Strain in the Polymer


Steric strain in the polymer tends to make AHp less negative as
a result of bond stretching, bond-angle deformation, or interaction
between nonbonded atoms. This factor, steric hindrance, is pre-
sumably the reason for the low heats of polymerization of the
a,a-disubstituted ethylenes, such as vinylidene chloride, isobutene,
a-methyl styrene, and methyl methacrylate (Table 2.2), because the
crowding together of the groups is greater in the polymer than in the
monomer.
Table 2.2
Variations in Heats of Polymerization
Heat of polymerization, Ceiling temperature
Monomer kcal/mole in bulk, °C

Styrene 16 235
a-Methyl styrene 7 61
Methyl aery late 20
Methyl methacrylate 13 164
Formaldehyde 13 126
Acetoaldehyde 0 -31
Acetone -6
Vinyl chloride 32
Vinylidene chloride 14
Ethylene 26 407
Propylene 21 300
Isobutene 17 50

Polyoxymethylenes are much more sensitive to side-group steric


hindrance, and side groups larger than hydrogen appear to lower T c
to room temperature or below. From the earlier discussion, it is
clear that if AHp changes from negative toward zero, the ceiling
temperature will be reduced accordingly. Since side-group steric
hindrance would be expected to decrease the rotational entropy of
the polymer, ASp would become even more negative in the presence
of bulky side groups. This would serve to lower the ceiling tempera-
ture further with increasing steric repulsion of the side groups.
Evans and Polanyi [6] showed that, when models were constructed
using van der Waals radii for the substituent group, these groups
came to within distances smaller than their normal van der Waals
344 HIDEO SAWADA

separation in the final product. This could lead to repulsion energy


between substituent groups, and, hence, to a higher energy of the
final product.
Inspection of models shows that substituents on alternate carbon
atoms of a'—C—C—C— chain interfere with each other to a greater
extent than those located on adjacent atoms. The construction of
scale models of a,/3-disubstituted olefin polymers, such as poly-
(butene-2) is easier than the construction of scale models of poly-
mers with two substituents alternating on the same carbon as in
polyisobutylene. This factor accounts for the lower heat of polym-
erization of a,a-disubstituted olefins as compared with that of
analogous ot,/3-disubstituted olefins. The hypothetical polymer,
poly(butene-2), must have two modes of conformation—the isotactic
and the syndiotactic—because every carbon in the chain —<5(CH3)H—
is asymmetric. The isotactic chain is rather strained and difficult
to construct even in the helical configuration, but the syndiotactic
is almost completely strain-free.
From the written formulas for polymers of a,a-disubstituted
monomers, greater steric hindrance might have been expected in
the head-to-head or tail-to-tail structure than in the head-to-tail
structure. However, the construction of scale models of the head-to-
head or tail-to-tail structure is easier than the construction of
scale models of the head-to-tail structure. Therefore, the alternate
head-to-head or tail-to-tail polymer should be comparatively free
from steric repulsions between substituents.
As shown in schematically in Fig. 2.2, this is also true in the
case of the activated complexes. A polymer radical in which the odd
electron is on the —CH2— group [(I) and (II) in Fig. 2.2] will, there-
fore, prefer to add onto the —CH2 end rather than the —CXY end (n).
Similarly, a polymer radical with the odd electron on the CXY group
will prefer to add onto the —CXY end of the monomer (IV) rather
than onto the —CH2 end (HI). Hence, the alternate head-to-head or
tail-to-tail polymer tend to be preferred if there is much steric
hindrance. In both the acrylate and methacrylate series the change
from methyl to ethyl to n-butyl leads to a steady increase in the heat
of polymerization.
McCurdy and Laidler [7] have shown that in these series the in-
creasing steric hindrance leads to an increase in the proportion of
the alternate head-to-head or tail-to-tail polymer, so that in the
polymers there is actually less hindrance with the more bulky sub-
stituents than with the smaller ones. In agreement with this, Marvel
and Cowan [8] have shown that the monomers
THERMODYNAMICS OF POLYMERIZATION. I 345

INITIAL STATE ACTIVATED STATE

4 /Y X

v
H .H
H Y x

X "< H
H
A
X Y H H

V .
A A
X Y X Y
- i<"/(
X Y X Y
""

vV
P: // vv
A '! H H XY

Fig. 2.2 Initial and activated states in polymerization [7].

Cl Br
CH2=C and CH2=C

where considerable steric effects are expected, lead predominantly


to the alternate head-to-head or tail-to-tail polymer.
The magnitude of the energy of steric repulsion cannot be calcu-
lated, and the difference between observed and calculated heats of
polymerization may be considered to represent the energy of steric
repulsion between neighboring groups within the polymer chain [9].
For monosubstituted units the difference is small, only for styrene
does it definitely exceed 1 kcal/mole. Heats of polymerization of
346 HIDEO SAWADA

disubstituted monomers are 3-9 kcal/mole lower (in magnitude) than


the calculated values, indicating severe compression of chain con-
stituents.
It is difficult to calculate the increase in AHp due to steric repul-
sions. However, the general principles on which quantitative calcula-
tions of steric effect for simple molecules can be performed are out-
lined in a recent monograph [10]. Attempts have been made to calcu-
late potential functions for organic polymers with the use of appre-
ciable approximations [11], but an accurate solution for polymer
molecules is a difficult task.
B. Conjugation and Hyperconjugation
Differences in stabilization energy in monomer and polymer due
to conjugation or hyperconjugation lower the heat of polymerization
to an extent equal to the resonance energy of the monomer, which
may be evaluated from the difference in heats of hydrogenation of
ethylene and the vinyl compound.
Conjugation to the ethylenic bond makes an appreciable contribu-
tion to the stabilization of styrene and its ring-substituted deriva-
tives, which keeps the heat of polymerization rather low (-AHp =
16.0-16.5 kcal/mole). Similar stabilization of the monomer state by
conjugation is apparent in butadiene or isoprene (about 3 kcal), vinyl
pyridine (2.5-4 kcal), acrylonitrile, and others. Conjugation with a
carboxyl or a carbonyl group in acrylic acid, acrylic esters, acro-
lein, and methyl vinyl ketone also seem to lower the AHic of these
monomers.
When the substituent is nonconjugating, as it is in vinyl acetate
and higher vinyl ethers, the AHp values would not be much altered.
In a series of a-methyl-substituted monomers, such as propylene
and ot-methyl styrene, the contribution of hyperconjugation in stabil-
izing the monomer is apparent.
It is rather difficult to account for the particularly high heats of
polymerization (exceeding that of ethylene itself) of the derivatives
in which there is a strongly electronegative group, such as vinyl
chloride (-AHic = 22.9 kcal), nitroethylene (-AH lc = 21.7 kcal),
vinylidene fluoride (-AHi c = 3 1 kcal), and tetrafluoroethylene
(—AHic = 37 kcal). These increased heats of polymerization may
be due to a reduction, on polymerization, of repulsion between non-
bonding electrons associated with the electronegative groups. Alter-
natively, they may arise because of stabilization of the polymer
caused by a degree of ionic covalent resonance.
In a-methyl styrene, all the factors lowering the AHp have com-
bined; namely, the resonance of the benzene ring, the —CH3 hyper-
THERMODYNAMICS OF POLYMERIZATION. I 347

conjugation stabilizing the monomer, and the a,a-disubstitution de-


stabilizing the polymer. In vinylidene chloride the electronegativity
of the chlorine substitution has caused the compensating effect of
raising the value, since otherwise the chlorine substituent, being
almost the same size as the methyl, should have lowered the AHp
to 13 kcal as for isobutene.
C. Hydrogen Bond and Solvation
It must be emphasized that since hydrogen-bonding and solvation
effects are small, the explanations in this section are necessarily
tentative. The hydrogen atom can form hydrogen bonds of O • • • H—O
or N—H • • -O type. The hydrogen-bonding sites in free monomer
molecules, more rigidly fixed than among polymer chains, would
lead to higher intermolecular association in the monomer, thus
lowering of the heat of polymerization. This point was recognized
and stressed by Joshi [12].
Hydrogen-bonded monomers include acrylic acid (-AHic = 16.0kcal),
methacrylic acid (-AHic = lO.lkcal), acrylamide (-AHSS= 14.4 kcal
in benzene and 13.8kcalin hexane),and methacrylamide (-AHSS =
8.4 kcal in benzene). If a sufficiently dilute aqueous or alcoholic
solution of a monomer is polymerized, hydrogen-bonding influences
may be appreciably reduced or eliminated almost completely. The
AHss (water or methanol) for all these hydrogen-bonded monomers
have been found normal in relation to structure and were uniformly
higher by about 3 to 5 kcal than those of the associated liquid state.
2-Hydroxyethyl methacrylate and 2-hydroxypropyl methacrylate
both exhibit lower heats of polymerization than do the monomers
having no hydroxyl groups. The difference seems too large to ex-
plain by steric hindrance. The number of hydroxyl groups solvated
in the polymer is generally less than in the monomer. The heats of
solvation of alcohols are in the range from 2.5 to 4 kcal/mole, so
that the effects observed could be explained if only part of the solva-
tion shell is removed on polymerization [7].
Factors, such as heats of emulsification of the monomers or heats
of wetting of the polymers, might also play a part.

III. EMPIRICAL ESTIMATION OF HEAT OF POLYMERIZATION


Heats of polymerization may be either experimentally determined
by each of three distinct methods or calculated by empirical methods.
Details of the experimental techniques, such as combustion methods,
direct-reaction calorimetry, and the thermodynamic equilibrium
techniques, have been fully treated in a recent review [13].
348 HIDEO SAWADA

All empirical methods of estimating heats of polymerization of


vinyl compounds have been based on empirical relations between
the heat of formation and structure of low molecular hydrocarbons.
Thus, using the relations deduced by Prosen et al. [14], Flory [15]
obtained values of AH|g = —22 to —24 kcal/mole for a number of
simple olefins. Jessup [16] has computed values of AH° and AG°
for ethylene polymerization over a range of temperatures and
chain lengths.
Heat of polymerization (assuming no steric hindrance in the poly-
mer) may also be predicted from the parallel reaction of hydrogena-
tion, which involves a change from trigonal to tetrahedral hybridiza-
tion of the carbon atoms, as in polymerization.
Flory [9] has considered the following steps:

CH 2 =CXY CH3—CH2XY
AH,

(A)

The heat of polymerization is given by

. AHp=AHh+AHe

If X = Y = H, AHe is 10.41 kcal. The heat of hydrogenation of ethyl-


ene is -32.73 kcal at 25°C; hence, the heat of polymerization for
the unsubstituted ethylene AH g g is - 2 2 . 3 kcal/mole. If X is an alkyl
group and Y = H, the reference compound (A) is a straight-chain
hydrocarbon, and the second process above consists in its transfor-
mation to a branched unit (B) having a single substituent. Hence,
AH e is 9.6 kcal. If both X and Y are alkyl groups, the reference com-
pound is a single-branched hydrocarbon, and it is converted to a
double-branched structural unit in the hypothetical dehydrogenation
process. So, AHe is 9.2 kcal. The heats of polymerization of the
various hydrocarbon monomers would, therefore, be about 9 to 11
kcal less negative than of the corresponding hydrogenation.
Recently Joshi and his associates [17] extended the Somayajulu-
Zwolinski generalized bond-energy scheme to 15 olefinic polymer
structures for estimating their enthalpies of formation and enthal-
pies of polymerization. The agreement with available experimental
data is quite satisfactory, provided that allowance is made for the
structural energy differences arising from tacticity of the a-olefin
polymers.
THERMODYNAMICS OF POLYMERIZATION. I 349

References
[1] J. D. Cox, Tetrahedron, 18, 1337 (1962).
12] R. J. Orr, Polymer, 5, 187 (1964).
[3] F. S. Dainton, T. R. E. Deilin, and P. A. Small, Trans. Faraday Soc,
51, 1710 (1955).
[4] P. A. Small, Trans. Faraday Soc, 51, 1717 (1955).
[5] F. S. Dainton, K. J. Ivin, and D. A. G. Walmsley, Trans. Faraday Soc,
56, 1784 (1960).
[6] A. G. Evans and M. Polanyi, Nature, 152, 738 (1943).
[7] K. G. McCurdy and K. J. Laidler, Can.J. Chem., 42, 818 (1964).
[8] C. S. Marvel and J. C. Cowan, j.Am. Chem. Soc, 61, 3156 (1939).
[9] P. J. Flory, Principles of Polymer Chemistry, Cornell Univ. Press,
Ithaca, N.Y., 1953, pp. 246-256.
[10] F. H. Westheimer, in Steric Effects in Organic Chemistry (M. S. New-
man, ed.), Wiley.New York, 1956, Chap. 12.
[11] P. DeSantis, E. Giglio, A. M. Liquori, and A. Ripamonti, J. Polymer
Sci.,A1, 1383 (1963).
[12] R. M. Joshi, J. Polymer Sci., 60, s56 (1962).
[13] R. M. Joshi and B. J. Zwolinski, in Vinyl Polymerization (G. E. Ham,
ed.), Vol. 1, Dekker, New York, 1967, Chap. 8.
[14] E. J. Prosen, W. H. Johnson, and F. D. Rossini, J. Res.Natl. Bur. Std.,
37, 51 (1946).
[15] P. J. Flory, J.Am. Chem. Soc, 59, 241 (1937).
[16] R. S. J e s s u p . J . Chem.Phys., 16, 661 (1948).
[17] R. M. Joshi, B. J. Zwolinski, and C. W. Hayes, Macromolecules, 1,
30 (1968).
350 HIDEO SAWADA

Appendix: Heats of Polymerization of Various Monomers a-

Stan-
dard -AH, Tempera-
Monomer states kcal/mole ture, °C Ref.

Acenaphthylene ss 23.5 26.9 b


ss 24.0 26.9 b
cc 19.6 26.9 b
ss 17.6 74.5 c
ss 16.9 74.5 c
Acetaldehyde lc 0 25 d
Acetone 1c -6 25 d
Acrolein lc 19.1 74.5 c
Acrylamide ss 19.8 (water) 26.9 b
ss 14.1 (benzene) 74.5 e
ss 16.9 (acetone) 74.5 e
sc 19.5 (water) 74.5 e
sc 13.8 (hexane) 74.5 e
Acrylic acid lc 16.0 74.5 c
ss 18.5 (water) 20 f
ss 18.4 (water) 25 g
sc 17.6 (benzene) 74.5 c
sc 17.2 (carbon 74.5 c
tetrachloride)
sc 17.8 (hexane) 74.5 c
Acrylic acid, n-butyl ester Is 18.5 74.5 e
lc 19.1 25 g
Acrylic acid, ethyl ester lc 18.8 25 g
lc 18.6 74.5 h
Acrylic acid, methyl ester lc 18.7 76.8 i
lc 18.8 74.5 h
lc 18.6 25 g
ss 20.2 20 f
ss 19.4 74.5 h
Acrylonitrile lc 17.3 76.8 i

lc 18.3 25 j
lc 18.3 74.5 k
sc 18.5 74.5 k
Allyl chloride Is 18.5 74.5 c

Biphenyl, p-isopropenyl ss 8.1 -15 k


1,3-Butadiene gg 17.4 25 1

gg 18.7 25 1
lc 17.6 25 in

J
THERMODYNAMICS OF POLYMERIZATION. I 351

Appendix (continued)

Stan-
dard -AH, Tempera-
Monomer states kcal/mole ture, °C Ref.
1-Butene gg 20.7 25 n
lc 20.0 25 n
cis-2-Butene gg 19.1 25 n
lc 17.9 25 n
trans- 2-Butene gg 18.1 25 n
lc 17.0 25 n
Chloral gc 17 25 o
lc 9 50 o
sc 8 50 o
Chloroprene lc 16.2 61.3 p
Ethylene gg 22.35 25 q
go 25.4 25 q
gc 25.9 25 r
gc 24.2 25 V
gc 25.5 25 t

gc 25.9 25 u
Ethylene, tetrafluoro- gc 41.5 25 V

37 25 w
gg
lc 39 25 X

Ethylene, nitro- lc 21.7 25 y


Formaldehyde gc 17.2 25 z
gc 12.2 25 aa

gc 13 25 aa
gc 13.2 25 u

gc 16.3 80 bb
25 n
1-Heptene gg 20.6
1-Hexene lc 19.8 25 n

Isobutene gc 17.2 25 u
lc 11.5 25 r,cc
ss 12.8 25 cc
ss 12.9 -50 dd
lc 12.6 25 t

Isoprene gg 16.9 25 1
lc 17.9 25 ee
Is 17 74.5 ff
Is 15.7 34.6 gg
Itaconate, dimethyl- ss 14.5 26.9 b

Maleic anhydride Is 14 74.5 ff


352 HIDEO SAWADA

Appendix (continued)

Stan-
dard -AH, Tempera-
Monomer states kcal/mole ture," C Ref.

ss hh
Maleimide 16.1 (chlorobenzene) 74.5
ss 21.4 (dioxane) 74.5 hh
ss 21.2 (acetonitrile) 74.5 hh
ss 20.9 (dimethyl
hh
formamide) 74.5
Methacrolein lc 15.6 74.5 c
Methacrylamide ss 13.4 (water) 74.5 e
ss 10.2 (chloroform) 74.5 e
ss 9.4 (acetone) 74.5 e
ss 8.4 (benzene) 74.5 e

Methacrylic acid lc 10.1 74.5 c

ss 15.8 (water) 20 f
ss 13.5 (water) 25 g
sc 13.6 (methanol) 74.5 c

Methacrylic acid,
lc ii
benzyl ester 13.4 76.8
Methacrylic acid,
n-butyl ester lc 13.9 25 g
lc 13.5 76.8 ii

lc 13.7 74.5 h

Is 14.3 26.9 b
Methacrylic acid,
tert-butyl ester Is 13.0 26.9
Methacrylic acid,
cyclohexyl ester lc 12.2 76.8 j]
Is 12.7 26.9 b
Methacrylic acid,
p-ethoxyethyl ester Is 14.8 26.9 b
lc 13.7 74.5 h
Methacrylic acid,
ethyl ester lc 14.4 120 kk
lc 13.8 25 g
lc 14.2 74.5 h

Is 13.8 26.9 b
Methacrylic acid,
n-hexyl ester lc 14.0 25 g
Is 14.4 26.9 b
Methacrylic acid,
/3 -hydroxyethyl ester lc 11.9 25
THERMODYNAMICS OF POLYMERIZATION. I 353

Appendix (continued)

Stan-
dard -AH, Tempera-
Monomer states kcal/mole ture, °C Ref.
Methacrylic acid,
/3-hydroxypropyl ester lc 12.1 25 g
Methacrylic acid,
isobutyl ester 1c 14.3 74.5 h
Methacrylic acid,
isopropyl ester lc 14.3 74.5 h
Methacrylic acid,
methyl ester lc 13.9 76.8 P
lc 13.4 130
lc 12.9 (water) 20 f
lc 13.6 (water) 25 g
lc 13.3 74.5 d,h
Is 13.8 26.9 b
ss 12.9 130 II j mm
ss 14.0 (acetonitrile) 74.5 h
ss 13.7 74.5 h
ss 14.0 (hexane) 74.5 h
Is 13.1 24
Methacrylic acid,
phenyl ester lc 12.3 76.8 ii
Methacrylic acid,
n-propyl ester lc 13.7 74.5 h
Methacrylonitrile lc 13.5 74.5
ss 15.3 130
Methyl vinyl ketone lc 17.7 74.5
Naphthalene,
2-isopropenyl- ss 8.7 -5
cis-2-Pentene gg 19.2 25
trans-2-Pentene gg 18.1 25
Propylene sc 16.5 -78
gc 24.9 25
gg 20.7 25
lc 20.1 25
Styrene gg 17.8 25
lc 16.7 25
lc 16.1 76.8
lc 16.5 74.5
lc 16.4 26.9 b
lc 17.4 127 mm ,uu
354 HIDEO SAWADA

Appendix (continued)

Stan-
dard -AH, Tempera-
Monomer states kcal/mole ture, °C Ref.

Styrene (cont'd.) Is 17.5 25 ss


Is 17.4 26.9 b
Is 17.7 26.9 b
ss 15.9 -60 dd
Is 16.8 24 oo
Styrene, o-ehloro- 1c 16.4 76.8 tt

Styrene, p-chloro- lc 16.0 76.8 tt


Styrene, 2,5-dichloro- 1c 16.5 76.8 tt
Styrene, ar-ethyl lc 16.3 76.8 tt

Styrene, a-methyl lc 8.4 25 w


lc 8.2 -20 rrnij ww

ss 8.0 -20 irnijWw

ss 7.0 30 XX

ss 8.5 -20 k
ss 8.0 -20 WW

Styrene, 2,4,6-trimethyl lc 16.7 26.9 b


Vinyl acetate lc 21.3 76.8 i
Is 20.0 25 yy
Is 21.4 25 zz
Is 21.6 24 OO

lc 21.0 74.5 e
ss 21.5 74.5 e
ss 20.7 74.5 e
ss 20.5 74.5 e
Vinyl benzoate lc 20.2 74.5 h
Vinyl n-butyl ether lc 14.4 40-60 aa.i
N-Vinylcarbazole sc 15.2 74.5 c
Vinyl chloride gc 31.5 25 bbb , ccc
gc 30.7 25 bbb,ddd
lc 17 25 1
lc 26.7 25 bbb
Is 22.9 74.5 eee
Vinylidene chloride lc 14.4 76.8 i
lc 18.0 25 bbb
lc 17.5 74.5 h
gc 24.0 25 bbb
lc 17.7 74.5 c
Vinylidene fluoride gc 35 25 fff
Vinyl propionate lc 20.5 74.5 e,h
THERMODYNAMICS OF POLYMERIZATION. 1 355

Appendix (continued)

Stan-
dard -AH, Tempera-
Monomer states kcal/mole ture, °C Ref.

2-Vinyl pyridine lc 17.1 74.5 c


Is 18.0 74.5 c
sc 17.6 74.5 c
4-Vinyl pyridine lc 18.7 74.5 e
ss 18.7 74.5 e
a
Monomers are listed alphabetically; cyclic monomers are listed in
Chapter 5 of this review, to be published in this journal in the future.
b
F . S. Dainton, K. J. Ivin, and D. A. G. Walmsley, Trans. Faraday Soc,
56, 1784 (1960).
C
R. M. Joshi, Makromol. Chem., 55, 35 (1962).
d
V. A. Kargin, V. A. Kabanov, V. P. Zubov, and I. M. Papisov,
Dokl.Akad. Nauk S.S.S.R., 134, 1098 (1960).
e
R. M. Joshi, J. Polymer Set., 56, 313 (1962).
{A. G. Evans and E. Tyrrall, J. Polymer Set., 2, 387 (1947).
8 Reference [7]
h
R.M. Joshi, Makromol. Chem., 66, 114 (1963).
1
L. K. J. Tong and W. O. Kenyon, J.Am. Chem Soc, 69, 2245 (1947).
J J. H. Baxendale and G. W. Madaras, J. Polymer Sci., 19, 171 (1956).
k
H. Hopff and H. Lussi, Makromol. Chem., 62, 31 (1963).
ID. E. Roberts, J. Res. Natl. Bur. Std., 44, 221 (1950).
m
R. A. Nelson, R. S. Jessup, and D. E. Roberts, J . Res. Natl. Bur. Std.,
48, 275 (1952)
r
F. S. Dainton, J. Diaper, K. J. Ivin, and D. R. Sheard, Trans. Faraday
Soc, 53, 1269 (1957)
°W. K. Busfield and E. Whalley, Trans. Faraday Soc,59, 679 (1963).
PS. Ekegren, S. Ohrn, K. Granath, and P. O. Kinell, Ada Chem. Scand.,
4, 126 (1950).
qR. S. Jessup, j . Chem. Phys.,16, 661 (1948).
r
J . W. Richardson and G. S. Parks, J. Am. Chem. Soc, 61, 3545 (1939).
S
F. A. Quinn and L. Mandelkern, J.Am. Chem.Soc, 80, 3178 (1958).
l
G. S. Parks and J. R. Mosely, J. Chem. Phys., 17. 691 (1949).
U
G. S. Parks and H. P. Mosher, J. Polymer Sci., Al, 1979 (1963).
V
D. W. Scott, W. D. Good, and G. Waddington, J. Am. Chem. Soc, 77,
245 (1955).
W
\V. M. D. Bryant, J. Polymer Set., 56, 277 (1962).
X
F. S. Dainton and K. J. Ivin, in Experimental Thermochemistry
(H. A. Skinner, ed.), Wiley-Interscience, New York, 1962, p. 251.
VJ. Grodzinski, A. Katchalski, and D. Vofsi, Makromol. Chem., 44-46,
594 (1961).
356 HIDEO SAWADA

Appendix (continued)
J
J . B. Thompson, Formaldehyde, 3rd ed., Reinhold, New York, 1964,
p. 180.
aa
F . S. Dainton, K. J. Ivin, and D. A. G. Walmsley, Trans. Faraday Soc.,
55, 61 (1959).
bb
Y. Iwasa and T. Imoto, J. Chem. Soc. Japan, Pure Chem. Sect.,
84, 29 (1963).
CC
A. G. Evans and M. Polanyi, Nature, 152, 738 (1943).
dd
R. H. Biddulph, W. R. Longworth, J. Penfold, P. H. Plesch, and
P. P. Rutherfold, Polymer, 1, 521 (1960).
ee
R. S. Jessup and A. D. Cummings, J. Res.Natl. Bur. Std., 13, 357
(1934).
ff
R. M. Joshi, Makromol. Chem., 55, 35 (1962).
gg
A. A. Korotkov and E. N. Marandzheva, Russ.J.Phys. Chem. (English
lransl.),Z7, 135 (1963).
"'R. M. Joshi, Makromol. Chem., 62, 140 (1963).
11
L. K. Tong and W. O. Kenyon,,/. Am. Chem. Soc, 68, 1355 (1946).
JJ
L. K. J. Tong and W. O. Kenyon, J. Am. Chem. Soc, 68, 1355 (1946).
• ^ R . E. Cook and K. J. Ivin, Trans. Faraday Soc, 53, 1132 (1957).
11
S. Bywater, Trans. Faraday Soc, 51, 1267 (1955).
111111
S. Bywater, Makromol. Chem., 52, 120 (1962).
™ K. J. Ivin, Trans. Faraday Soc, 51, 1273 (1955).
00
H. Miyama, Bull. Chem. Soc Japan, 29, 711 (1956).
PP
S . Bywater, Can.J. Chem., 35, 552 (1957).
qq
C. M. Fontana and G. A. Kidder, J. Am. Chem. Soc, 70, 3745 (1948).
" F. S. Dainton and K. J. Ivin, Trans. Faraday Soc, 46, 331 (1950).
" D. E. Roberts, W. W. Walton, and R. S. Jessup,./. Res. Natl. Bur. Std.,
38, 627 (1947).
" L. K. J. Tong and W. O. Kenyon, J. Am. Chem. Soc, 69, 1402 (1947).
UU
S. Bywater and D. J. Worsfold, J. Polymer Sci., 58, 571 (1962).
w
D. E. Roberts and R. S. Jessup, J. Res. Natl. Bur. Std., 46, 11 (1962).
WW
D. J. Worsfold and S. Bywater, J. Polymer Sci., 26, 299 (1957).
XX
H. W. McCormick, J. Polymer Sci., 25, 448 (1957).
^ W . I. Bengough, Trans. Faraday Soc, 54, 54 (1958).
" W. I. Bengough, Trans. Faraday Soc. 54, 1560 (1958).
" a M . F. Shostakovskii and I. F. Bogdanov, J.Appl. Chem. USSR, 15, 249
(1942).
bbb
G. C. Sinke and D. R. Stull, J.Phys. Chem., 62, 397 (1958).
ccc
J. R. Lacher, E. E. Merz, E. Bohmfalk, and J. D. Park, J. Phys. Chem.,
60, 492 (1956).
ddd
J. R. Lacher, H. B. Gottlieb, and J. D. Park, Trans. Faraday Soc,
58, 2348 (1962).
eee
R . M. Joshi, Indian J. Chem., 2, 125 (1964).
111
W. D. Good, J. L. Lacina, B. L. De Prater, and J. P. McCullough,
J. Phys. Chem., 68, 579 (1964).
Chapter 3. Thermodynamics of Radical Polymerization

I. GENERAL ASPECTS 358


A. Energetics of Radical Polymerization 358
B. Degree of Polymerization 359
C. Activation Energies of Elementary Reactions 361
II. GENERATION OF FREE RADICALS 361
III. PROPAGATION REACTION 366
A. The Polanyi Relation 366
B. Reactivity and Heat of Polymerization 368
C. Ceiling Temperature 371
IV. INTERACTION OF RADICALS 371
A. Combination and Disproportionation Reactions . . . . 371
B. Interaction of Small Hydrocarbon Radicals 372
C. Interaction of Large Hydrocarbon Radicals 373
D. Interaction of Some Large Radicals 374
V. FREE ENERGIES OF FORMATION O F POLYETHYL-
ENE AND POLYTETRAFLUOROETHYLENE 376
A. Free Energies of Polyethylene Synthesis 376
B. Free Energies of Polytetrafluoroethylene Synthesis . 381
' REFERENCES 385

357
358 HIDEO SAWADA

I. GENERAL ASPECTS
A. Energetics of Radical Polymerization
Radical polymerization mechanisms can be summarized as follows:
1. Initiation,
I — 2R- vx = 2kif[I]
R- + M — P1- vj =ki[R-][M]
2. Propagation,
P n -+ M - Pntl- v p =k p [P-][M]
3. Termination,
Pn' Pm" ~~* Pn + m v
td:
+ v
*n *m ^n+m tc
4. Chain transfer,
:
p • + M -* P n + P, • v

Let us represent graphically the state of the reacting system by


plotting its potential energy vs the reaction path. Then the energy
of activation of the process is determined by the height of the energy
barrier that the reacting system must surmount, as shown in Fig. 3.1.
If the reaction is exothermic, the energy of activation is simply the
height of the potential barrier E p . The difference between the activa-
tion energies of the backward and forward reactions is equal to the
heat of the forward process [1].

E d - E p = AHp (1)

The apparent activation energy E a obtained from the slope of a


plot of the logarithm of the rate vs l / T will be related to the individ-
ual activation energies as follows [2]:

E a = Ed/2 + (Ep - Et/2) (2)

in which Ed, E p , and Et are the energies of activation for decompo-


sition of initiator, propagation of polymerization, and termination of
polymerization, respectively.
THERMODYNAMICS OF POLYMERIZATION. I 359

ACTIVATED STATE /

A I
REACTANTS

PRODUCTS

REACTION COORDINATE

Fig. 3.1. Energy vs reaction coordinate diagram.

The activation energy for the spontaneous decomposition of ben-


zoyl peroxide is about 30 kcal/mole. For many polymerization r e -
actions, (Ep - Et/2) i s about 5 ~ 6 kcal/mole. The apparent activa-
tion energy for initiated polymerization i s , therefore, slightly
greater than 20 kcal/mole. This corresponds to a two- to three-fold
increase in rate for a 10°C temperature change.
B. Degree of Polymerization
When the degree of polymerization is large, the fraction of mono-
mer removed by steps other than the propagation step is negligible,
so that the rate of polymerization i s given by

(3)

and the degree of polymerization P n is given by the rate of removal


of monomer divided by the rate of formation of pairs of dead polymer
molecule ends, i.e.,

v + v +
td tc •m
k p [M] (4)
k t [ P - ] + k m [M]
where fy = + k{ C .
3 6 0 H I D E O SAWADA

If we assume steady-state conditions, [P •] = Vvj/Rt and VJ = vx =


jfp], The degree of polymerization is given by

The rate constants for the reactions can be expressed in an Arrhenius


form:

Therefore,

/ - E m + E D \ _,_ . / - E t / 2 - E./2 + E D \ ,„,


m
7. = a exp ( P I + b exp ( l / 1/
P) (6)
\ RT / \ RT /
where a = A m / A p , b = V2f[I]A1At/([M]Ap). By differentiating this
expression with respect to temperature, we obtain

_dz = -a(-Em+Ep) / - E m + E p \ _ b(-E t /2 - E,/2 + Ep)


dT RT2 \ RT / RT2
x exp ( - E t / 2 - E l / 2 +Ep\
V RT /
When Pfi has a maximum value the temperature is given by the re-
quirement that z be a minimum with respect to the change in T. Thus,
we find at dz/dT = 0

_ E m - E t /2 - Ex/2
a(E m - E p )
Rln
ib(E p -E t /2-E 1 /2)( (8)

Setting TM > 0, we get following relation:

. E m > Ep > E l±_ E i (9)

Hence, a maximum in the degree of polymerization vs tempera-


ture curve for homogeneous radical polymerization will be observed,
if we get the above relation. If the rate of initiation is constant, the
relation is expressed as follows:
E m > E p > Et/2 (10)
THERMODYNAMICS OF POLYMERIZATION. I 36i_

For thermal polymerization,

E i
E m > Ep > ^Et (11)

where Ej is the energy of activation for initiation.


The occurrence of a maximum in the molecular weight vs polym-
erization temperature curve for polyvinyl acetate and for polymethyl
methacrylate seems to demonstrate that the relation E m > E p > E^/2
can be accepted [3,4].
If transfer reactions are negligible, the temperature coefficient
of molecular weight will depend upon the initiation process. If an
initiator is used, it follows that

d In P n = E p - E t /2 - Ed/2
(12)
dT RT5

Since E p - Et/2 is 5-6 kcal and E$ is about 30 kcal, this quantity will
be negative and the molecular weight will decrease with increasing tem-
perature. The same is true in thermal polymerization. In photopolym-
erization in the absence of transfer, d In Pn/dT = (Ep - Ed/2)RT2,
which is positive. This is the only case in which molecular weight
increases with temperature.
C. Activation Energies of Elementary Reactions
.. Values of AHp, Ep, and Et for several polymerizations are shown
in Table 3.1. Except for the dienes, E p is near 7 kcal/mole for most
monomers, and Ej. varies from 3 to 5 kcal/mole. The frequency
factors Ap vary over a fifty-fold range, which suggests that steric
effects may be somewhat more important than the activation energy.
For example, Ap is much lower for methylmethacrylate with two
substituents on the same ethylenic carbon atom than it is for less
hindered monomers. As shown by the data of Table 3.1, Semenov [1]
has pointed out that the activation energy E p clearly increases even
when the heat of polymerization decreases. However, this relation
does not hold in the case of the data collected by Flory [2].

n . GENERATION OF FREE RADICALS


This is the type of reaction, which produces radicals, and its
occurrence is governed by the strengths of the bonds concerned. In
general, the thermal dissociation of a molecule can lead to the for-
mation of free radical fragments. The simplest type of such a reac-
ON

Table 3.1
Heats of Reaction and Activation Energies of Some Monomers

AH p , A E p , kcal t, kcal
P .
(x 10"T)
Monomer kcal/mole Flory [2] Semenov [1] Flory [2] Semenov [1] Flory [2]

Vinyl acetate 21 24 4.5 7.3 0 5.2

Methyl aery late 19 10 5.2 7.1 1.6 5

Styrene 16 2.2 G.9 7.8 2.8 2.4

Methyl
methacrylate 13 0.51 5.7 G.3 1.2 2.8

Butadiene 17.8 12 9.3

Isoprcne 17.3 12 9.8


m
o
THERMODYNAMICS OF POLYMERIZATION. I 363

tion is when only one bond is broken in thermal dissociation; for ex-
ample, the thermal dissociation of hexaphenylethane leads to two
triphenylmethyl radicals, thus,

For a split into radicals, the bond dissociation energy, or heat of


reaction, will be equal to the difference in the activation energies of
the forward and reverse reactions:

where Ex and E2 are the activation energies of the forward and re-
verse reactions, respectively, and D is the dissociation energy of
the Rx—R2 bond. The reverse reaction involves the recombination
of two radicals and, in general, E 2 £ 0 and, hence, Et = D. The acti-
vation energy E of the primary bond dissociation is equal to the
bond dissociation energy D, conveniently symbolized as D(R, - R2).
In this way bond dissociation energies have been measured and the
factors influencing these energies have been investigated [5-7]. The
dissociation energy of a particular bond type is not a constant but
depends upon the molecular environment of the bond. It must be em-
phasized that the thermochemical bond energy (or simply bond
energy) is a mean quantity and that the bond dissociation energy is
the energy required to break an individual bond.
Bond dissociation energies, or heat of the above reaction at 0°K,
have been expressed in terms of heats of formation Qf of molecules
and radicals:
Qf(R2-) - Qffo - R2) (13)
364 HIDEO SAWADA

For example, the dissociation energy for the CH3—Clbond,D(CH3—Cl),


may be calculated. Thus, for the reaction

CH3—Cl CH3 Cl

D(CH3-C1) = Qf(CH3-) + Qf(Cl •) - Qf(CH3-C1) (14)

Substituting the appropriate heats of formation, we obtain

D(CH3—Cl) = 32.0 + 29.0- (-19.6)

=i 80.6 kcal/mole (15)

Table 3.2 summarizes available data on bond dissociation energies


for molecules in organic free radical reactions (taken from the re-
view of Kerr [8]). It is seen that bond dissociation energies are in-
fluenced by the molecular environment. The delocalization energy of
the radical may confer stability on the radical and lead to a low bond
dissociation. Another factor which may influence bond dissociation
energies is the relief of steric strain as dissociation occurs. Repul-
sion forces between large groups in the molecule may give instability
to the molecule and lead to a low bond dissociation energy.

Table 3.2
Selected Bond Dissociation Energies [8] .

Bond D238, kcal Bond D298, kcal

CH-H 108 CH3-C1 84


CH2—H 104 CC13—Cl 73
CH 3 -H 104 CH3-I 56.3
CH2=CH-H 104 HO - O H 51
HO-H 119 CH3O—OCH3 36.1
HS-H 90 CH 3 COO-OOCCH 3 30
CH 3 S-H 88 C 2 H 5 COO-OOCC 2 H 5 30
CH 2 =CH 2 167 (CH 3 ) 2 N-N(CH 3 ) 2 42
(CH 3 ) 2 CH-CH(CH 3 ) 2 78 (CH 3 )NH-NH(CH 3 ) 47
CH3—CH3 88 •CH=CH-H 42
CH3CO— CH3 82 • CH2-H 108
CH3-OH 91 •O-H 101
C2H5—OH 91 •S-H 81
CH 3 O-CH 3 80
THERMODYNAMICS OF POLYMERIZATION. I 365

In general, D(C—H) values for radicals are low compared to those


of saturated molecules. The presence of an odd electron in a mole-
cule has a weakening effect on the surrounding bonds. This phenomenon
seems to account for the ease of radical decomposition, rearrange-
ment, and disproportionation reactions.
To give radicals by thermal dissociation at relatively low temper-
atures, a molecule must contain a weak valency bond such as that
between oxygen atoms in peroxides. A variety of such basic struc-
tures are known, involving chiefly C—C, N—N, O—O, and S—S bonds
and their combinations, and containing various substituent groups
that decrease the bond dissociation energies by resonance stabiliza-
tion of the resulting radicals. The basic requirement for the thermal
formation of free radicals at ordinary temperatures in the liquid
phase is a structure possessing a covalent bond with a dissociation
energy of 20-40 kcal. Many thermal sensitizers require an activa-
tion energy of roughly 30 kcal/mole for dissociation. As these radi-
cals are unstable and reactive, the activation energy of the primary
bond dissociation is equal to the bond dissociation energy.
The energies of activation for dissociation also depend upon the
nature of solvent—these can be attributed to differences between the
extents to which the initial and transition states are solvated.
The empirical Morse potential [9] for diatomic molecules, as
shown in Fig. 3.2, is given by

V=D0{l-exp[-a(r-re)]}2 (16)

in which a is an empirical constant, and r is the interatomic distance.


Equation (16) yields

V = Do when r = «

Since the minimum,

V= 0 when r = r e

Do is, thus, the energy of dissociation D plus the zero-point energy.


: In calculating Do, we first note that it is the maximum height of
the potential curve for large r. If a molecule had this amount of
energy, its atoms could just become independent of each other. There-
fore,

(Evib)max = D
o
366 HIDEO SAWADA

re r

Fig. 3.2 Morse potential for diatomic molecules.

Since the zero-point energy is roughly (£)h!/0l we now have


D
= ( E vib)max- (17)

where h is Planck's constant, and v0 is the fundamental vibrational


frequency.

HI. PROPAGATION REACTION


A. The Polanyi Relation
In Fig. 3.3 are shown two types of propagation reactions having
different heats of polymerization. Curve A corresponds to a reac-
tion such as AAACH2—CH2- + CH2=CH2, and curve B corresponds
to a reaction such as AAACH2—CHX- + CH=CHX, where X repre-
sents phenyl group. It is assumed for purposes of illustration that
the reactants have the same potential energy. The energy of activa-
tion and the heat of polymerization are indicated in the diagram. Be-
cause of resonance stabilization, the potential energy of the polysty-
rene radical is appreciably lower than that of the polyethylene radi-
cal. Curve B is lower than curve A. Therefore, the intersection of
the potential curves is shifted (only) vertically by a change of Q.
THERMODYNAMICS OF POLYMERIZATION. I 367

CO
a.
Mi

REACTION COORDINATE

Fig. 3.3. Intersections of potential curves for processes with heats of


reaction.

If the pairs of curves (A and B) have very similar shapes and if


the effects due to stabilization of the transition state are the same,
an increase in heat of polymerization by the amount AQ will cause a
smaller decrease in activation energy AE. Hence,

AE=-aAQ 0 < a < l (18)

where a is positive but less than 1, its magnitude depending upon the
shapes of A and B near the point of intersection.
Evans and Polanyi [10] were the first to find existence of the above
relation for exothermic reactions. This relation was proposed and
verified experimentally somewhat later using the reactions between
Na atoms and a homologous hydrocarbon series. Integration gives

E = A - aQ (19)

where A is a constant for a homologous series. This type of relation,


usually referred to as the Polanyi relation, is only applicable to re-
actions of radicals or atoms with a series of closely related com-
pounds.
368 HIDEO SAWADA

B. Reactivity and Heat of Polymerization


The reaction considered here may be of the following type:

H H H H
\ _ / II
R---C, — C« R—C,—Co—
/ l \ i i
H X H X
— rx r2— — rx r2—
During this reaction the following changes occur: (a) the interaction
between the 77 electrons of the double bond is broken and a o bond
formed between C1 and R, (b) the u bond between Cx and C2 is ex-
tended from a distance of 1.45 to 1.54 A, (c) the C—H links are
changed from trigonal to tetrahedral symmetry about the carbon
centers. Of these changes, (a) and (b) have been studied, and a
simple model involving the free electron on the radical R and the
ir electron on the C=C has been used. Evans, Gergely, and Seaman
[11] have computed such energy of the system as a function of the
distances rx and r 2 . Figure 3.4 shows the results obtained. On these
contour surfaces, the point I represents the energy and configuration
of the initial state of the system, and F, that of the final state. The
problem is to determine the factors influencing the height of the
activation energy barrier T. Two methods are available for the
treatment of such problems.
The main conclusions are as follows: (1) The activation energies
of such reactions are low, of the order of 3 to 10 kcal; (2) the weaker
the bond formed between the attacking radical and the carbon center
of the double bond, the higher the activation energy; and (3) the more
exothermic the reaction, the lower the activation energy of the pri-
mary radical attack.
In a reaction of the type
/w\CHY- + CH2=CHX * - /v^CHYCH2CHX •
the heat of reaction can be expressed as [10]
Q = Qo - R a + Rf - R ra <2°)
in which Qo is the heat of the reaction
/w\CH 2 - + CH2=CH2 »- /w\CH2—CH2—CH2 •

and Ra, Rf, and R ra are the resonance energies of the attacking radi-
cal, the radical formed, and the monomer, respectively.
THERMODYNAMICS OF POLYMERIZATION. I 369

20
A
^

o
F )
.6
T
. —
— . -
10
"1-5 20 25 30 35
Tc CH,. A

20
B

F V
* — T
( \ '

10
10 15 20 < 25 30
0H= r l . A

Fig. 3.4. Energy surfaces showing (A) the potential energy contour of the
system CH3- • •CH2=CH2- • -CH2—CH2—CH2— as a function of the inter-
nuclear distance and (B) the potential energy contour of the system
OH + CH2=CH2— HO—CH2—CH2— as a function of the internuclear
distance. The initial transition and final state of the system are represented
by the points I and F, respectively; T is the activation energy barrier [11].

The potential energy of initial state is expressed in the form

Uinit = - ( A c = c + Rm + Ra> (21)

where A c = c is the 77-bond dissociation energy of C=C linkage. The


potential energy of final state Ufinaj is
c-c + Rf) (22)

where A c _ c is the cr-bond dissociation energy of C—C linkage. Thus,


the difference of heat of reaction can be expressed as
370 HIDEO SAWADA

Q = u init ~ ufinal
= (2AC_C - A c = c ) - R a + Rf - R m

= Qo - R a + Rf - R m (23)

and, hence,

AQ = Q - Qo = - R a + Rf - R m (24)
The difference of activation energy is given by the Polanyi relation
as
AE = E - Eo = - k ( - R a + Rf - R m ) (25)
where k is a constant. The activation energy of an initiation or prop-
agation step should vary from one system to another according to
E = Eo - k(Rf - R a - R m ) (26)
where Eo is the activation energy for a reaction corresponding to QQ.
This treatment leads to the conclusions outlined below.
Let us compare activation energies of homopolymerization for
monomers Mx and M2, in which the attacking radical is similar to
the radical formed:
Ex = Eo + kR mi
and, similarly,
E2 = Eo + kR mz
Thus,
(E, - E2) = k(R mj - R mz ) . • (27)
These considerations are consistent with the empirical observa-
tion that the chief factor governing the reactivity of a vinyl monomer
is the extent to which the double bond is conjugated with other un-
saturated groups.
(mj styrene,k p = 176 [12]
(m2) methyl methacrylate,kp= 367 [12]
(m3) methyl acrylate,k p = 2090 [12]
(m4) vinyl acetate,k p = 3700 [12]
THERMODYNAMICS OF POLYMERIZATION. I 371_

Although values of R m are not known accurately, it is very likely that


they are in the sequence R m i > R m2 » Rm;j > R m4 . Thus, high re-
activity of the nonresonance-stabilized vinyl acetate radical, as com-
pared with the three "conjugated" radicals of styrene, methyl methac-
rylate, and methyl acrylate, is very evident.
C. Ceiling Temperature
In 1938, Snow and Frey [13] pointed out that copolymerization of
low alkanes and sulfur dioxide would not proceed above a certain tem-
perature, which they termed the "ceiling temperature." This ceiling
temperature was found to be a characteristic of the olefin and inde-
pendent of the catalyst system. There have since been reports in the
literature of polymerizations, mainly at high temperatures, which
show deviations from normal kinetics. They will not proceed to com-
plete conversion and produce a final equilibrium concentration of
monomer which is dependent only on the reaction temperature.
Bywater [14] studied the photosensitized polymerization of methyl
methacrylate in solution over the range 100-150°C. The reactions
did not proceed to completion; at each temperature, the final value
of the concentration of monomer was independent of the initial value
and corresponded to the concentration of monomer in equilibrium
with polymer. At 132.2°C, for example, the limiting value of the con-
centration of monomer was close to 0.3 mole/liter. For polymers
derived from vinyl monomers, the concentrations of monomer in
equilibrium with polymer are very low at ordinary temperatures;
for example, at 25'C the calculated values for polystyrene and poly-
methyl methacrylate are, respectively, 10"6 and 10~3 mole/liter.
However, at 0°C the equilibrium concentration of a-methyl styrene
is 0.76 mole/liter. For many systems, the equilibria cannot be e s -
tablished at higher temperatures because of side-reactions.
The polymerization of a-methyl styrene does not proceed to com-
pletion even at -40°C. For a long time it was thought that a-methyl
styrene could not be polymerized by a free-radical mechanism. How-
ever, in 1958, Lowry [15] proved that this was due to the low ceiling
temperature and that this monomer can be polymerized at low tem-
peratures.

IV. INTERACTION OF RADICALS


A. Combination and Disproportionation Reactions
Pairs of radicals can interact by either combination or dispropor-
tionation, and the relative importances of these two processes might
be afforded by a study of the thermodynamic relations of the system.
372 HIDEO SAWADA

As part of a study of this problem, it will be worth while to investi-


gate the changes.in enthalpy, entropy, and free energy accompanying
the alternative reactions.
Bevington [16] has studied this problem and has been able to make
deductions concerning the effect of temperature, concentration, and
molecular size on the relative importance of the two modes of ter-
mination.
B. Interaction of Small Hydrocarbon Radicals
The reactions possible for two ethyl radicals are

C2H4 + C2H8 (la)


2C2H5
n-C4H10 (lb)

The difference in the changes of free energy accompanying these r e -


actions is

(AGC-AGD) or (AHC - A H D ) - T ( A S C - ASD)

The subscripts C and D refer to combination and disproportionation,


respectively; (AHC - AHD) may be evaluated either from the heats
of combustion of ethylene, ethane, and n-butane or from their heats
of formation; the entropy term may be calculated either from the
standard entropies of formation. Similar calculations can be carried
out for the interaction of other radicals.
For 2 moles of ethyl radicals at 25°C and 1 atm reacting at con-
stant pressure, according to reactions (la) and (lb), we have

AHC - AHD = - 2 2 . 1 k c a l
and
ASC - ASD = -33.4 cal/deg

The entropy term must be corrected since an increase in entropy


occurs when 1 mole of ethane is mixed with 1 mole of ethylene each
at 1 atm to give a mixture also at 1 atm. This increase in entropy
is -R(N 1 In nj + N2 In n 2 ), where N refers to the number of moles of
a substance and n to its mole fraction in the mixture. In this case
Nj = N2 = 1 and nx = n2 = \ , so that AS m j x j n g is 2.76 cal/deg and
(ASC - ASD) is -36.2 cal/deg. The only effects on (AHC - AHD) are
those due to gas imperfections and they are neglected; the entropy
THERMODYNAMICS OF POLYMERIZATION. I 373

change (S2 - Sx) per mole for a change in pressure is R In ( P - J / P J ) .


Noting that there are 2 moles of disproportionation products and
only 1 of combination product, (ASC — ASD) is changed to—36.2 —
R In 106 or -63.6 cal/deg.
Combination is more exothermic than disproportionation, but the
entropy of the products of the latter reaction is greater than that of
the combined product; at 25° C and a pressure of 1(T8 atm, (AGC - AGD)
is - 3 . 1 kcal, indicating that the enthalpy term outweighs the entropy
term. If (AHC - AHD) and (ASC - ASD) do not vary appreciably with
temperatures, there must be a temperature at which (AGC - AGD)
is zero; for ethyl radicals this "crucial" temperature is 337°C if
the pressure is 1 atm, and 74°C if it is 10"6 atm.
C. Interaction of Large Hydrocarbon Radicals
The hypothetical case of long straight-chained hydrocarbon radi-
cals interacting at 25°C in the gas phase to give gaseous products
is considered first; this corresponds to mutual termination process
for the free radical polymerization of ethylene under these idealized
conditions. The interacting radicals are, for simplicity, considered
to be of the same size; the competing reactions are

RCH=CH 2 + RCH2CH3 (2a)

RCH2CH2CH2CH2R (2b)

where R is a long hydrocarbon chain.


There are two distinct methods for evaluating (AHC — AH D ),
namely, by extrapolation of data for hydrocarbons of low molecular
weight and by comparison with the heat of polymerization. Expres-
sions have been given [17] for heats of formation in kilocalories per
mole for gases at 1 atm and 25°C.
n-Paraffins: AH = -10.908 - 4.926n
' Olefins with terminal CH=CH 2 : AH = 19.592 - 4.926n
where n is the number of carbon atoms, supposed large. Therefore,

AHC - AHD = -10.908 - 4.926 x 2n + 10.908 + 4.926n - 19.592


+ 4.926n = - 1 9 . 6 kcal (28)

If differences in strength of C—H bonds of the various types are


neglected, disproportionation is equivalent to the breaking of one
C—C bond and the making of one C = C bond, and combination to the
374 HIDEO SAWADA

making of one C—C bond. In the propagation step of the polymeriza-


tion of ethylene, one C=C bond is broken and two C—C bonds are
made, therefore (AHC — AHD) ought to equal AHp.
The expression (-23.49n + 6.4) cal/deg has been given for the
molar entropy of formation of a gaseous normal paraffin with n car-
bon atoms in the standard state [18], and (-23.4n + 37.2) cal/deg
for the similar quantity for a 1-alkene [19]. These expressions lead
to (ASC - ASD) = -37.2 cal/deg, but, allowing for the increase in
entropy due to the mixing of the products of disproportionation, the
final value is approximately-40 cal/deg. If (AHC - AHD) is -21kcal
the "crucial" temperature is about 252°C; at 25°C, (AGC - AGD) is
9.1 kcal [16], whereas Bryant [20] gives -8.0 kcal, as discussed in
Section V.A below. Considering the difficulties the agreement is good .
The term (ASC — ASD) can be calculated by another method if suf-
ficient is known of the molecules concerned. Calculation of (ASC - ASD)
is resolved into three distinct parts, namely, estimates of the trans-
lational, vibrational, and rotational contributions; restricted internal
rotational contributions are included in the vibrational term, and
electronic contributions are neglected.
To summarize, for reaction of radicals of weight 1015 at 25°C,
(ASC - ASD) is equal to -40 cal/deg for the gas phase at 1 atm,
and —42 cal/deg for a solution in an ideal solvent when the volume
fraction of radicals is 10~5. Further, it may vary with the size of
the interacting radicals, the temperature, and the pressure in the
gas phase or the concentration in solution.
D. Interaction of Some Large Radicals
The calculation of the difference in free energy changes for the
reactions

AAACH=CH(C 6 H 5 ) + AA/\CH 2 CH 2 (C 6 H 5 ) (3a)

A^\CHCH(CH)CH(CH)CH (3b)

which are the alternatives for mutual termination in the styrene polym-
erization, is less satisfactory than the equivalent calculation for poly-
ethylene radicals. The enthalpy term cannot be set equal to the heat
of polymerization since combination of radicals is a head-to-head
reaction, whereas the growth of polymer is almost certainly head-to-
tail. The reactions used as models for the reactions of the large
radicals are given by
THERMODYNAMICS OF POLYMERIZATION. I 375

,CH 3 CH 2 C 8 H 5 CH 2 =CH(C 6 H 5 ) (4a)


2CH 3 CH(C 8 H 5 )
\
CH 3 CH(C 8 H 5 )CH(C e H 5 )CH 3 (4b)
The long polystyrene chains which play no part in the reactions are
replaced by hydrogen atoms. The quantity (AHC - AHD) is calculated
from the heats of combustion of the three products.
Here the heats of combustion are styrene, 1046 kcal/mole (liquid
at 18°C); ethyl benzene, 1091 kcal/mole (liquid at 18°C); and dibenzyl,
1811 kcal/mole (crystals at 18°C). Since the molar latent heat of
liquid dibenzyl is 4 kcal/mole, the heat of combustion of the liquid
at that temperature will be 1815 kcal/mole. Since the substitution
of a methyl group for hydrogen raises the heat of combustion by
156.5 kcal/mole, the heat of combustion of dimethyldibenzyl is taken
as 2128 kcal/mole. This gives (AHC - AHD) = -9.0 kcal/mole. The
calculations of (AHC - AHD) for a series of systems are summarized
in Table 3.3.
Table 3.3
Values of (AHC-AHD) for Some Monomeric Reactions [16]
Heats of
combus-
tions, (AHC-AHD),
Monomer Substances in model reactions kcal/mole kcal/mole

Acrylic CH 2 =CH-COOH 327.5


acid CH3—CH2 —COOH 365 -17.5
CH3CH(COOH) CH(COOH) CH3 675
Aci-yloni- CH2=CH-CN 420.5
trile CH3—CH2-CN 456.4 -16.0
CH3CH(CN)CH(CN)CH3 861
Metha- CH 2 =C(CH 3 )COOH 484
crylic CH 3 -CH(CH 3 )COOH 517.4 -11.5
acid (CH 3 ) 2 C (COOH) C (COOH) (CH 3 ) 2 990
Styrene CH 2 =CH-C 6 H5 1046
CH3 CH2 CgH5 1091 -9.0
CH3CH(C6H5) CH(C6H5) CH3 2128
a-Methyl CH 2 =C(CH 3 )-C 6 H 5 1203
styrene CH 3 -CH(CH 3 )-C 6 H 5 1247 -7.0
(CH 3 ) 2 C(C 6 H 5 )C(C 6 H 6 )(CH 3 ) 2 2443
376 HIDEO SAWADA

The entropy term cannot be estimated with certainty but it is


likely to be similar in magnitude for all radicals. If the monomer
has an a-methyl group, (ASC - ASD) is made slightly more negative
because disproportionation can occur in two ways, giving the un-
saturated product as either RCH=C(CH3)X or RCH2CX=CH2. At a
given temperature and concentration of radicals, the magnitude of
(AGC - AGD) varies appreciably from one system to another. Varia-
tions are due mainly to differences in (AHC — AHD), but for radicals
derived from monomers with a-methyl groups, the additional con-
tribution to ASD tends to make (AGC - AGD) more positive.

V. FREE ENERGIES OF FORMATION OF POLYETHYLENE


AND POLYTETRAFLUOROETHYLENE
A. Free Energies of Polyethylene Synthesis
1. Heats of Formation of Hydrocarbon Free Radicals. The calcu-
lation of the heats and entropies of free radicals has been carried
out for the case of hydrocarbons and applied to polyethylene synthe-
sis by Bryant [20]. The heat of formation of a hydrocarbon free
radical is most readily calculated from that of the parent hydrocar-
bon, in combination with the heat of association of atomic hydrogen
and the bond dissociation energy.
("t TT . f-f TT _i_ TT
C
nH(2n*2) c
n H (2nn) + H

H. .. 1 TT

The heat of formation of methyl free radicals i s , for example,


AH3
C + 2H2 — CH4 -17,889
CH4 — CH 3 - + H- +102,000
. H- — |H2 -52,089

C + f H2 -~ CH 3 - +32,000 cal/mole

Heats of formation of primary n-alkyl free radicals are given in


column two of Table 3.4 for radicals through n-dodecyl.
2. Molecular Entropies and Entropies of Formation of Hydrocar-
bon Free Radicals. The entropies of low-molecular-weight gaseous
THERMODYNAMICS OF POLYMERIZATION. I 377

Table 3.4
Free Energy of Formation of Gaseous n-Alkyl Free Radicals [20]

Free radical AH° AS" AG°

CH 3 . 32,022 -2.07 32,600


C2H5- 25,175 -21.64 31,600
n-C3H7 • 18,091 -46.94 32,100
n-C4H9 • 12,099 -69.04 32,700
n-C 5 H n • 6,411 -92.44 34,000
n-C 6 H, 3 . 1,151 -115.83 35,700
n-C,H,5. -3,775 -139.21 37,700
n-C 8 H 17 • -8,701 -162.60 39,800
n-C 9 H 19 • -13,627 -185.99 41,800
n-C 10 H 2 i- -18,553 -209.37 43,900
n-C u H 2 3 . -23,479 -232.76 45,900
n-C12H25- -28,405 -256.14 48,000

hydrocarbons can be calculated from spectroscopic data. Methods


for the calculation of the entropies of low-molecular-weight gaseous
hydrocarbons have been worked out in detail by Pitzer [21,22]. The
calculation of the molecular entropies of hydrocarbon free radicals
was based on a consideration of the following entropy contributions:
translational, rotational, electronic, vibrational, and restricted in-
ternal rotational. Translational entropies were calculated by means
of the Sackur-Tetrode equation expressed as follows:

S t r = 6.8635 log M + 25.9917 (29)

where M is the molecular weight. The rotational effect for the mole-
cule as a whole was calculated by the usual equation for a three-di-
mensional rigid rotator:
S r o t = 2.2878 log I I ' I " - 4.5757 log a + 284.6299 (30)
where I, I', and I" are the principal moments of inertia, and a is the
symmetry number. The moments of inertia were estimated approxi-
mately by a graphic analysis, assuming a planar arrangement of
groups about the trivalent carbon atom with normal bond lengths and
•120° bond angles.
Except in the case of methyl free radical, the vibrational entropy
at 298.2°K was assumed equal to that of the parent hydrocarbon. A
378 HIDEO SAWADA

similar assumption was made in the case of the restricted internal


rotational of ethyl, n-propyl, isopropyl, and tert-butyl radicals. The
vibrational entropy of the methyl free radical was calculated by means
of the Planck-Einstein approximation, assuming the following crude
frequency assignments: w = 900(1), 1400(2), and 3000(3), where the
unit is the wave number and the numerals in parentheses are the
multiplicities.
Using the molecular entropies the entropy of formation of methyl
free radical is, for example,
AS° =S°CH3. -(S' c +|S° H2 )
= 46.11- {1.36 +1(31.21)}
= -2.07 cal/deg mole
The entropies of formation of several hydrocarbon free radicals at
25°C are given in the third column of Table 3.4
3. Free Energies of Formation of Hydrocarbon Free Radicals.
In previous sections we have discussed methods for obtaining AH0
and AS0 at any given temperature. Since
AG° = AH0 - T AS0
it is easy to calculate AG° at a given temperature if AH° and AS° are
known.
The free energy of formation of CH3 • at 25°C is, thus,
AG° = AH0 - T AS0
= +32,000 - (298) x (-2.07)
= 32,600 cal/mole
In the last column of Table 3.4 are given free energies of formation
at 25°C for several hydrocarbon free radicals.
The magnitudes of the heats and free energies of formation of the
short-chain hydrocarbons and radicals are summarized in Fig. 3.5.
The free radicals have the largest positive AH0 and AG° values, with
the alkenes next, and finally the alkanes, values of which are mostly
on the negative side of zero. An interesting feature of the AG° curve
for the alkyl radicals is the near equality of the first four members
of the series. This is followed by a gradual slope upward toward
more positive values. The AH° curves, on the other hand, slope away
THERMODYNAMICS OF POLYMERIZATION. I 379

AG. n-ALKYL FREE RADICALS

IS
a
ui
z
u

-20
o

UJ -40
X

2 4 6 8
CARBON ATOMS PER MOLECULE

Fig. 3.5. Free energies and heats of formation of homologous hydrocarbon


series, including the alkyl free radicals [20].

fairly rapidly with curvature toward the lower positive and higher
negative values. In all three series the free energies become more
positive with increasing molecular weight, and the heats of forma-
tion more negative.
4. Thermodynamics of Polyethylene Synthesis. The average heat
and free energy involved in the polymerization of ethylene have been
calculated by Jessup [23] for products of increasing degrees of
polymerization, x, per mole of monomer. The results with x equal
. to infinity are as follows:
xC2H4(g) — (C2H4)x(g) AH° = -22.3kcal, AG° =-12.2 kcal
(31)
These values apply to the overall reaction by which polythene is
formed but tell very little about the individual reaction steps, since
the latter are believed to take place by a series of free radical
mechanism.
The three following reactions are examples of initiation by free
radicals:
380 HIDEO SAWADA

H- + C2H4 = C2H5- AH° = -39.4 kcal, AG° =-33.3 kcal (32)


CH3- + C2H4 = n-C3H7- AH0 =-27.0 kcal, AG° =-16.8 kcal (33)
C2H5- + C2H4 = n-C4H9- AH" = -25.6 kcal, AG° =-15.2 kcal (34)
The large-to-moderate negative free energy changes of reactions (32),
(33), and (34) indicate that they are thermodynamically favorable.
This type of reaction is also kinetically quite rapid and has a very
low activation energy. The highly reactive character of the hydrogen
atom, and of the methyl and ethyl free radicals, leaves little doubt
that these substances can be active initiators of polyethylene synthesis.
There is no fundamental difference between the initiation and chain
propagation steps. Both processes involve the addition of a free radi-
cal to ethylene to form a second free radical containing two additional
carbon atoms. The free energy change for chain propagation apparently
reaches a constant value with the formation of the n-octyl free radi-
cal.
n-C10H21 • + C2H4 = n-C12H25 •
AH0 =-22.3 kcal, AG° =-12.2 kcal (35)

The figures are, of course, the same as those for the overall polym-
erization step.
Chain transfer with hydrocarbons capable of forming secondary
or tertiary free radicals are often favorable:
n-C12H25- + C3H8 = n-C12H28 + (CH3)2CH •
AH° = -4.2 kcal, AG° = -3.9 kcal (36)
n-C12H25 • + (CH3)3CH = n-C12H2S + (CH3)3C •
AH0 = - 7 . 1 kcal, AG° = - 6 . 6 kcal (37)

Transfer reactions involving a primary free radical and a closely


related alkane have negligible heats and free energies:

n-C12H25 • + CI0H22 = n-C12H26 + n-C10H21 •


AH0 = 0.0 kcal, AG° = 0.0 kcal (38)

It is well known that chain transfer is important in the polyethylene


THERMODYNAMICS OF POLYMERIZATION. I 381

synthesis. Transfer to monomer is, however, unfavorable if the


vinyl radical is formed and doubtful if the ethyl radical results:

n-C12H25 • + C2H4 = n-C12H28 + C2H3 • (39)


0
AH = + 12.8 kcal, AG° = + 13.0 kcal
n-C12H25- + C2H4 = n-C12H24 + C2H6' (40)
AH0 = + 1.6 kcal, AG° = + 0.4 kcal
The most probable kind of chain transfer is that involving the re-
action of a growing primary free radical with a secondary hydrogen
from the side of a formed polymer chain.
n-C12H28 + C2H5- = n-C6H13CHC5Hu + C2H6 (41)
AH° = -10.3 kcal, AG° = -9.7 kcal
This indicates a favorable, although not extremely rapid, type of re-
action responsible for the start of chain branching.
The highly unfavorable character of the monomolecular dissocia-
tion to form an olefin and a hydrogen atom is shown below:
n-C12H25 • = n-C12HM + H • AH° = + 41.0 kcal, AG° = + 33.7 kcal
(42)
Chain termination by pairs of free radicals, either by combination or
disproportionation, is extremely favorable thermodynamically. This
thermodynamic advantage is offset by the low concentrations of free
radicals usually present in a polymerizing mixture. Two reaction
schemes for termination are

2n-C12H25 • = n-C24H50 AH° = -71.8 kcal, AG° = -59.1 kcal (43)


2n-C12H25 • = n-C12H24 + n-C12H26
AH" = -52.5 kcal, AG° = -51.1 kcal (44)
B. Free Energies of Polytetrafluoroethylene Synthesis
1. Heats of Formation of Fluorocarbon Radicals. With both the
heats of formation and the C—F bond dissociation energies available
for various fluorocarbons, the heats of formation of the correspond-
ing radicals can be calculated. The enthalpy of formation of the tri-
fluoromethyl radical, for example, is available from two independent
sequences of reactions [24]. The first is
382 HIDEO SAWADA

AH°
CF4(g) - CF,'(g) + F-(g) +123 kcal
C(graphite) + 2F2(g) — CF4 (g) -217.8 kcal
F-(g) — £F 2 (g) - 18.9 kcal

C(graphite) + f F2(g) - CF3 • (g) -113.7 kcal


and the second is
AH3
CHF3(g) - CF,-(g) + H-(g) +102 kcal
C (graphite) + fF 2 (g) + }H2(g)
— CHF3(g) -162.6 kcal
H-(g) — -|H 2 (g) - 52.1 kcal

C(graphite) + { F,(g) - CF3 • (g) -112.7 kcal


The two sources agree within 1 kcal. Enthalpies of formation for a
series of fluorocarbon radicals are given in the original paper [24].
2. Molecular Entropies and Entropies of Formation. Parks and
Huffman [25] found that linear equations for S° of fluorocarbons as
a function of the number of carbon atoms will fit reliable experi-
mental data within a few tenths of an entropy unit. The following
equation of the Parks and Huffman type was derived by Bryant [24].

C n F 2n+2 (g): S°= 46.41 + 16.066n (45)


where n is the number of carbon atoms in the molecule.
Entropies of fluorocarbon free radicals were computed from those
of the parent fluorocarbons by considering the following factors:
(1) increase in entropy on passing from a singlet to a doublet ground
state; (2) change in symmetry number; (3) approximate reduction in
entropy due to loss of a fluorine atom.
Comparison of entropies of the lower aliphatic fluorocarbons and
their radicals lead to the following relationship for monoradicals
C n F 2n+1 • (g) derived from perfluoroparaffins:
THERMODYNAMICS OF POLYMERIZATION. I 383

S° = S° (fluorocarbon) + 1.375 (due to greater multiplicity)


+ 1.375 (due to loss of symmetry) - 4.25 (due to loss of
fluorine atom) = S° (fluorocarbon) - 1.50 ± 1.50 (46)

Numerically, for normal perfluoroalkyl radicals, Eq. (46) becomes

C n F 2 n + 1 • (g): S° = 44.91 + 16.006n (47)

for radicals above CF 3 -. The entropies of perfluoroalkyl radicals


were calculated with the aid of Eq. (47). Then the molecular entro-
pies of fluorocarbons and their radicals maybe combined withS°c and
Sp to yield the corresponding entropies of formation.
3. Free Energy Changes. Where a large number n of monomer
units is involved, the overall synthesis may be summarized by the
following equation which ignores the individual steps of the mecha-
nism:

C2F4(g)- l/n[C 2 F 4 ] n (g)


AH° = -37.10 kcal, AG° = - 25.32 kcal (48)

The enthalpy of polymerization for the ideal gaseous state is given


in Eq. (48).
The corresponding value based on the solid polymer is

C2F4 (g) - l/n [C 2 F 4 ] n (s) AH0 = -41.12 kcal (49)

The difference in AH° between Eqs. (48) and (49), 4.02 kcal, is of the
correct magnitude for the heat of sublimation.
4. Initiation and Chain Propagation. The first step in a chain of
. free radical reactions is that of initiation. This is illustrated by the
following propagation sequence:

F • (g) + C2F4 (g) — CF3CF2 • (g)


AH0 = -79.00 kcal, AG" =-69.24 kcal (50)
or
CF 3 • (g) + C2F4 (g) - CF3CF2CF2 • (g)
0
AH =-41.90 kcal, AG° =-29.67 kcal (51)
384 HIDEO SAWADA

CF3CF2 • (g) + C2F4 (g) - CF3CF2CF2CF2 • (g)


AH° =-37.10 kcal, AG° = -25.32 kcal (52)
n-C10F21 • (g) + C2F4 (g) - n-C12F25 • (g)
AH0 =-37.10 kcal, AG° = -25.32 kcal (53)
The constant values of enthalpy and free energy of propagation in
Eqs. (52) and (53) coincide with those for the overall polymerization
process if the latter is associated with a product of high molecular
weight, since the contributions of initiation and termination become
vanishingly small. The free energy of the propagation step is one of
the most favorable reactions, although the large enthalpy value sug-
gests that the reaction may become less favorable fairly rapidly
with rising temperature.
5. Termination, Chain Transfer Reactions, and Chain Branching.
Termination by mutual combination of two free radicals is highly
exothermic:
2n-C12F25 • (g) - ' n-C24F50 (g)
AH3 =-81.50 kcal, AG° = -68.50 kcal (54)
The corresponding termination by disproportionation is also strongly
exothermic:
2n-C12F25*(g) — n-C12F28(g)+ n-C12F24 (g)
AH0 = -56.80 kcal, AG° = -55.60 kcal (55)
Instead of a perfluoroparaffin of double the number of carbon atoms,
the products here are a molecule each of perfluoroparaffin and per-
fluoroolefin.
There is no indication of a tendency for a fluorocarbon radical to
go over to a perfluoroolefin with the loss of a fluorine atom at ordi-
nary temperature:
n-C12F25 • (g) - n-C12F24 (g) + F • (g)
AH3 = + 66.60 kcal, AG° = + 56.92 kcal (56)
' The following examples suggest qualitative similarity to the com-
parable reactions involving hydrocarbon radicals:
n-C12F28 (g) + CF3 • (g) - n-C12F25 • (g) + CF4 (g)
AH3 = -0.60 kcal, AG° = -0.19 kcal (57)
THERMODYNAMICS OF POLYMERIZATION. I 385

n-C3F7- (g) + C3F8(g) - i-C 3 F 7 - (g) + C3F8(g)


AH° = -13.40 kcal, AG= = -13.00 kcal (58)
Thermodynamically, with respect to two possible reactions, chain
transfer with the monomer is even more clear-cut than in the polym-
erization of ethylene:
n-C12H25 • (g) + C2F4 (g) - n-C12F24 (g) + C2F5 • (g)
AH° =-12.40 kcal, AG° = -12.32 kcal (59)
n-C12F25 • (g) + C2F4 (g) - n-C12F28 (g) + CF 2 =CF • (g)
AH3 = + 1.60 kcal, AG° = + 1.58 kcal (60)
However, kinetics may play a more important role here than thermo-
dynamics. It is doubtful whether chain transfer with the monomer is
important in the polymerization of tetrafluoroethylene, in view of the
customary high molecular weight of the product.

References
[1] N. N. Semenov, Some Problems in Chemical Kinetics and Reactivity
(M. Boudart, transl.), Vol. 1, Princeton Univ. Press, Princeton, N.J.,
1958, Chap. 1.
[2] P. J. Flory, Principles of Polymer Chemistry, Cornell Univ. Press,
Ithaca, N.Y., 1953, Chap. 4.
[3] H. Sawada, Chem.High Polymers (Tokyo), 20, 561 (1963).
[4] H. Sawada, Chem. High Polymers (Tokyo), 21, 251 (1964).
[5] C. Walling, Free Radicals in Solution,Wiley, New York, 1957, Chap. 2.
[6] C. T. Mortimer, Reaction Heats and Bond Strengths, Pergamon,
London, 1962, Chap. 1.
[7] M. G. Evans, in Fibres from Synthetic Polymers (R. Hill, ed.),
Elsevier, Amsterdam, 1953, Chap. 3.
[8] J. A. Kerr, Chem. Rev., 66, 465 (1966).
[9] P. M. Morse, Phys.Rev., 34, 57 (1929).
[10] M. G. Evans and M. Polanyi, Trans. Faraday Soc, 34, 11 (1938).
[11] M. G. Evans, J. Gergely, and E. C. Seaman, J. Polymer Sci., 3, 866
(1948).
[12] M. S. Matheson, E. E. Auer, E. B. Bevilacqua, and E. J. Hart, J.Am.
Chem. Soc.,71, 497, 2610 (1949).
[13] R. D. Snow and F. E. Frey,hid. Eng. Chem. (Ind. Sect.),30, 176
(1938).
[14] S. Bywater, Trans. Faraday Soc, 51, 1267 (1955).
386 HIDEO SAWADA

[15] G. G. Lowry, J. Polymer Sci., 31, 187 (1958).


[16] J. C. Bevlngton, Trans. Faraday Soc, 48, 1045 (1952).
[17] E. J. Prosen, W. H. Johnson, and F. D. Rossini, J. Res.Natl. Bur. Std.,
37, 51 (1946).
[18] K. S. Pitzer, Client. Rev., 27, 39 (1940).
[19] J. E. Kilpatrick, E. J. Prosen, K. S. Pitzer, and F. D. Rossini, J . Res.
Natl. Bur. Std., 36, 559 (1946).
[20] W. M. D. Bryant, J. Polymer Sci., 6, 359 (1951).
[21] K. S. Pitzer, J. Chem.Phys., 5, 473 (1937).
[22] K. S. Pitzer and D. W. Scott, J.Am. Chem. Soc, 65, 803 (1943).
[23] R. S. Jessup, J. Chem. Phys., 16, 661 (1948).
[24] W. M. D. Bryant, J. Polymer Sci., 56, 277 (1962).
[25] G. S. Parks and H. M. Huffman, Free Energies of Some Organic Com-
pounds , Chemical Catalog Co., New York, 1932.
Chapter 4. Thermodynamics of Polycondensation

I. GENERAL ASPECTS 387


II. DEGREE OF POLYMERIZATION 388
III. EQUILIBRIUM CONSTANT 389
IV. RING FORMATION IN POLYCONDENSATION 394
REFERENCES 395

I. GENERAL ASPECTS
The polycondensation reaction is a random one and the rate con-
stant is independent of molecular size. If the simplest case is con-
sidered, i.e., the condensation of an w-hydroxy acid or a dibasic
acid and dihydric alcohol in equimolar concentration. In this case,
the carboxyl and hydroxy groups are in equilibrium with the ester
group and water; i.e.,

[/w\OH] — [/w\COO/w\] + [H2O]

The equilibrium constant for the reaction is

where nH is the concentration of water and p is the extent of reaction.


The equilibrium constant for polyester is about one-hundredth of that
387
388 HIDEO SAWADA

for polyamide. Therefore, it is impossible to obtain high-molecular-


weight polyester by polycondensation without reduced pressure.

n. DEGREE OF POLYMERIZATION
The number-average degree of polymerization P n is a function [1]
of the extent of reaction p:

and, hence,
_ 2K/nH
Pn =
VI + 4K/nH
and if K/nH » 1, then

therefore,
JL | 3 In K
2 I 3(1/ T )

The free energy change is given by AG = AH - T AS = -RT In K. If


log K is plotted against reciprocals of absolute temperature, a linear
relationship holds roughly and AH is given. Hence, at constant water
concentration, we have
91ng
" = - a AH (5)
(/)
where a is a constant, and AH is the enthalpy change of the reaction.
The formation of condensation polymers in the molten state is
exothermic [2], and hence AH < 0. Thus, the molecular weight of the
polymer made by a melt process increases as the temperature is
lowered [2,3].
However, under the same vapor pressure of water, the degree of
polymerization of polycapramide is increased as the temperature
, rises [4]. These results show that the change in activity of water
with temperature has much greater effect on the degree of polymeri-
THERMODYNAMICS OF POLYMERIZATION. I 389

zation than the change in equilibrium constant with temperature. The


heat of polycondensation of polycapramide, AH = - 6 . 8 kcal, when the
liquid state is chosen for the standard state of water, seem quite
resonable in view of the heats of reactions between acetic acid and
ammonia. If the vapor p r e s s u r e of water is employed in place of n w ,
the value of AH becomes 11.5 kcal (endothermic). Thus, because
of the unexpectedly high heat of vaporization of water, the calculated
value for the heat of condensation becomes markedly endothermic in
spite of the exothermic nature of the amide formation.

III. EQUILIBRIUM CONSTANT


Wiloth [5] and Hermans [6] have reported that the equilibrium
constant K becomes smaller with the increase of water content and
the reason is ascribed to using the concentration instead of the
activity in kinetics.
The reaction in the e-caprolactam polymerization is very com-
plicated, nevertheless the reaction equilibria concerning to chain
or ring monomer can be shown as follows:

H2O

H2O

Yumoto and Ogata [7] have reported that the values of K2 a r e


smaller than those of K, and the values of K become smaller with
the increase of water content and approach to the mean value of K 2 .
This was explained [7] as follows. The equilibrium constants be-
come greater a s the chain length becomes longer; therefore, the
equilibrium constant K becomes smaller as the mean degree of
polymerization becomes smaller because of the increase of the
chain monomer and oligomer contents. The difference among the
equilibrium constants means that the reactivity of the functional
end-group i s not equal, especially for the chain oligomers.
Meggy [8] calculated equations for the activity coefficients of
water and polymer in the system water-e-caprolactam. The degree
of polymerization is determined by the equilibrium:
x - m e r + y - m e r = (x + y)-mer + H2O
= (H2O)([x+y]-mer) (6)
'(x-mer) (y-mer)
390 HIDEO SAWADA

The polymerization does not go to completion; 5-10% of the lac-


tam remains in the equilibrium mixture. The lactam may be regarded
as being in equilibrium with 6-aminohexanoic acid and water:

K (lactam)(H2O)
K
2
(6-aminohexanoic acid)
Lactam is also in equilibrium with polymer:
lactam + x-mer = (x + l)-mer
_ ([x+ l]-mer)
* 3 " (x-mer)(lactam) l
'
In order to develop these equations, it is necessary to make two
assumptions: (a) in Eq. (6) the value of Kx at any temperature is the
same for all values of x and y, including x or y = 1 (this is the prin-
ciple of "equal reactivity of all groups" postulated by Flory); (b) in
the presence of a diluent, if a? is the activity of the whole polymer,
and m x is the molecular fraction of the x-mer in the polymer, then
the activity of x-mer is a2tnx.
On the basis of these two assumptions, it is possible to develop
Eqs. (6) to (8). Let the activity of water in the system be ax, that of
the polymer be a?, that of lactam be a3, and the degree of polymeri-
zation be n. The activity of x-mer is a j l - l/n) /n x , the activity
of the y-mer is a^l - l/n^'Vny, and of the (x + y)mer is
^ - l/n) x+y "Vn 2 . Substituting in Eq. (6), we have

The molecular fraction of monomer(x = 1) in the polymer is l/n 2 .


Hence,

K2 = a1a3n2/a2 (10)
The ratio (x + l)-mer/x-mer = ( n - l)/n.

K3 = ( n - U / n a , (11)
Equations (9), (10), and (11) are not independent; by combining any
two, the third is obtained:
THERMODYNAMICS OF POLYMERIZATION. 1 3SM

The composition may be defined conveniently as follows:


(mols. of water)
1
~ (mols. of water + mols. of unit)

= (mols. of unit as polymer)


N2 (mols.
jls. of water + mols. of ur
unit)
(mols. of unit as lactam)
N,3 = (mols. of water + mols. of unit)

Nx + N2 + N3 = 1 (12)

From the Gibbs-Duhem equation, we have

+N2/91nM + N3/91naA =0 (13)

From Eqs. (6) to (13),

where v1 ='a 1 /N 1 , i>2 = ag/Nj, and K^ and K3 are apparent equilib-


rium constants defined by Kx = (Nx /^I2) P n (Pn ~ 1) and K 3 = ( p n - 1)/
PnN3-
The values of In u1 and In vz can be evaluated by graphical inte-
gration. The value of the true equilibrium constants Kx is given by
K
i = ("i/^KT (16)

Although Kx becomes smaller with the increase of water content, Ky


has a constant value at a constant temperature—Kx is 440 at 221.5°C
and 379 at 253.5°C. From these values, log Kx is given by
logKi = 1.569+ 531/T (17)
AH = -2.42 kcal/mole, AS = 7.16 cal/mole deg.
392 HIDEO SAWADA

Starting from the principle of equal reactivity the entire, revers-


ible polycondensation of polyethylene terephthalate (PET) may be
symbolized by

E+ E ^ Z+ G
where
E = — COOCH2CH2OH
Z = — COOCH2CH2OOC—
G = HOCH2CH2OH

The forward reaction proceeds by ester interchange between the


alcoholic hydroxyl group of a 2-hydroxyethyl ester end-group, and
the ester linkage of another 2-hydroxyethyl ester end-group and
linking the ends of two chain molecules under elimination of glycol.
The reverse reaction is a glycolytic degradation and proceeds by
ester interchange between a hydroxyl group of glycol and an ester
link of an ethylene diester group in a PET chain.
For comparison with other ester interchange reactions, Challa [9]
has expressed the rate constants (k and k') and the equilibrium con-
stant (K) in terms of the elementary reaction between a single hy-
droxyl group and a single ester link. Therefore, the following rate
expression is used:

-^=Ke2-4k'gz (18)

where the corresponding, small letters denote the concentrations of


the reactants and products. The rate of the forward reaction is rep-
resented by ke2, since one molecule of glycol can be produced by re-
action between any hydroxyl group and any ester link belonging to
different 2-hydroxyethyl ester end-groups. The rate of the reverse
reaction is represented by 4k'gz, since the reaction between any bi-
functional ethylene diester group and any bifunctional glycol mole-
. cule can proceed in four equivalent ways.
For dg/dt = 0, Eq. (18) yields the equilibrium conditions:

Replacing the concentrations by the number of equivalents or moles


present in the equilibrium mixture results in
THERMODYNAMICS OF POLYMERIZATION. I 393

(2Q)

A more practical form of Eq. (20) can be obtained by introducing the


equilibrium extent of end-group reaction p e , i.e., the fraction of
original end-groups that has reacted at equilibrium:

p e = (2U0 - E e )/2U 0 (21)


or
E e = (1 - pe)2U0 (22)
where Uo is the number of moles of repeating units per gram of the
sample.
Z equals half the number of end-groups E that have reacted. Thus,
Z e = (2U0 - E e )/2 = Pe U 0 (23)
Gg is given by
Ge = Go + (Eo - E e )/2 (24)
where Go denotes the number of moles of free glycol added before
sealing the reaction tube.
P^ = 2U0/E0 = 1/(1 - p0) (25)
where p0 is the well-known Flory extent of reaction.
In the case of previous addition of glycol, q is defined as
q = Go/Uo (26)
From Eqs. (22), (25), and (26), Ge is given by

Ge = (q + p e - Po) Uo (27)
Finally, substitution of Eqs. (22), (23), and (27) into Eq. (20) results
in
K = Pe(q +Pe-Po) (28)
(1-Pe) 2
From the known starting conditions (p0 and q) and the analytical
results, the equilibrium extent of reaction p e can be computed. Then,
394 HIDEO SAWADA

substitution of p0, q, and pe into Eq. (28) yields the polycondensation


equilibrium constant K; K increases with the extent of reaction and,
thus, also with the average degree of polymerization. Challa [9]
proved that the principle of equal reactivity does not fully hold in
PET. His result is shown in Fig. 4.1. The influence of temperature
on K appears to be of minor importance. Application of the thermo-
dynamic relationships
AH = RT2 d In k/dT (29)
AS = d(RT In k)/dT (30)
yields AH = - 2 kcal/mole and AS = - 5 cal/deg mole.

1-2 .

08-

0-4-

0-2 04 0-8

- R,
Fig. 4.1. Polycondensation equilibrium constant K at 262.0°C as a function
[9] of l - p e .

IV. RING FORMATION IN POLYCONDENSATION


Stbckmayer and Jacobson [10] have attempted to formulate a
general theory of the polycondensation reaction. Their treatment of
the problem is formally identical with that of the thermodynamics .
of a perfect Bose-Einstein gas, and the critical phenomena may be
compared with the condensation of such a gas.
The only arbitrary assumption made in this treatment is that the
reactivity of the groups is the same and independent of the molecular
THERMODYNAMICS OF POLYMERIZATION. I 395

size. Furthermore, it is necessary that the elementary steps of the


reaction are completely reversible. This theory predicts that the
formation of rings depends on the dilution.

ACKNOWLEDGMENTS
I wish to express my appreciation to Professor K. F. O'Driscoll
of State University of New York at Buffalo for his interest and will-
ingness to read and criticize the manuscript, and to state my indebt-
edness of Mr. M. Dekker of Marcel Dekker, Inc. for his interest and
valuable counsel. I am also indebted to the management of Daicel
Ltd. for permission to write this review. Thanks are also due
Mrs. M. Sakakibara who typed large sections of the manuscript.

References
[1] P. J. Flory, Principles of Polymer Chemistry, Cornell Univ. Press,
Ithaca, N.Y., 1953, p. 81.
[2] J. Zimmerman, J. Polymer Sci., B2, 955 (1964).
[3] H. Sawada, and A. Yasue, J. Chem. Soc. Japan, Ind. Chem. Sect., 67,
1442 (1964).
[4] O. Fukumoto,J. Polymer Sci., 22, 263 (1956).
[5] F. VViloth, Makromol. Chem., 15, 98 (1955).
[6] P. H. Hermans, J. Appl. Chem., 5, 493 (1955).
[7] H. Yumoto, and N. Ogata, Makromol. Chem., 25, 91 (1957).
[8] A. B. Meggy, J . Chem. Soc., 1956, 4876.
[9] G. ChMa, Makromol. Chem., 38, 105 (1960).
[10] W. H. Stockmayer and H. Jacobson,J. Chem.Phys., 18, 1600 (1950).
J. MACROMOL. XI.-REVS. MACROMOL. CHEM., C5(1), 151-174 (1970)

Thermodynamics of Polymerization.
II. Thermodynamics of Ringopening
Polymerization

HIDE0 SAWADA
Central Research Laboratory
Daicel Ltd.
Tsurugaoka, Oi, Irurnagun
Saitarna, Japan

I. GENERALASPECTS ......................... 151


11. HOMOCYCLIC COMPOUNDS. .. . .. .. . .. . . .. . . . . . 152
A. AngleStrain . . .. . . . . . . . . . . . . .. . . . . . . . . . . 152
B. Conformational and Transannular Strain . . . . .. . . . . 162
C. Steric Effect of Side Group . .. . . . ... . . . . ... . . . 163
III. HETEROCYCLIC COMPOUNDS . . . . . . . . . . . . . . . . . . 165
A. Cyclic E t h e r s . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
B. Lactams .... . . . .. . . . . ... . .. .. . . . . . .. .. . 168
C. L a c t o n e s . . .. . . . . . . . .. . . . . . . . . . . . . . . . . . . 170
D. Miscellaneoris Heterocyclic Compounds . . . . . . ... . 170
IV. SUMMARY.. . .. . .. . .. . . . .... . . .. . ......... 172
ACKNOWLEDGMENTS. . . . . . . . . . . . . . . . . . . . . . . . 172
REFERENCES ............................. 172

I. GENERAL ASPECTS

The heat of polymerization for ring-opening polymerization i s


affected by the following: (1) angle strain, which is very important
for the three- and four-membered rings (i.e., those rings that have
the greatest angle strain), (2) conformational strain (crowding of
151
Copyright 0 1970 by Marcel Dekker, Inc.
152 HIDE0 SAWADA

eclipsed adjacent hydrogen atoms) and transannular strain (strain


and crowding of hydrogen atoms across the ring); and (3) steric
effect of side group.
On the other hand, the entropy changes of polymerization do not
depend on angle strain, but a r e susceptible to configurational in-
fluence.
Polymerization of ring compounds to linear polymers may show
ceiling temperatures as in the case of tetrahydrofuran and 1,3-di-
oxolane or floor temperatures as in the case of sulfur and selenium.
The equilibrium polymerizations of ring compounds will be discussed
in a later review.

II. HOMOCYCLIC COMPOUNDS


The cyclic compounds for which most data a r e available a r e homo-
cyclic compounds.
A. Angle Strain
Angle strain, defined as the distortion of bond angles from their
most favored orientation, serves to increase the internal energy of
a molecule. Those molecules in which the skeletal bond angles are
closest to the preferred values will, therefore, have the lowest in-
ternal energy.
Ring-opening polymerization takes place through opening of the
ring, and the linear polymers so formed a r e virtually free from
strain. The heat of polymerization, therefore, provides a direct
measure of the strain energy in the monomer ring [l]. Heats of
polymerization for ring monomers a r e summarized in Table 1.
Von Bayer [2] pointed out that if the normal angles between the
four valencies of a tetrahedral carbon atom a r e assumed to be 109”28‘,
i.e., the valencies are equally spaced about the atom, then, in three-
and four- membered rings, severe distortion of the normal valency
angles must be involved. Thus, in a cyclopropane ring each bond
must be distorted from the tetrahedral angle by 24O44’.
If all the rings a r e assumed to be planar, the distortions for each
ring size a r e shown in Table 2. Small ring compounds are generally
reactive due to high strain in their ring structure. The presence of
“strain” and hence of decreased stability has been shown by the en-
hanced values of the molecular heat of combustion per methylene
group of cyclopropane and cyclobutane compared t o the values for
the larger rings. The next to the last column in the table gives the
strain per methylene group, and this quantity is a measure of the
distortion present. The last column gives the “strain energy” in the
THE R MODY NAMlCS OF PO LYME R I ZATl ON. I I 153

TABLE 1
Heats of Polymerization for Ring Monomers

No. of
atoms Standard -AH, Temp.,
Monomer in ring states* kcal/mole "C Ref.

a. Cycloalkanes
a
Cyclopropan e 3 lc 27.0 25
a
Methylcyclopropane 3 lc 25.1 25
1,l-Dimethylcyclopropane 3 lc 23.3 25 d

Cyclobutane 4 lc 25.1 25 b
Methylcyclobutane 4 lc 23.9 25 a
1,l-Dimethylcyclobutane 4 Ic 22.3 25 a

Cyclopentane 5 lc 5.2 25 b
Methylcylopentane 5 lc 4.1 25 a

1,l-Dimethylcyclopentane 5 lc 3.2 25 a

Cyclohexane 6 lc -0.7 25 b
Methylcyclohexane 6 lc -2.2 25 a
1.1- Dimethylcyclohexane 6 lc -1.8 25 a

Cycloheptane 7 lc 5.1 25 b
Cyclooctane 8 lc 8.3 25 b
Cyclononane 9 IC 11.2 25 C

Cyclodecane 10 lc 11.5 25
Cycloundec an e 11 lc 10.8 25 C

Cyclododecane 12 lc 3.4 25 C

Cyclotridecane 13 lc 5.3 25 C
C
Cyclotetradecane 14 lc 1.7 25
C
Cyclopentadecane 15 lc 2.9 25
Cyclohexadecane 16 lc 1.9 25 C

Cycloheptadecane 17 lc 2.0 25 C

b. Cyclic Ethers
Ethylene oxide 24.9 25 d
3 gg
lc 22.6 25 e

h o p y l e n e oxide 18.0 25 epf

styrene oxide 3 lc 24.3 26.9 d


3-Nitrostyrene oxide 3 Ic 24.1 26.9 d

Oxetane 4 ss 19.3 -9 g
154 HIDE0 SAWADA

TABLE 1 (continued)
No. of
atoms Standard -AH, Temp.
Monomer in ring states* kcal/mole OC Ref.

3,3-Di (chloromethyl) -
d
oxetane 4 lc 20.2 26.9
3.3-Di (phenoxymethyl) -
oxetane 4 ss 19.8 26.9
3,3- Dimethyloxetane 4 ss 16.1 -9

Dioxolane 5 lc 6.2 20 h
5s 5.2 9 i

Tetrahydrofuran 5 gg 5.0 20
gg 2.9 25
Ic 9.1 25
1s 4.3 40
1s 5.3 40
Is 4.0 25
lc 3.0 50
lc 5.3 25

Tetrahydropyran 6 gg 0.4 20 j, k , ~

Trioxane 6 gc 3.83 25 9
cc -0.57 25 9

m- Dioxane 0.0 20 h. P

- 65
r
1.3-Dioxepan 7 ss 3.6 -+ 5
4.7 20 h
gg

7 lc 1.8 26.9
ss 2.1 26.9
6s 1.9 26.9

t
Tetraoxane 8 cc 0.7

h
1,3-Dioxocane 12.8 20

c. Cyclic Amides
a-Pyrrolidone 5 lc 1.1 75 U

lc 1.3 25 V

1-Methyl-a-pyrrolidone 5 lc 0.8 25 W
THERMODYNAMICS OF POLYMERIZATION. II 155

TABLE 1 (continued)
No. of
atoms Standard -AH, Temp.
Monomer in ring states* kcal/mole "C Ref.
~~ ~~ ~

a -Piperidone 6 lc 2.2 75 U

lc 1.1 25 V

1-Methyl-a-piperidone 6 lc -0.5 25 W

E- Caprolactam 7 lc 3.8 75 I

lc 3.3 25 V

1s 3.6 250 Y
1s 4.5 240 L

Is 4.0 250 aa
1s 3.3 230 bb
1-Methyl-c- caprolactam 7 lc 2.3 25 W

5- Methyl-c- caprolactam 7 lc 3.8 75 U

7- Methyl-€- caprolactam 7 Ic 3.8 75 U

U
E- Enantholactam 8 lc 5.3 75
lc 5.7 25 V

Ic 5.2 230 bb
1-Methyl-c-enantholactam lc 3.9 25 W

q- Capryllactam 9 lc 7.8 230 bb

d. Inorganic Monomers
Selenium 8 1s -2.3 400 c,c
Sulfur 8 1s -3.2 200 dd.ee

e. Cyclic E s t e r s
P-Propiolactone 4 lc 19.2 - ff

Pivalolactone 4 lc 20.1 25 66

*Monomer and polymer states: 1. liquid; s, solution; c, condensed; g, gas.


aF.S.Dainton,T.RE.Devlin, and P.A.Smal1, Trans, Faraday'Soc., 51, 1710
(1955) .
bS. Kaarsemaker and J.Coops, R e c . Trau. Chim., 71, 261 (1952).
H.van Kamp, J.Coops, W.A.Lambregts, B. J.Visser, and H.Dekker, Rec.
Trau. Chim., 79, 1226 (1960).
dF.S.Dainton, K. J.Ivin, and D.A.G. Walmsley, Trans. Faraday SOC., 66,
1784 (1960).
H.C.Raine, R.B.Richards, and H.Ryder. Trans. Faraday Soc., 41. 56
(1945) .
'P.Cray and A.Williams, Trans. Furuduy SOC., 55, 760 (1959).
'J.B.Rose, J . Chem. SOC., p.546 (1946).
156 HIDE0 SAWADA

TABLE 1 (continued)
~

hS.M.Skuratov, A.A.Strepikheev. S.M.Shtekher, and A.V.Volokhina, Dokl.


Akad. Nauk SSSR. 117, 263 (1957).
'P.H.Plesch and P.H.Westermann, J . Polymer Sci C , 16, 3837 (1968).
IS. M.Skuratov, A.A.Strepikheev. and M.P.Kozina, Dokl. Akad. Nauk SSSR,
117, 452 (1957).
kR.C.Cass, S . E.Fletcher, C.T.Mortimer, H.D.Springal1, and T.RWhite,
J . Chem. SOC.,p.1406 (1958).
'D.Sims, J . Chem. SOC., p.864 (1964).
mC.E.H.Bawn, R.M.Bel1, and A.Ledwidth,Polymer, 6, 95 (1965).
"K.J.Ivin and J.Leonard, Polymer, 6, 621 (1965).
OG.A.Clegg, D.RGee, T.P.Melia. and A.Tyson,Polymer, 9, 501 (1968).
'A.Snelson and H.A.Skinner, Trans. Faraday S O C . , 57, 2125 (1961).
T.P.Melia, D.Bailey, and A.Tyson, J . Appl. Chem., 17, 15 (1967) .
P.H.Plesch and P.H.Westermann, Polymer, 10, 105 (1969).
F.S.Dainton, J.A.Davies, P.P. Manning, and S.A. Zahir, Trans. Faraday
SOC., 53, 813 (1957).
K.Nakatsuka, H.Suga, and S.Seki, J . Polymer Sci. B , 7 , 361 (1969).
" A.A.Strepikheev, S.M. Skuratov, 0.N.Kachinskaya. R S . Muramova,
E.P.Brildina, and S.M.Shtekher, Dokl. Akad. Nauk SSSR, 102, 105 (1955).
"V.P.Kolesov, I.E.Paukov, and S.M.Skuratov. Zh. Fiz. Khim., 36, 770
(1962);Russ. J . Phys. Chem., 36, 401 (1962).
WM.P.Kozina and S.M.Skuratov, Dokl. Akad. Nauk S S S R , 127, 561 (1959).
* S.M.Skuratov, A.A.Strepikheev. and E.N.Kanarskaya, Kolloidn. Zh., 14,
185 (1952).
YA.B.Meggy, J . Chem. SOC., p.796 (1953).
'P.F.Van Velden, G.M.Van d e r Want. D.Heikens, C.A.Kruissink,
P.H.Hermanb, and A.J.Staverman, Rec. Traw. Chim., 74, 1376 (1955).
aaA.V.Tobolsky and A.Eisenberg, J . Am. Chem. SOC.,81, 2302 (1969).
"A.K.Bonetskaya and S.M.Skuratov. Vysokomolekul. Soedin., All, 532
(1969) .
"A.Eisenberg and A.V.Tobolsky, J . Polymer Sci., 48, 19 (1960).
ddF.Fairbrother, G.Gee, and G.T.Merral1, J . Polymer S c i . , 16, 459 (1955).
eeA.V.Tobolsky and A.Eisenberg, J . Am. Chem. SOC., 81, 780 (1959).
"B.Boyesso, Y.Nakase, and S.Sunner, Acta Chem. Scand., 20. 803 (1966).
"H.K.Hall,Jr., Macromolecules, 1. 488 (1969).

molecule, which is the difference between its observed heat of for-


mation and its value estimated from the group additivity relations
using the group values which have been, of course, derived from
the unstrained standards [3]. The angle strain is seen to be quite
severe in cyclopropane, less severe in cyclobutane, and still less in
cyclopentane. Cyclohexane lies at the energy minimum, and for the
larger rings the strain increases to a maximum at cyclononane and
then falls off as shown in Table 2.
TABLE 2
Distortions in Cycloalkanes

Molecular heat
Distortion f r o m Heat of of combustionb Strain
No. of atoms Angle between normal valency polymerization p e r CH2 group, energy,,
in ring valency bonds angle a AHp,, kcal/mole kcal/mole kcal/mole
~~~

(2) (ethylene) 0" 54044" 24.2d 168.7 (22.6)


3 60" 24"44" 27.0 166.6 27.6 (27.6)
4 90" 9044" 25.1 164.0 26.2 (26.4)
5 108" O"44" 5.2 158.7 6.3 ( 6.5)
6 120" -5"16" -0.7 157.4 0.2 ( 0.0)
7 128"34" -9"33I' 5.1 158.3 6.4 ( 6.3)
8 135" -12"46" 8.3 158.6 9.9 ( 9.6)
9 - - 11.2 158.8 12.8 (12.6)
10 - - 11.5 158.6 (12.0)
11 - - 10.8 158.4 (11.0)
12 - - 3.4 157.7 ( 3.6)
13 - - 5.3 157.8 ( 5.2)
14 - - 1.7 157.4 ( 0.0)
I -
15 2.9 157.5 ( 1.5)
16 - - 1.9 157.5 ( 1.6)
17 - - 2.0 157.2 (-3.4)

aDistortion = (normal valency angle-actual angle between bonds). The distortion caused is assumed to be equally
shared between the two bonds.
bThe heat of combustion of the gaseous cycloalkane divided by the number of methylene groups. Data from
E.L. Eliel, N.L. Allinger, S.J. Angyal, and G.A. Morrison, Conformational Analysis, Wiley-Interscience, New York,
1965,p. 193.
Strain energies within parentheses a r e calculated according to the following equation:
(the molecular heat of combustion of cycloalkane -
strain energy = n 157.4)
n
where n i s the number of methylene groups.
Monomer, gas polymer, condensed.
-
158 HIDE0 SAWADA

The calculated heat, entropy, and free energy of polymerization


of the liquid cycloalkanes to linear polymer at 25°C are given in
Table 3. Using heats of formation of monomer and polymer the
heats and entropies of polymerization have been calculated by Dain-
ton et al. [ 4 ] . In Table 3 these A H and AS values show that A H makes

TABLE 3
Heats, Entropies, and Free Energies of Polymer-
ization of Cycloalkanes a t 25°C [4]

AH^,, - AS.9 - A G ~c ,
X kcal/mole cal/deg mole kcal/mole

L(CHz).d
3 27.0 16.5 22.1
4 25.1 13.2 21.2
5 5.2 10.2 2.2
6 -0.7 2.5 -1.4
7 5.1 3.8= 3.9
8 8.3 0.8a 8.2
CH~~.H(CH~)~-,~H~
25.1 20.2 19.1

-
23.9 17.2 18.8
4.1 15.3 -0.5
-2.2 7.6 -4.5
(CHS),C( CH,)x-zCH,
23.3 22.3 16.6
22.3 18.0 16.0
3.2 15.7 -1.5
-1.8 8.5 -4.3

'Data f r o m H.L.Finke, D.W.Scott, M. E.Gross, and


G.Waddington, J. Am. Chem. SOC., 78, 5469 (1956).

the main contribution to A G for three- and four-membered rings,


but the heat and entropy contributions are equally important for five-,
six-,and seven-membered rings. These results showed that cyclo-
hexane was the most resistant to polymerization, since AG was posi-
tive. For cyclopropane, cyclobutane, cyclopentane, cycloheptane,
and cyclooctane, A G for polymerization was calculated to be negative.
Thermodynamic feasibility, however, does not always guarantee the
practical realization, and no high polymers of cyclopropane and cy-
clobutane are known [5].
THERMODYNAMICS OF POLYMERIZATION. II 159

From a qualitative point of view, the angle strain in three- and


four-membered carbon ring systems is large: it is small in five- and
six-membered rings, and then increases with further increase in
ring size, and this is mirrored in the rise and fall of AH>c.
On the other hand the entropy changes do not show much marked
dependence on angle strain, but are susceptible to configurational
influence. The entropy change of polymerization can be given as a
function of the probability of ring closure. Thus the entropy change
of polymerization is

Asp = - b In P - a (1)

where P is the probability of ring closure and a and b a r e constants,


the values of which depend on the nature of the monomer. The prob-
ability of ring closure for a chain with n repeating units can be given
as a function of the probability that the chain ends will come together.
This probability is usually given in t e r m s of a value (?), the root-
mean-square of end-to-end distance. Obviously, for ring closure t o
occur, r must be less than a few AngstriSm units. The entropy change
would have a large negative value for the ring opening of a three-
membered ring. For larger rings, as the ends required t o react dur-
ing the closure go farther and farther apart, the entropy of ring closure
should become less and less favorable, leading to a small negative
value for the ring opening. When the entropy changes were plotted
against the ring size for cycloalkanes (omitting cyclohexanef), a
linear relationship was found (Fig. 1). The larger rings had more
favorable entropies of ring opening than did the small rings, but
very much less favorable enthalpies.
From a statistical mechanics treatment, it has been shown that
the entropy change of ring closure is [6,7]

ASr = R In {PV/2xVsN} (2)

where P is the probability of ring closure, i.e., the fraction of chains


that will close to form ring structures, V is the total volume of the
system, V, is the volume of a constrained skeletal atom prior to
bond breaking, x is the number of monomer units in the ring, and N
is Avogadro's number.
Three approaches have been followed for calculating the entropy

*The only reason for this omission i s that, if the chain contains six atoms
and the bond angle at each is 120', then there i s a high probability that cycli-
zation to a six-membered ring will occur. Thus the entropy change of p o l y m e r
ization of cyclohexane is an exceptionally small negative value.
160 HIDE0 SAWADA

3 4 5 6 7 8
X

Fig. 1. Relationship between the entropy changes and the number of ring
atoms of cycloalkanes. (Data f r o m Ref. 141.)

change of ring closure: (1)In the Stockmayer-Jacobson development,


P was obtained by assuming a Gaussian distribution of end-to-end
distances and integrating over the volume element Vs [6] ; (2) P was
obtained from the detailed structural model for the chain with fixed
bond length, fixed bond angles, and fixed rotational states-ring clo-
sure is assumed to occur when the end-to-end distance is less than
one bond length (Carmichael-Kinsinger model [7]); (3) another ap-
proach [8 J has been employed in that the equilibrium between simple
rings and chains is described by means of a partition-function cal-
culation of the equilibrium constant relating each of the variously
sized rings to the chain population.
The main defect of these approaches is that they assume that the
separation of the end group is determined only by statistical con-
siderations. These approaches neglect the very real possibility that
the terminal units of a long chain may be held together by ionic
charges or by catalysts [9].
It is worth mentioning a case of strain in acenaphthylene [lo]:
THERMODY NAMl CS OF POLYMERIZATION. I I 161

which polymerizes through the double bond of the five-membered


ring, with the exceptionally high heat of 24 kcal/mole. If the atomic
configuration and valency angles of the naphthalene moiety of this
monomer were identical with that of naphthalene, the length of the
C-C bond in the ethylenic part would be 2.43 A which is much larger
than even the C-C single bond (1.54 A), and, therefore, the angle
strain in this five-membered ring must be very large. On polymeri-
zation, the double bond in the five-membered ring is converted to a
single bond and some of this angle strain is relieved. The large heat
of polymerization of maleimide (- 21 kcal/mole) by addition polymer-
ization through the double bond is also attributed to the changes in
the internal strain of the monomer ring accompanying polymerization.
Calculated values of strain energies for numerous homocyclic and
heterocyclic compounds are assembled in Table 4. Cycloalkenes ap-
pear to have slightly less strain energies than the corresponding
TABLE 4
Strain Energies of Cyclic Compounds [16]

Strain energy, Strain energy,


Compound kcal/mole Compound kc a1/mole

Cyclopropane 27.5 Thiacyclopentane 1.0


Cyclobutane 26.1 Tetrahydropyran 2.2
Cyclopentane 6.1 1,3- Dioxane 2.9
Cyclohexane 0.1 1,4-Dioxane 4.0
Cycloheptane 6.1 Cyclopentene 4.9
Cyclooctane 9.7 Cyclohexene -1
Cyclononane 12.5 Cycloheptene -5
Ethylene oxide -28 Cyclooctene -6
Propylene oxide
Ethylene imine
Ethylene sulfide
- 26.4
23
18.6
Cyclooctatetraene

Benzene --22
15.2

Thiacyclobutane 18.9 Thiophen --16

---
Pyrrolidine 5.5 Furan 8
Tetrahydrofuran 6.7 Pyridine -19
1,3-Dioxolane 7.3
I

cycloalkanes. This finding is surprising at first sight, but it may be


that the insertion of a double bond into the cycloalkane ring is ac-
companied by only a small increase in angle strain. This strain may
be outweighed by a reduction in nonbonded interactions between hy-
drogen atoms, of which there are two fewer in the olefin.
162 HIDE0 SAWADA

However, a small ring, such as cyclopropene, is an example of


two trigonal carbons in a highly strained three-membered ring; an
estimate of a least 8 kcal/mole has been made for the extra strain
energy of cyclopropene over the cyclopropane [ l l ] . The fully con-
jugated ring compound cyclooctatetraene is seen to be appreciably
strained.
The strain energy in benzene is negative, that is, - 22 kcal/mole.
This is conjugation energy which describes the energy associated
with the delocalization of r-electrons. Therefore, the ring-opening
polymerization of benzene has not been achieved. Conjugation ener-
gies should not be equated with resonance energies, for reasons dis-
cussed by Dewar and Schmeising [12].
B. Conformational and Transannular Strain
The cyclopentane ring has a completely eclipsed conformation
with ten "oppositions" between adjacent hydrogen atoms:

Although there is almost no angle strain the planar model, there is


conformational strain caused by the completely eclipsed conforma-
tions. The somewhat higher value for the heat of combustion per
methylene group obtained for cyclopentane (as compared with cyclo-
hexane or cycloheptane) is ascribed to the conformational strain in
the molecule, as shown in Table 2.
In cyclic compounds of more than six members, the strain due to
the deviation of the bond angle is relieved because the ring assumes
a puckered configuration. However, another steric hindrance arises
from the nonbonded intramolecular interaction between the atoms,
including interactions between atoms which formally may be regarded
as being on opposite sides of the ring in the x = 6-11 range. This in-
tramolecular interaction between atoms attached to different parts of
the ring in alicyclic compounds is described as transannular inter-
action. It gives rise to transannular strain in these molecules. In
rings of fifteen or more carbon atoms this effect is absent. In these
molecules the ring atoms arrange themselves into two roughly par-
THERMODYNAMICS OF POLYMERIZATION. II 163

allel chains and thus more nearly resemble aliphatic compounds:

It would, therefore, be expected that above x = 6, AH would be-


come more negative, pass through a shallow minimum, and then in-
crease t o a limiting value. In view of the lesser structure sensitivity
of AS, the AG values might show a dependence on x which is qualita-
tively similar t o that of AH. The free-energy changes for rings of
more than eight members have not been estimated, but the value
might approach zero as the ring size is increased.
C. Steric Effect of Side Group
The influence of side-group steric repulsions on the relative
thermodynamic stability of polymers and on the relative polymeriz-
ability of cyclic monomers will now be discussed in more detail.
In an equilibrium between cyclic monomers and linear polymers,
steric interference between the side groups o r between side groups
and chain atoms will change the equilibrium t o favor the low molecu-
lar weight cyclic monomers. Steric repulsions of the b and c types,
illustrated in Fig. 2, are more serious in an open-chain polymer

Fig. 2. Changes in side-group repulsion on polymerization.


than in a cyclic monomer. Since steric repulsions must raise the
internal energy, interactions of this type will raise the enthalpy of
the polymer relative to the cyclic monomer, and so AH will be made
more positive, and possibly be changed from negative t o zero.
On the other hand, steric hindrance between two side groups on
the same carbon atom should cause the external bond angle between
those groups to widen. This, in turn, should bring about a narrowing
of the opposite skeletal bond angle owing t o a hybridization change.
164 HIDE0 SAWADA

Clearly, narrowing of the skeletal angle will favor the formation of


small rings rather than larger rings. It will also force the various
components of the polymer closer together. If appreciable intra-
molecular crowding is present, small changes in the skeletal angle
will have a relatively large effect on the enthalpy of the polymer,
but less effect on the cyclic monomer. This general effect will be
simply called the =gem-dimethyl effect” in this review.
The gem-dimethyl effect can be quantitatively interpreted in t e r m s
of the thermodynamics of the conformations involved [13,14]. The
calculation of AH for the ring-closure reaction will be considered
first. The effect of substituents on the enthalpy of ring closure is
interpreted in t e r m s of the change in the number of gauche interac-
tions in going from the reactant to the product. The substitution de-
creases the change in gauche interactions on ring closure. The en-
thalpy of the ring-closure reaction is reduced for the substituted
case and the ring closure is thus favored by substitution. This is
reflected in the lower heat of polymerization for the substituted
cyclic monomer.
To estimate the overall effect of substituents on the ring-closure
reaction, the entropy change on cyclization must also be taken into
account. A substituent will not alter the entropy of the cyclic com-
pound relative t o the unsubstituted cyclic compound very much. The
substituent, however, has a much greater effect on the entropy of the
open-chain compound. The substituent restricts rotation of the car-
bon chain due mainly to the increased height of the barriers t o in-
ternal rotations, and thus decreases the entropy of the open-chain
compound. The net effect of a substituent is t o make the entropy of
ring closure more positive. Bulky substituents which greatly re-
strict rotation in the open- chain compound will facilitate ring closure
through the entropy effect.
Thus it can be seen qualitatively that both the entropy and enthalpy
effects of a substituent make the free energy of the ring opening re-
action more positive relative to the unsubstituted case. The effects
of methyl group substitution are to make AHic more positive and ASic
more negative, both causing upward displacements of the AG against
x curve without great alteration of shape as shown in Fig. 3.
The main effect of steric repulsion is t o reduce the rotational en-
tropy of the polymer and to raise the enthalpy of the polymer rela-
tive to the cyclic monomer. The larger the dimensions of the side
group, therefore, the more the equilibrium should be shifted to the
cyclic monomer. Thus, as AH for polymerization changes from nega-
tive toward zero, the ceiling temperature will be lowered until de-
polymerization will occur at any temperature.
THERMODYNAMICS OF POLYMERIZATION. II 165

2 3 4 5 6 7 8
X
Fig. 3. Free energy of polymerization of liquid cycloalkanes. ( a ) 1.1-Di-
methyl substituted; ( b ) methyl substituted; ( c ) unsubstituted; 25'C; x i s the
number of atoms in the ring. (Data from Ref. [41.)

III. HETEROCYCLIC COMPOUNDS


Small [15] has considered the effect of ring size on A G for hetero-
cyclic compounds. The general shape of the relation between A G and
x is expected to be similar to that calculated for the cycloalkanes,
particularly if the heteroatom does not differ too much from carbon
in size and bond angles (for example, in oxygen and nitrogen com-
pounds but not in sulfur compounds, as shown in Table 5).
TABLE 5
Bond Lengths and Bond Angles
~~

C-C length 1.54 A C-C-C angle 109"28'


C-N length 1.47 A C-N-C angle 109"
C-0 length 1.44 A C-0-C angle 111"
C-S length 1.82 A c-s-c angle 100"

Since bond lengths and bond angles of C-N and C-0 do not differ
much from those of C-C bonds, replacement of a carbon atom in a
cycloalkane ring by a heteroatom such as nitrogen o r oxygen would
166 HIDE0 SAWADA

not produce a large change in angle strain. The C-S bond length,
however, is much larger than that of the C-C bond and would pro-
duce a much larger change in structure than replacement of carbon
by oxygen o r nitrogen. In addition, a much larger change in vibration
frequencies is to be expected for replacement by sulfur than for re-
placement by oxygen and nitrogen.
The strain energies of some cyclic compounds a r e shown in Table 4.
The strain energies of ethylene and propylene oxides a r e nearly the
same as the strain energy of cyclopropane, whereas the strain energy
of ethylene imine is only a little lower; the strain energy of ethylene
sulfides is, however, appreciably less than the strain energy of cy-
clopropane.
Amongst the five-membered rings, the order of strain energies
[16] is 1,3-dioxolane > tetrahydrofuran > cyclopentane > pyrroli-
dine > thiacyclopentane. A similar picture is disclosed by the data
for the six-membered rings where the oxa compound is slightly
strained, the dioxa compounds is much more strained, and the aza
and thia compounds are unstrained.
The results of the polymerization of cyclic ethers, esters, ure-
thans, ureas, and imides a r e given in Table 6 [17] . In Table 7 are
collected a number of AH values for ring compounds. A summary
of the results follows:
1. The polymerizability of five- and six-membered cyclic mono-
mers depended markedly on the class of compound.
2. Four-, seven-, and eight-membered rings polymerized in al-
most every case.
3. Substituents on a ring always decreased polymerizability.
4. Substitution of heteroatoms in the ring had in general little
effect on the ease of polymerization as compared to the parent
monomer.
A. Cyclic Ethers
1. 3-Membered Rings. Epoxides polymerize readily owing to
angle strain.
2. 4-Membered Rings. Oxetanes polymerize smoothly with cat-
ionic catalysts. Angle strain and repulsion of adjacent hydrogens
furnish the driving force.
3. 5-Membered Rings. Tetrahydrofuran polymerizes readily ow-
ing to the repulsions of eclipsed hydrogens, as in cyclopentane. Di-
oxolane also polymerizes well. Substituted tetrahydrofurans, namely
the 2- methyl, 3-methyl, and 2-chloromethyl derivatives, do not
polymerize. Substitution in a heterocyclic compound invariably de-
creases its polymerizability, again as predicted for the cycloalkanes
in terms of the gem-dimethyl effect [18].
THE RMODY NAMl CS OF PO LYME R IZATlON. I I 167

TABLE 6
Polymerizability of Five- and Six-Membered Cyclic
Monomers [17I

Polymer iz ability a
Class of monomer 5-Ring 6- Ring

Lactam +
Lactone
Urethan
Urea
Imide
Anhydride

a Experimental conditions: polymerization was

c a r r i e d out at several temperatures between the melting


point of the monomer and 250°C. (+) Polymerization;
(-) no polymerization.
Data f r o m J. Furukawa and T. Saegusa,Polymer-
ization of Aldehydes and Oxides,Wiley-Interscience,
New York, 1963, p.19.

TABLE 7
Comparison of Heats of Polymerization AHpc f o r Ring
Compounds [l1”
~

NO. of atoms in ring


Repeat unit in
polymer 5 6 7 8

-(CH2)nOCHzw 6.2 0.0 4.7 12.8


-(CH2)nCONH- 1.1 2.2 3.8 5.3
-(CHanCON(CHs)- 0.8 -0.5 2.3 3.9
- (CHJnS+ 6.3 0.5 2.5 3.8
- ( CHz)n- 5.2 -0.7 5.1 8.3

=Heats of polymerization expressed in kilocalories per


mole.

4. 6-Membered Rings. Tetrahydropyran, 1,3-dioxane, and 1,4-di-


oxane do not polymerize, in keeping with their strainless chair struc-
tures. However, trioxane polymerizes easily owing to the resonance
stabilization of the linear, initiated and propagating zwitterions:
168 HIDE0 SAWADA

5. Larger Rings. The seven-membered cyclic formal polymer-


i z e s easily, The strain in this and larger rings is that caused by
repulsion of hydrogens a c r o s s the rings. The eight- membered cyclic
ethers, namely tetraoxane and 1,3-dioxocane, also polymerize well.
The variation in the heat of polymerization of cyclic monoethers
is similar t o that found for the cycloalkanes, in that the values be-
come less negative as the ring size increases, x = 3-5, becoming
positive at x = 6. The values a r e shown in Table 8, together with
TABLE 8
Changes in Heat of Polymerization Caused by Replacement of CH2 in
Cycloalkanes with Ether Linkage

No. of Change in AH,


ring atoms Cycloalkanesa Cyclic ethersa kcal/mole
~~ ~~~ ~~~

3 27.0 (Ic) 22.6 ( l c ) 4.4


4 25.1 (lc) 19.3 ( 6 s ) 5.8
5 5.2 ( l c ) 3.0 ( l c ) 2.2
6 -0.7 (Ic) 0.4 (a) -.0.3

aMonomer and polymer states a r e denoted in parentheses: 1, liquid; s ,


solution; c, condensed (liquid o r amorphous solid); and g, gas.

the difference between the heats of polymerization of cycloalkanes


and cyclic e t h e r s having the same number of atoms in the ring. The
difference is large for the three- and four-membered rings where
the effect of introducing an oxygen atom into the ring is to reduce
strain considerably. For the five-membered ring the difference is
small, and almost zero for the six-membered strainless rings. The
heats of polymerization of the formal also follow the same pattern
as the cycloalkanes, becoming increasingly more negative as the
ring size increases, x = 6-8.
B. Lactams
The heat of polymerization of the lactams becomes steadily more
negative as the ring size increases from five- , through six- and
seven-, to eight-membered rings. The change is much less than for
the corresponding cycloalkanes series. Bonetskaya and Skuratov [19]
discussed thermodynamics of polymerization of lactams with 5 to 13
atoms in cycle. In Fig. 4 the AG& values are plotted against x, the
number of atoms in the lactam ring, and from positive A G it is
seen that it is thermodynamically impossible t o polymerize pyrroli-
done at 25” C.
The effects of substituents on the enthalpy of ring closure of capro-
THERMODYNAMICS OF POLYMERIZATION. II 169

20 -

5 7 9 II 13
X
Fig. 4. Free energy of polymerization of lactam. x is the number of atoms
in the ring. (Data from Ref. 1191.)

lactam were discussed in t e r m s of the change in gauche interactions


and configuration of the amide bond [14]. Substituents can also favor
cyclization through the entropy effect. Thus, substituents have the
maximum effect on the five- and six-membered rings, completely
preventing polymerization. Substituted four- membered rings polym-
erize as do substituted seven- and eight-membered rings [14].
According to Table 9, AH makes the main contribution to A G for
TABLE 9
Heats, Entropies, and Free Energies of Polymerization of Lactams [19]

-AG
No. of -AH, -AS, (T= 298.16"K),
ring atoms kcal/mole cal/deg mole kcal/mole

5 -0.1 7.3 -2.3


6 1.7 6.6 -0.1
7 3.3 -1.1 3.6
8 5.4 -4.0 6.6
9 7.8 - 10 10.8
10 5.6 -15 10.1
11 2.8 -20 8.8
12 -0.5 - 25 7.0
13 0 -30 9.0
170 HI DEO SAWADA

seven- and eight-membered rings, but the heat and entropy contribu-
tions are equally important for nine-, ten-, and eleven-membered
rings. On the other hand, AS makes the main contribution to AG for
other lactams with more than twelve-membered rings [19].
C. Lactonee
In the cyclic ester series, four-ring propiolactone polymerizes,
but Carothers [20] found that neither y -butyrolactone nor ethylene
carbonate could be polymerized; all unsubstituted six- ring esters
polymerized, as well as the seven-membered €-caprolactone ; thus
in this series the maximum in AG& occurs at five atoms in the ring.
Brown et al. [21] suggested that cyclic carbonyl compounds with
six ring atoms were markedly less stable than the corresponding
compounds with five ring atoms. They cited the polymerizability of
&valerolactone, as opposed to the nonpolymerizability of y -butyro-
lactone, in support of this thesis.
D. Miscellaneous Heterocyclic Compounds
The polymerizabilities of a variety of atom-bridged bicyclic lac-
tams, lactones, carbonates, ureas, urethans, imides, ethers, and an-
hydrides were found to be dependent on ring strain [22].
The bicyclo [2:2:2] octane and bicyclo [3:2:2] nonane series, in
which the cyclohexane ring occurs in the boat form, underwent
polymerization readily. Monomers of the bicyclo [3:2:11 octane
group, which consists of a chair cyclohexane fused to a cyclopentane
ring, underwent polymerization. The bicyclo [3:3:1] did not polym-
erize, in keeping with the general stability of two fused chairs.
These results a r e summarized in Table 10.
The polymerization of a variety of spiro-oxetanes has been de-
scribed [18] :

The behavior of such spiro ethers resembles that of the oxetane


ring.
THERMODYNAMICS OF POLYMERIZATION. I I 171

TABLE 10
Polymerizability of Bridged Bicyclic Monomers [ 18,22] a

Bicyclic System [ 2:2: 21

M.P P

Bicyclic System [3: 2: 11

Bicyclic System [ 3: 3: l]

Bicyclic System [ 3: 2: 21

aM and P denote monomeric and polymeric forms, respectively.


172 HIDE0 SAWADA

IV. SUMMARY
The heat of polymerization AH to an open-chain polymer affords
a direct measure of the strain energy of the ring. On the other hand,
the entropy change of polymerization AS does not depend on the ring
strain but is susceptible to the configurational influence.
In small rings, AH much affects AG, whereas in medium rings
the effects of AH and AS are comparable although both have small
values. However, AS makes the main contribution to AG for large
rings.

ACKNOWLEDGMENTS

I wish to express my appreciation to Professor K. F. O’Driscoll


of the State University of New York at Buffalo for his interest and
willingness t o read and criticize the manuscript. I a m also indebted
t o the management of Daicel Ltd. for permission t o write this re-
view. Acknowledgment is a l s o due Miss Y. Nishikawa who typed
large sections of the manuscript.

References
[l] F. S. Dainton, K. J. Ivin, and D. A. G. Walmsley. Trans. Faruday SOC.,
56, 1784 (1960).
[Z] A. von Bayer, B e r . , 18, 2277 (1885).
131 S. W. Benson, Themochemical Kinetics, Wiley, New York, 1968, p. 48.
[41 F. S. Dainton, T. R. E. Devlin, and P. A. Small, Trans. Faraday SOC., 51,
1710 (1955).
151 J. Furukawa and T. Saegusa, Polymerization of Aldehydes and Oxides,
Wiley-Interscience, New York, 1963, p. 16.
(61 B. Jacobson and W. H. Stockrnayer, J . Chem. Phys., 18, 1600 (1950).
[7]J. B. Carmichael and J. B. Kinsinger, Can. J . Chem., 42, 1966 (1964).
[8] R M. Levy and J. R. van Wazer, Polymer Preprints, 7(2), 938 (1966).
[9] N. R. Allcock, Heteroatom Ring Systems and Polymers, Academic, New
York, 1967, p. 88.
(101 R. M. Joshi, Makromol. Chem., 62, 140 (1963).
1111 R M. Joshi and B. J. Zwolinski, in Vinyl Polymerization (G. E . Ham, ed.),
Vol. 1, Dekker, New York, 1967, p. 487.
[12] M. J. S. Dewar and H. N. Schmeising, Tetrahedron, 5 , 166 (1959); 11, 96
(1960).
[13] N. L. Allinger and V. Zalkow, J . Org. Chem., 25, 701 (1960).
[14] R. C. P. Cubbon,MakromoZ. Chem., 80, 44 (1964).
[15] P. A. Small, Trans. Faraday SOC., 51, 1717 (1955).
[16] J. D. Cox, Tetrahedron, 19, 1175 (1963).
[17]H. K. Hall, Jr., and A. K. Schneider, J . A m . Chem. S O C . , 80, 6409 (1958).
THE RMODY NAMlCS OF POLYMER I ZATl ON. I I I73

[la] H. K. Hall, Jr., Polymer P r e p n ' d s , 6 ( 2 ) , 535 (1965).


[19]A. K. Bonetskaya and S. M. Sukuratov, Vyskomolekul. Soedin., A l l , 532
(1969).
[20] H. Mark and G . S. Whitby, eds., Collected Papers of Wallace H .
Carothers on Polymerization, Wiley-Interscience, New York, 1940, p. 107.
[Zl] H. C. Brown, J. H. Brewster, and H. Schechter, J . A m . Chem. SOC., 7 6 ,
467 (1954).
[22] H. K. Hall, Jr., J . A m . Chem. SOC., 80, 6412 (1958).
J. MACROMOL. SCI.-REVS. MACROMOL. CHEM., C7(1), 161-187 (1972)

Thermodynamics of Polymerization.

HIDE0 SAWADA*
Central Research Laboratory
Daicel Ltd.
Tsurugaoka, Oi,Irumagun
Saitama, Japan

I. GENERAL ASPECTS . . . .. . . . . . . .. . . . .. .. .. . . . 162


11. FORMATION O F CARBONIUM ION . . .... . . . . ..
. . . 164
A. Ionization Potential . . .. . . .. .. . ..
, ..
.. . . .... 164
B. Proton Affinity .. .... . . ...... .... . . .. . .... 165
C.Acidity . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
, . . 168
D. F r e e Energy Change of Formation of Carbonium
Ion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
. ... .. . .. .. . . . . ...
E. Ions and Ion P a i r s . . .. . . . 171
F. Energetics of Solvation. ... .
. . . . . . . . . .. . . . . . . 172
111. INITIATION O F CATIONIC POLYMERIZATION . . . . . . 174
A. Energetic Consideration of Initiation Reaction
by Halogen Acid . , , . . , . . . . . . . . . . . . . . . . . . . . 174
B. Catalytic Activity in Cationic Polymerization
by Lewis Acids . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
IV. PROPAGATION O F CATIONIC POLYMERIZATION . . . . 177
A. Energetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
B. Heats of Reaction of Cations with Olefins . . . . . . . . . . 179
*Present address: Filter Laboratory, Daicel Ltd., Teppocho, Sakai,
Osaka, Japan.
161

Copyright 0 1972 by Marcel Dekker, Inc. NO PARTof this work m y be reproduced or utilized in ony
form or by ony meuns. electronic or mechanical, including xerography, photocopying, microfilm, and
recording, or by any information storage and retrieval system, without the written permission of the
publisher.
162 H. SAWADA

........
C. Activation Entropy Changes of Propagation 182
D. Thermodynamics of Formation of Zwitterions ...... 184
V. CHAIN TRANSFER AND TERMINATION ........... 186
REFERENCES ............................. 186

I. GENERAL ASPECTS
The polymerizations considered in this review will be restricted
to the cationic reactions, i.e., those in which the growing polymer
chain carries a positive charge-whether a s a free cation o r as the
positively charged member of a partially dissociated ion pair. In
general, ionic polymerizations initially involve the dissociation of
a particle into a positive and a negative ion in an organic environ-
ment. Propagation then occurs through successive additions of
monomeric units to the charged o r reactive ends of the growing
chains. The polymerization steps a r e then analogous to a free
radical mechanism,
The most important reactions of a growing carbonium ion a r e
(a) propagation, in which further monomer units a r e added to the
chain; (b) chain transfer, in which the positive charge is transferred
to another molecule (usually monomer), thus stopping the growth of
one polymer molecule and initiating another; and (c) termination, in
which the carrier ion is destroyed. These three reactions determine
the course of a cationic polymerization.
For a cationic polymerization, the general steps may be written:
Initiation:
ABH AB'H'

AB-H' + M ki % HM'AB'

Propagation:

Termination

MiAB- kt : M,+AB-H'

Transfer:

MiAB-+ M kmt =- M, + HM'AB-


THERMODYNAMICS OF POLYMERIZATION. Ill 163

The main characteristics of cationic polymerization are as fol-


lows [ I - ~ I :

1. The over -all energies of activation for cationic polymeriza-


tions usually fall within the range -10 to +15 kcal/mole, a s shown
in Table 1. If negative, then the reaction rate decreases as the tem-
perature rises. The activation energy of initiation is 5-7 kcal/mole,
and this value is very low compared with that of the usual radical
polymerization (20-30 kcal/mole). Although the cleavage of a cova-
lent bond of an initiator needs a large amount of energy in the radi-
cal initiation, the ionic initiator is polarized and the initiation reac -
tion in the cationic polymerization may not need high energy. There-
fore cationic polymerization reactions themselves are very rapid and
can take place at low temperatures-often almost instantaneously.
High molecular weight polymers a r e usually obtained at low temper -
atures when propagation is much more rapid than termination and
transfer reaction.

Table 1
The Over-all Energies of Activation for Cationic Polymerization [ 7 1

Monomer Initiator Cocatalyst Solvent E (kcal/mole)

Propylene AlBq 10.3


Isobutene TiC14 8
Tic14 - 7.5
SnC14 7
SnC14 3
Styrene SnC14 3
SnC14 4.5
SnCl, 5.5
Tic14 - 8.5
Tic14 CCIsCOzH -1.5
AlClS 1.2
a-Methyls tyr ene SnC14 -3.5

2. The dielectric constant and solvating tendency of the medium


can have a profound effect upon the course taken by a given ionic
polymerization. A medium of low dielectric constant and low solvat-
ing power for the ions will favor the formation of a covalent bond.
A medium of high dielectric constant and high solvating power will
favor the solvent-separated ion pair.
3. The monomers susceptible to polymerization by acids and
164 H. SAWADA

Friedel-Crafts halides are those olefins, cyclic amines, and oxides


where the active intermediate in chain growth is a carbonium ion,
an ammonium ion, o r an oxonium ion, respectively.

R' + CH,=CHY RCH,-CHY

R+ + H)N
:: - R H N ~ c
'~ )

R+ + 0
:
)

The first group has been most extensively studied, but there is in-
creasing interest in the possibilities of polymerization of the cyclic
compounds and these were dealt with in P a r t 11 of this review.
The olefins and vinyl compounds susceptible to cationic polymer -
ization are those having substituents tending to induce an electron-
rich double bond, e.g., isobutene, styrene, and the vinyl alkyl
ethers. Those vinyl monomers having electrophilic substituents,
e.g., vinyl acetate, vinyl chloride, and the acrylic derivatives, do
not respond to cationic initiators, indeed vinyl chloride is so inert
that it has been used as a solvent for the low-temperature cationic
polymerization of isobutylene.
4. The substances able to initiate the cationic polymerization
a r e all strong Lewis acids, i.e., powerful electron acceptors. They
can be systematically classified as:
(a) Simple protonic acids;
(b) Friedel-Crafts halides plus cocatalyst;
(c) Carbonium salts (mainly perchlorates and perfluorides);
(d) Cationogenic substances (e.g., triphenylmethyl chloride).

II. FORMATION OF CARBONIUM ION


A. Ionization Potential
The ionization potential, I, is defined as the energy required to
remove an electron from a molecule o r atom in the dilute gas phase.
R -R++e: AH=I

In principle, any number of ionization potentials correspond to vari-


ous energies of the liberated electron and to different states of the
resulting cation. Unless otherwise qualified, the ionization potential
THERMODYNAMICS OF POLYMERIZATION. Ill 165

is assumed to refer to the most weakly bound electron in the molecule.


The ionization potentials of alkyl radicals are listed in Table 2.
The comparison of methyl and ethyl radical ionizations is shown in
Fig. 1. Carbonium ion stabilization by a methyl group is equal to
35 kcal/mole, in comparison to radical stabilization by the methyl
substituent of 8 kcal/mole.

Table 2
Ionization Potentials of Alkyl Radicals [13 1

Ionization potential
Radical (kcal /mole)

229.4 f 0.7
202.5 f 1.2
200.4 i 1.2
182.2 f 1.2
199.2 f 1.2
182.9 f 1.2
171.8 f 1 . 2
164.2 f 2.3
181.3 f 1 . 2
164.2 f 2.3
192.1 f 2.3
217.9 f 1 . 2
188.2 f 0.7
185.2 f 1.2
177.8 f 1.2
190.2 i 1.8

B. Proton Affinity [8]


It is important to estimate the basicities of various olefins in
terms of proton affinities and carbonium-ion affinities as defined

H' + M
Mi + M
-
in the following equations:

HM'
Mi+l
P = proton affinity

C = carbonium-ion affinity

The proton affinities of olefins will be determined by the follow-


ing four steps [g], the proton in this example being added to carbonl.
166 H. SAWADA

202 Kcal/molo

-CHaCHg.
- -T - - -
A 0 8 Kcal/mols
Flg. 1. A comparison of the ionization potentials for methyl and ethyl radi-
cals. The stabilization energy of the carbonium ion i s shown a s AC+ and

-
that of the radical a s AC..

(1) H+ + e H- AH, = +i

+ -
-
(3) Ha C-CH, R1lC-CH, AH3 = +D
R,’ R,’

(4) t>c -CH, R 1 l


R,/
C +-CH, +e AH^ = -I

The quantity i is known as the electron affinity and is determined


by the reverse procedure of measuring the ionization potential of a
proton, fl is the energy of opening the second half of the double bond,
D is the carbon-hydrogen bond strength, and I is the ionization poten-
tial of the radical formed by addition of hydrogen atom to olefin.
For any nonsymmetrical olefin the proton affinity for addition to
carbon 1 will be different from that for addition to carbon 2. In
Table 3 a r e listed values of the proton affinities for ethylene, pro-
pylene, and isobutene. The carbonium ion affinities (C, and C,
THE RMODY NAMlCS OF POLYMERIZATION. III 167

Table 3
Proton Affinities (kcal/mole) and Atom Affinities (kcal/mole) of Olefins [81

CH2CH2 57.5 97.5 97.5 200 200 152 152 40 40


CH$2H(CH,) 52.5 95 89 179 180 175.5 168.5 42.5 36.5
CH2C(CHs)2 52.0 94 86 165 178 189 168 42 34

corresponding to P, and P,) will follow the same sequence as the


proton affinities.
P, (and therefore C,) increases markedly from ethylene to iso-
butene. This would account for the increase in the ease of polymeri-
zation as the olefin changes from ethylene, which is unaffected by
cationic catalysts, to isobutene, which is highly reactive. As shown
in Table 3, a decrease of the ionization potential of the radical
causes an increase of the proton affinity. Figure 2 shows the linear

9.01

-
90 95 10.0 105

I, (monomer) (ev)

Fig. 2. Comparison of the ionization potential of radical with that of monomer


[lo]. Monomer (radical); (1)C€12=CH2 (-CH2CHS), (2) CH2=CHCH3
( . C H ~ C H ~ C H S )( .3 ) CHFCHCH2CHs (*CH2CH2CH2CHS), ( 4 ) CH2zC- (CH3)2
(.CH?CH(CHJ),), ( 5 ) CHJCH=CHCH~(*CH(CHs)CH,CH,) (6)
( 7 ) CHsCH=C(CH3)2 (*CH(CHs)CH(CHs)2).
(.o),
168 H. SAWADA

relation between the ionization potential of the radical and that of the
monomer [lo]. Therefore the proton affinity in cationic polymeriza-
tion increases a s the ionization potential of the monomer decreases.
It is also seen from Table 3 that P, is much greater than P, for
isobutene. This means that a proton will add to the isobutene double
bond more readily at the CH, than at the C(CH,), end. Thus if the
addition of HX to a double bond involves an initial proton attack,
then the H will add onto the carbon with the greater number of hy-
drogen atoms on it. We may thus interpret Markovnikov's rule in
terms of bond strengths and ionization potentials.

C , Acidity
Brbnsted [ll] defined acids as species which had a tendency to
lose a proton, and bases as species which had a tendency to gain a
proton. An acid-base pair
BH' = B + H+ (1)

consists of the acid (BH') and its conjugate base (B) or, alternatively,
of the base (B) and its conjugate acid (BH'). The equilibrium of
Eq. (1) cannot be observed; all reactions of an acid involve the trans-
fer of the proton to a base which may be the solvent:

Aqueous solution is an important and convenient solvent for the study


of acids, with the water molecule acting a s a base for the removal
of the proton from the acid:

BH' + H,O = H,O+ + B (3)

Although the equilibrium constant for the reaction shown in Eq. (3)
is formally given by

it is usual to omit the water term because in dilute aqueous solution,


the change in concentration of the water due to the formation of the
hydronium ion, H,O+, is negligible. The commonly used expression
for the acid dissociation constant (the acidity constant of acid BH'),
K,, is given as:
THERMODYNAMICS OF POLYMERIZATION. I l l 169

The reciprocal of the acidity constant of acid BH+ is called the ba-
sicity constant of the conjugate base B. The wide range over which
Ka varies makes it more convenient to express the information in
logarithmic form. Thus pKa = log Ka.
Only strong acids can protonate the molecules of weakly basic
solvents. Perchloric acid, hydrogen bromide, sulfuric acid, p-
toluenesulfonic acid, and hydrogen chloride form a series of de-
creasing acid strength in acetic acid, although they are apparently
of equal strength in the more basic solvent, water. Meaningful com-
parisons of acid strengths must be made in a single solvent o r under
such conditions that medium effects a r e minimized.
The acidity function H,,[12] is defined by

It depends, therefore, upon the activity of hydrogen ion aH+and also


upon the ratio of the activity coefficients of a neutral base f, and its
conjugate acid f B H t . By combining Eq. (5) with Eq. (6) and remem-
bering that f = a/c, one obtains an alternative expression for H,

The acidity function is not identical with the pH, it merely becomes
equal to it in dilute aqueous solution.
D. Free Energy Change of Formation of Carbonium Ion
We shall briefly discuss some general aspects of the stability of
carbonium ions. The stability of an ion must be associated with the
free energy change accompanying its formation. However, for the
formation of carbonium ions from alkyl halides in the gaseous phase,
we only know the values of the enthalpy of formation of carbonium
ions which a r e determined from electron impact measurements and
hence the derived values of the heterolytic bond dissociation energy.
Since the corresponding entropy changes a r e usually small, it is
possible to discuss the variations in the free energy change in t e r m s
of the variations in the corresponding enthalpy change. The results
a r e listed in Table 4.
The energy required to dissociate an alkyl halide RX into the ions
R + and X - can be considered to be composed of the energy changes
in the three hypothetical steps:
170 H. SAWADA

Table 4
Enthalpies of Gaseous Ionization R-Br -
R+ + Br- 1131
(Heterolytic Bond Dissociation Energies)

D(R-Br) I(R) AHga


R (kcal/mole) (kcal/mole) (kcal/mole)

Methyl 67 229.4 214


Ally1 45.4 188.2 152
Benzyl 50.5 178.9 147
-p-c1 50.1 183.3 151
-0-Me 48.5 175.5 142
-m-Me 50.5 176.4 145
-p-Me 49.1 172.0 139
-m-NOp 48.4 197.4 164
-m-CN 48.1 197.9 164
-p-CN 49.7 192.8 161

(1) RX --
aCalculated a s AHg = D(R-Br) + I ( R ) - E(Br), with E(Br) = 82 kcal.

Re + X- AH, = D

(2) R. + X.

(3) R' -
+ e + X.
R' + e + X-
R' + X-
AH2 = I

AH3 = E

The quantity AH, is the homolytic bond dissociation energy, and


the term AH2 is called the ionization potential of the alkyl radical.
The quantity AH3 is the electron affinity of the X atom and so is in-
dependent of the nature of R. The ionization potential of the alkyl
radical and the enthalpy changes for formation of carbonium ions
a r e given for a series of alkyl halides in Table 4. These figures
all refer to completely separated ions in the gas phase. As can be
seen from Table 4, the dissociation energies of alkyl compounds
are not expected to be as sensitive to the nature of R as are the
ionization potentials.
The formation of gaseous carbonium ions by heterolysis of neu-
tral molecules requires high energies; heterolytic reactions in the
gas phase a r e not easily observed. Organic reactions involving
carbonium -ion formation in solution are much more frequently
found. More detailed discussions a r e given in Section 11-F.
THERMODYNAMICS OF POLYMERIZATION. I I I 171

E. Ions and Ion Pairs


Winstein [14] has developed the notion of distinguishable stages
of ionization where, between the covalent compound RX and the fully
dissociated (and solvated) ions, 111, there a r e two types of ion pairs:
I, the initimate ion pair, and 11, the solvent-separated ion pair.

RX [R'X-] R'X- R' + X-


I 11 111

Griffiths and Symons [I51 make the distinction more specific by a


slightly different terminology (e.g., species I is called a contact
ion-pair) and by proposing two subspecies in stage II-solvent-
shared (IIa) and solvent -separated (IIb) ion-pairs.
The heat of dissociation of an ion-pair may be determined directly
by heat of dilution measurements, o r may be estimated by application
of the equation

d In K/dT = AH/RT2

where K is the dissociation constant of the ion-pair, and AH is the


heat absorption corresponding to the dissociation of 1 mole of the
ion-pair into ions at infinite dilution. Most of the published figures
were obtained by the second method. However, the calorimetric
determination of AH values should be much more reliable.
For a system of spherical, nonpolarizable ions in a structureless
dielectric medium (dielectric constant, D), the probability of finding
an ion B at a distance r from ion A is

where c is the ionic concentration, 4 r r 2 d r is the volume of shell


considered, Q is the work of separation, k is Boltzmann's constant,
N is Avogadro's number, and T is the absolute temperature. The
work of separation Q is given, according to Coulomb's law, by

4 = -Z,Z,e2/Dr (11)

where Z, and Z B are the ionic charges and e is the electronic charge.
For ions of like sign, P is very small: for oppositely charged ions
P passes through a minimum value when the ionic separation is
172 H. SAWADA

At this point it will be seen that the work of separating the ions is
equal to 2kT, i.e., four times the mean kinetic energy per degree of
freedom. This result was first obtained by Bjerrum [16]. At separa-
tions greater than rmIn the thermal energy of ions is greater than the
electrostatic energy and the ions can be regarded as free. The con-
verse is true for separations less than rmln,the two ions behaving
a s essentially a single species o r ion-pair. Bjerrum [16]proposed
that ions separated by a distance smaller than rmlnshould be treaked
a s ion-pairs. Typical values of rmIn for 1:l electrolytes a r e 3.6 A
in water, 45 8, in acetic acid, and 120 8, in benzene. In media of low
polarity two ions can be regarded as an ion-pair by this definition
even though they a r e separated by one o r more solvent molecules.
A free energy vs. reaction coordinate diagram for the formation of
carbonium ion-pair intermediate is shown in Fig. 3.

t
(3
a
w
z
W

W
W
a
LL

I I
I

Tmin
REACTION COORDINATE

Fig. 3. Free energy vs. reaction coordinate diagram for the formation of a
carbonium ion-pair intermediate.

F. Energetics of Salvation
Carbonium -ion reaction in solution inevitably involves solvated
ions; the carbonium ion and its counter-ions are stabilized by inter-
action with the solvent. Since such stabilization is primarily asso-
ciated with the presence of electric charges, it follows that the ionized
form will be favored relative to the un-ionized form by increased
solvation. All solvents exert some stabilization, relative to the gas
phase.
THERMODYNAMICS OF POLYMERIZATION. Ill 173

The ionization of an organic molecule RX in the gas phase (g) and


in solution (s) is represented in Fig. 4.
R+. .X’ is the transition state of ionization and R+ and X- are the
infinitely separated ions. It is clear that AGO, the standard free en-
ergy of formation of ions in solution from gaseous alkyl halide, must
be the sum of AG; and AGO,. From Fig. 4 it is evident that

AG:

Flg. 4. Free energy cycles for heterolysis of RX.

AG: = AG; + AG; + AGO, (13)

AGf = AG; + AG; + AG: (14)

This simple relation (often referred to as the Born equation) has


been used to calculate the free energy of solvation of the ion. The
free energy change on transferring the ion of radium r and charge e
from the gas phase to the solvent (D)is thus:

According to Eq. (15) the solvatingenergy of an ion willbe greater


in solvents of high dielectric constant. Although the dielectric con-
stant is generally used a s an indication of the solvating power of a
solvent, it is not necessarily a quantitative measure. Specific solva-
tion effects and polarizability a r e also of importance. Furthermore
it would seem that ions of small radius would be most strongly sol-
vated, and this is indeed true for inorganic cations. Therefore, the
174 H. SAWADA

interaction between the ion and the solvent should be strongest for
small r and large D. Generally speaking, the solvation energies of
carbonium ions will therefore be smaller than those of single charged
metallic cations.
Thermodynamic quantities of activation for the ionization of tert -
butyl chloride in various solvents are shown in Table 5. It is clear

Table 6
Thermodynamic Quantities for the Ionlzation of tert-Butyl
Chloride in Various Solvents a t 25°C [17]'

AG" AH" ASa AGG AGg


Solvent c (kcal/mole) (kcal/mole) (e.u.) (kcal/mole) (kcal/mole)

€3 2 0 78.5 19.5 23.2 12.2 6.7 - 135


HC02H 56 21.5 21.0 - 1.7
HCONH2 100 23.5 22.4 - 3.8
CHsOH 32.6 25.8 24.9 - 3.1 2.64 - 125
CHsCOOH 6.2 26.5 25.8 - 2.5 2.50 -124
CzH50H 24.3 27.1 26.1 - 3.2 2.42 -123

aAGi = 155 kcal/mole.

that AGO, is numerically very much smaller than AGO,in all solvents,
and solvents such a s water o r acetic acid may contribute about 125
kcal to the dissociation process. The solvation energies in three
hydroxylic solvents with the exception of water are very similar;
in this case (as in many others) the bulk dielectric constants are a
very poor measure of solvating power.
Finally, it must again be emphasized that carbonium ions in solu-
tion a r e accompanied by an equal number of counter-ions. It is there-
fore impossible to study carbonium ions in solution in isolation: for
example, the free energy of ionization of a molecule Rx will include
the solvation energies of both R* and X-.

III. INITIATION OF CATIONIC POLYMERIZATION


A. Energetic Coneideration of Initiation Reaction by Halogen Acid
Let us consider the reaction of halogen acid HA with olefins.

\ 0 \ +/
HA + ,C=C,- / CH-C, * A-
THERMODYNAMICS OF POLYMERIZATION. I l l 175

The process of formation of cation may be divided into three steps:

(1) HA-H' + A- +El

\ 0 \ +/
(2) Ht + ,C=C,- ,CH-C, -

-
€2

\
(3) ,CH-C,
+/
+ A'
\
,CH-C,
+/
- * * A- - €3
The energy for process (1) is the heterolytic bond dissociation en-
ergy of HA, and the quantity E , is the proton affinity of olefin. The
potential energy between ion pairs separated at a distance r is

c3 = e2/rD (17)
where e is the unit of electronic charge and D the dielectric constant.
According to our convention, the negative sign refers to unlike ions
(attraction) and thus E, is negative.
The total energy change of the system will be

The enthalpy change of initiation of cationic polymerization depends


upon the acid strength of the initiator E,, the proton affinity of the
monomer c,, and solvating powers of the solvent E ~ .
The free energy change of initiation is
AG, = AH - TAS
and upon substitution

AG, = e l - E, - (ez/rD) - TAS (19)


The entropy change, of course, is negative, and hence AH must be
negative in order to make AG, a negative value.
Let us now consider the initiation of propylene by HC1 [18]. En-
thalpy of gaseous ionization of HC1 is about 330 kcal mole and proton
affinity of propylene is 175.5 kcal/mole. If r = 2.5 and D = 1, e3
is 130 kcal/mole. Thus AG, must be positive and no polymerization
d
occurs. The formation of gaseous carbonium ions by heterolysis of
neutral molecules requires high energy (Table 4). However, E , must
176 H. SAWADA

be greatly reduced by solvation. The high energy required for heter-


olysis seems to be offset by interaction between the ions and their
surroundings, in particular solvent molecules and other ions. is
found to be about 25 kcal/mole for HC1 in H,O. Thus, for example,
the halogen acids HC1 and HBr have been reported as effective ini-
tiators for olefins in solvents.
The olefins and vinyl compounds found susceptible to cationic
polymerization a r e those having substituents tending to induce an
electron-rich double bond, e.g., alkyl, alkoxy, and aryl groups.
For these monomers in which E , are large positive quantities, AG,
is negative and hence polymerization will readily occur. Thus elec-
tron-releasing substituents favor the formation of a cationic site:

It is concluded that the rate of initiation increases with increasing


electron-density in ethylene double bond by the electron-donating
effect of the substituent. When E, is small (corresponding to a posi-
tive free energy change), polymerization will not occur. Those vinyl
monomers having electrophilic substituents, e.g., vinyl acetate,
vinyl chloride, and the acrylic derivatives, do not respond to cationic
initiators.
According to Eq. (17) carbonium ions will be less stabilized and
thus more difficult to form as the dielectric constant of the medium
is decreased. A medium of low dielectric constant and low solvating
power for the ions will favor the formation of a covalent bond, and
only the addition of an initiator to monomer will occur. Therefore
polymerization does not occur well by cationic methods in such cases.
B. Catalytic Activity in Cationic Polymerization by Lewis Acids
In the cationic polymerization of styrene catalyzed by Lewis
acids, the cationic activity of Lewis acids is in the following order
[19]:
SbC1, >> TiCl, > AlBr, > SnC1, > FeCl, > ZnC1, > HgC1,
(252) (175) (163) (133) (141) (57) (52)

Kagiya et al. [20]used the shift of the characteristic band of the


carbonyl group in the IR spectrum of the Lewis acid-xanthone com-
plex and reported that an increase in the shift showed 'an increase in
THERMODYNAMICS OF POLYMERIZATION. I l l 177

the acidity of Lewis acid. The values in parentheses in the above


relation represent the magnitude of the carbonyl shift. There is a
good correlation between the catalytic activity and the carbonyl
shift in the Lewis acid-xanthone complex. Therefore it is expected
that an increase in the cationic activity of Lewis acid will arise
from an increase in its activity.
Similarly, the activity of cocatalysts in the polymerization sys -
tem of the same monomer and the same Lewis acid catalyst, e.g.,
in the cationic polymerization of isobutene catalyzed by tin chloride,
decreases in the order [21]:

CC1,COOH > CH,ClCOOH > CH,COOH > CH,CH,NO, >


(pKa = 0.64) (2.87) (4.76) (9)
CH,NO, > C,H,OH > H,O
(11) (10) (15)
As the pKa value of cocatalyst decreases, its activity increases.

W . PROPAGATION OF CATIONIC POLYMERIZATION


A. Energetics
Cationic polymerizations a r e quite exothermic since the reaction
involves the conversion of n-bonds to o-bonds. The heat of polymer-
ization for any particular monomer is essentially the same irre-
spective of the mode of initiation (if the monomer can be polymerized
by both radical and cationic catalysts).
The apparent activation energies a r e given by

Erst, = -2.303Rd log (rate)/d(l/T) (20)


EDp = -2.303Rd log (DP)/d(l/T) (21)

The over -all energies of activation for cationic polymerizations


Erate usually fall within the range -10 to +15 kcal/mole. ED, is
always negative, with values from -3 to -7 kcal/mole. These
quantities are empirically convenient ways of representing the tem-
perature dependence, but must be treated with caution because they
a r e composite quantities and do not necessarily have any exact
theoretical significance.
In the cationic polymerization the counter-ion derived from an
initiator exists near the growing chain end and gives an effect not
178 H. SAWADA

only on the initiation reaction but also the propagation reaction. If


the polymerization proceeded via an ion-pair mechanism,, the value
of kp would be changed by changing the initiator and solvent, even
with a fixed monomer and polymerization temperature.
Also, in this model the monomer molecule is highly polarized in
the transition state of the propagation reaction, and it is expected
that the activation energy and frequency factor would be smaller than
that of radical polymerization. The activation energy of kp in cationic
polymerization decreases with increasing polarity of solvent.
The activation energies EDp and Eratofor the degree and rate of
polymerization, respectively, are obtained as

EDp = E, - Et
where Ei, Ep, and Et are the activation energies for the initiation,
propagation, and chain termination steps, respectively. Et will be
replaced by Etr when termination occurs by a transfer reaction. The
values of Ei and Et are greater than Ep in most cases.
For many polymerization systems Eratois negative and one ob-
serves the rather unusual phenomenon of increasing polymerization
rates with decreasing temperatures. The sign and value of Erst,
vary from one monomer to another. Even for the same monomer,
the value of Eratomay vary considerably depending on the catalyst,
cocatalyst, and solvent employed. The variations in Eratoare a con-
sequence of the differences in Ei, Ep, and Et caused by the differ-
ences in the catalyst and the solvating power of the reaction medium.
It should be noted that, irrespective of sign, the values of Eratoare
generally smaller than in radical polymerizations. The rates of
cationic polymerizations do not quantitatively change with tempera-
ture as much as those of radical polymerizations.
The activation energy EDPfor the degree of polymerization is al-
ways negative because Et is greater than Ep for all cases irrespec-
tive of the mode of termination. This means that the degree of polym-
erization decreases as the polymerization temperature is increased.
EDPhas greater negative values when termination is by transfer re-
actions than when termination is by spontaneous termination or by
combination since the transfer reactions have greater activation en-
ergies. As the polymerization temperature is increased, the mode
of chain breaking will shift from termination to transfer.
Figure 5 shows the dependence of log D P of polyisobutene on the
reciprocal temperature [22]. The plot starts to deviate from linearity
THERMODYNAMICS OF POLYMERIZATION. I l l 179

I 1
4.0 5.0 6.0 7.0 8.0 9.0 10.0

Fig. 5. Temperature dependence of DP of polyisobutene in propane solvent


1221.

around -lOO"C, then bends over to assume a lesser slope. The cor-
responding over-all activation energies a r e -3.54 and -0.22 kcall
mole, respectively. This has been attributed to a change in the
termination step from chain transfer to monomer below -100°C to
chain transfer to solvent above -100°C.
B. Heats of Reaction of Cations with Olefins
The propagation may be regarded as an electrophilic attack by
the ion at the n-electrons of the olefin; accordingly, the ease with
which it occurs is governed by the stability of the ion and by the
basicity of the olefin. In particular, information on the energetics
of the reaction is sparse: an indication of this affinity should be ob-
tained from the calculated values of AHo for the gas-phase reaction
as follows:

\ / I I
R' + /C=C,- R-C-C'
I I
180 H. SAWADA

These values are given in Table 6 for some simple ions and olefins.
In the propagation reaction the olefin monomer will be attacked
by the carbonium carbon of the growing polymer chain. By analogy
with the initiation reaction we might expect that at each step in the
propagation the attack of the carbonium ion on the monomer is at
its tail end, so that+at the end of the growing po+lymer chain there is
always a head CH,C(CH,) and not a tail C(CH,)CH,. This would re-
sult in a head-to-tail propagation process, and would exclude the
tail-to-head, the head-to-head, and tail-to-tail types of propagation.
The nucleophilicity of monomer increases with a decrease in its
ionization potential on the propagation reactions. The same is the
case with the initiation reactions. Figure 6 shows the relation be-
tween the rate constant of propagation in the cationic polymerization
of the styrene derivatives and their ionization potentials.
Evans and Polanyi [8] can give a more detailed calculation of the
propagation process as follows. For the head-to-tail addition they
[8] have calculated this value for the head-to-tail step as 19.5 kcal.
The tail-to-tail addition step was calculated as about 40 kcal exo-
thermic, and the head-to-head step as slightly endothermic.

R 7 1 7 1
R1, + / I
C -CH, + CH,=C, -+ CH,-C-C-CHi Q = 0 kcal
R
2
’ RZ I I
Rz Rz

+ CH,=C\ /
Rl\ + R‘ +2 1
-+
Rl\
CH-CH, CH -CHz-CHz-C \
RZ’ RZ R
Z
’ RZ
Q = 40 kcal

Thus, of the three possible steps for the carbonium-ion mecha-


nism, the tail-to-tail addition is most probable, the head-to-tail
addition is still quite probable, but the head-to-head addition is very
improbable. Since the tail-to-tail addition can only occur in chain
propagation if the head-to-head addition is also involved, this dis-
cussion indicates that for a carbonium-ion mechanism, however the
chaip starts, it will always proceed with the head a s the positive end
CH,C(CH,),, and this will lead to the head-to-tail type of addition
THERMODYNAMICS OF POLYMERIZATION. Ill 181

Table 6
Heat of Reaction of Cations with Olefins
in the Gas Phase [5]

A Ha (kcal/mole)
Ion E thy1ene Propylene Isobutene

Methyl -69.5 -90.5 - 103


Ethyl - 35 - 61 - 71
n-Propyl - 22.5 -45 -56
n-Butyl - 25 -47 -58
Isopropyl -8.5 -30.5 -42
sec-Butyl 0 - 22 -33
tert- Butyl +7.5 -14.5 - 24

with a constant heat of reaction of 19 kcal at each step. This is


shown in Fig. 7 [8]. In this figure we have also included the heats
of reaction for the consecutive steps of the head-to-head, tail-to-
tail mechanism, which alternate between slight endothermicity and
compensating strong exothermicity.

r
-
0,
0

Flg. 6. Relation between the rate constant of propagation in cationic polymer-


ization of styrene derivative and its ionization potential [20]: (1) styrene,
(2) p-chlorostyrene, (3) p-methylstyrene, (4) p-methoxystyrene.
182 H. SAWADA

z
0
a
a +10

v)

5 +30
W
+40
1 2 3 4 5
STEPS

Fig. 7. Heats of reaction for the steps involved in different types of propaga-
tion mechanisms [8]. ( A ) Head-to-tail radical o r carbonium mechanism,
when no steric hindrance is present.. (V)Head-to-tail radical o r carbonium
mechanism, when ateric hindrance is present. ( x ) Head-to-head, tail-to-tail
carbonium mechanism. (0)Head-to-head, tail-to-tail radical mechanism.

The reverse of the propagation reaction steps, i.?., depolymeri-


zation or depropagation, is generally not important in cationic
polymerization because the reactions are normally carried out well
below the ceiling temperatures. At higher temperatures, monomer-
polymer equilibrium would be difficult to detect because of the in-
stability of the chain ion, the importance of transfer processes, and
the fact that decomposition products are likely t o be the monomer
themselves.
No cationic vinyl polymerization has yet been found which con-
forms to equilibrium polymerization. However, equilibrium polym-
erization arising after cationic initiation of tetrahydrofuran has been
reported [23]. @-Methylstyrene, at room temperature, is near its
ceiling temperature, so that the depropagation reaction might be ex-
pected to be fast. Worsfold and Bywater [24] found that it does not
show typical equilibrium behavior when polymerized by BF,-Et,O.
These authors suggested a s explanation that the depropagation re-
action is too slow.

C. Activation Entropy Changes of Propagation


Pepper and Reilly [25] report that the change of activation entropy
in the propagation reaction is based on two phenomena: (a) the entropy
change of immobilization of a free-moving monomer into a polymer
chain, and (b) the difference of entropy in the solvation between the
reactants and the transition state.
THERMODYNAMICS OF POLYMERIZATION. Ill 183

AS' = AS' (immobilization of monomer) + ASf (solvation)

= -28 eu + AS' (solvation) (24)


ASf (immobilization of monomer) is estimated from the radical
polymerization. ASf (solvation) is a positive term, because in the
transition state the electric charge is more diffuse and the degree
of a solvation is lower than that of the more compact dipolar species
corresponded to reactants. In general the activation energy becomes
smaller and the frequency factor becomes larger as the dielectric
constant of the solvent increases.
The value of Ap of cationic polymerization is much smaller than
that of radical polymerization.

Ap (radical polymerization)/Ap (cationic polymerization) = 104-s


(25)
According to Eyring's theory, the propagation constant (kp) is ex-
pressed as
kp = Ap exp (-AEP/RT)
f

= (kT/h) exp (ASf/R) exp (-AEg/RT) (26)


and activation entropy AS* is the function of f 7 and f i , which are the
partition functions of the transition and initial states of the propaga-
tion step, respectively.

AS: = R In (f,/fi) + RT a In (f,/fi)/aT (27)

f r and fi can be reasonably assigned to the partition function of the


monomer in the transition state (f,(m)) and that of the monomer in
the initial state (fi(m)) because other parts of the growing chain are
common in the transition and the initial states of the propagation re-
action. Furthermore, the second term of Eq. (27) is negligible
compared with the first term. So the following equation can be de-
rived:
AS: = R In (f,(m)/fi(m)) (28)
then
Ap (radical polymerization)/Ap (cationic polymerization) z
f,(m) (radical polymerization)/f,(m) (cationic polymer-
ization) (29)
184 H. SAWADA

The vibrational and the rotational modes in the transition state


of the radical propagation and that of cationic propagation are con-
sidered to be
radical propagation cationic propagation
rotation 1 vibration
- * * C* . - -
M vibration ...c+.. .M'--'A-
/ vibration
/
rotation vibration

In the case of cationic propagation the rotation of the monomer unit


in the transition state is greatly restricted by the counter-anion.
Therefore,
Ap (radical polymerization)/Ap (cationic polymerization)
f 7(m) (radical poly merization)/f 7(m) (cationic polym -
erization) = frot2fvl,,/fri,,3 = frotz/fvIbz =

where k and h are Boltzmann's and Planck's constants, respectively,


I is the moment of inertia, and vi is the vibration frequency of the
monomer unit. I = g cm2 and v 1013 sec" are inserted in the
above calculation.
This result coincides rather well with experimental results [26].
It is therefore concluded that A of the cationic polymerization by
the ion-pair mechanism is mu,! smaller than that of the radical
polymerization since the counter-ion restricts the mobility of an
entering mohomer in the transition state of the propagation reac-
tion. On the other hand, in the polymerization of styrene by per-
chloric acid, Ap is as large as that of the radical polymerization;
that is, the propagation reaction proceeds through a free end.
Ap of the free-ion mechanism is larger than that of the ion-pair
mechanism, and Ap decreases with an increase in the polarity of a
solvent. The polymerization in a polar solvent shows an Ap value
with the same order as the radical polymerization.
D. Thermodynamics of Formation of Zwitterione
In cationic polymerization initiated by uncharged Lewis acids,
the formation of zwitterions in the initiation process has been postu-
lated [27]. Monomer addition to the cationic chain ends may lead to
macrozwitterions, i.e., polymer chains with a negative charge at
one end and a positive charge at the other. However, thermodynamic
THERMODYNAMICS OF POLYMERIZATION. Ill 185

consideration during propagation steps has been used as an argument


against this type of zwitterionic polymerization mechanism.
Considering the cationic polymerization of isobutene by AlCl,, we

AICl,(sol) + C,H,(sol) -
may write the over -all reaction as
XlC1, - CH,~(CH,),(SOI)

Our reaction of interest will consist of the following five steps, all
AHA

at the same temperature.

(1) AlCl,(sol) + C,H,(sol) - AlCl,(g) + C,H,(g) AH,, = 15 kcal/mole


(2) C4H,(g) - CH2C(CH,)2(g) p, = 50 kcal/mole
(3) CH,C(CH,),(g) - eH,6(cH3)2(g) I, + A, = 140 kcal/mole
(4) AlC1, + cH,6(CH,),(g) - a C 1 , CH,t(CH,),(g)
*

-D, = -60 kcal/mole


(5) fiCl,CH,6 (CH,),(g) - aCl,CH26(CH,),(sol) -AH82

The t e r m s AH,,, D, I, and A correspond to the heat of solvation,


the bond dissociation energy, the ionization energy, and the elec-
tron affinity, respectively; these terms were calculated theoreti-
cally by Plesch [28]. Since AH,, is supposed to be much smaller than
130 kcal/mole, the formation of macrozwitterions in cationic polym-
erizations is easily disturbed by side reactions.
However, such simple energetic considerations are not in general
a valid argument against the formation of macrozwitterions [29];
propagation macrozwitterions need not be linear chains (A) with an
increasing distance between the two ionic epd groups. They may be
cyclized to form ion-pairs from the two ionic chain ends (B) (Fig. 8).
If the chain ends form ion-pairs, macrozwitterions may greatly en-
hance the first propagating step.

(A) ( B)

Flg. 8. Macrozwitterions a s ( A ) linear chains and ( B ) ion pairs.


186 H. SAWADA

V. CHAIN TRANSFER AND TERMINATION


It is very difficult to get a high molecular weight polymer above
room temperature in cationic polymerization. This is due to the
frequent transfer reaction to transfer agents, monomers, and im-
purities. This transfer is the most important process governing
the molecular weights in carbonium ion polymerizations.
The ratio kp/km is somewhat temperature dependent; the activa-
tion energy of the transfer reaction is higher than that of propaga-
tion by 10 kcal/mole at most, but usually is about 2-3 kcal/mole.
Therefore the polymerization must be carried out at the'lowest
possible temperatures in order to obtain products of the highest
molecular weights. However, this requirement does not present
the practical difficulties which might be expected; because the over-
all activation energies for most cationic polymerizations are in the
range -5 to +10 kcal/mole, the polymerization rates do not change
rapidly with temperature. We can also see that the carbonium mech-
anism offers a possibility for the occurrence of polymerization at
the lowest temperatures, whereas a free-radical process would be
inevitably slowed down by extreme cold.
Since the spontaneous termination reaction cannot be distinguished
from the termination reaction by impurities, the termination con-
stant ratio (kt/kp) is only an apparent value and quantitative discus-
sion has not been carried out. Moreover, the existence of a true
termination reaction is observed in only a few systems, and many
systems seem to have no termination reaction [30].

Acknowledgments
I wish to express my appreciation to Professor K. F. O'Driscoll,
University of Waterloo, for his interest and willingness to read and
criticize the manuscript. I am also indebted to the management of
Daicel Ltd. for permission to write this review. Acknowledgment
is also due Miss Y. Nishikawa who typed large sections of the manu-
script.
References

[ l ] M. L. Burstall and F. E. Treloar, in The Chemistryof Cationic Polymen-


zation (P. H. Plesch. ed.), Pergamon, London, 1963, Chap. 2.
[2]F. S. Dainton, in Cationic Polymerization and Related Complexes
( P . H. Plesch. ed.), Heffer, Cambridge, 1953.
[3] T. Higashimura. in Kindai Kogyo Kagaku,Vol.16 (R.Oda et al., eds.).
Asakura, Tokyo, 1966, Chap. 4.
THERMODYNAMICS OF POLYMERIZATION. I l l 187

[4] D. C. Pepper, in Friedel-Crafts and Related Reactions'(G. A. Olah, ed.),


Wiley (Interscience), New York, 1964, p. 1293.
[5] A. M. Eastham, in Encyclopedia ofPolymerScience and Technology, Vol.
3, Wiley (Interscience), New York, 1965, p. 35.
[6] T. Higashimura, in Structure and Mechanism in Vinyl Polymerization
(T.Tsuruta and K. F. O'Driscoll, eds.), Dekker, New York, 1969, Chap.
10.
[7] Ref. 2, p. 158.
[ 8 ] A. G. Evans and M. Polanyi,J.Chem. SOC., 1947, 252.
[9] A. G. Evans, in F i b r e s from Synthetic Polymers (R. Hill, ed.), Elsevier,
New York, 1953, Chap. 3.
[ l o ] T. Kagiya, Y. Sumida, and T. Nakata, Bull.Chem. SOC. J a p . , 41, 2239
(1968).
[ l l ] J. N. B r h s t e a d , Rec.Trav. Chim. Pays-Eas.42, 718 (1923).
(12) L. P. Hammett, Physical Organic Chemistry, McGraw-Hill, New York,
1940, p. 267.
[13] D. Bethel1 and V. Gold,Carbonium Ions, Academic, New York, 1967,
p. 67.
[14] A. Ledwith, M. Hogo, and S. Winstein, Proc. Chem. Soc., 1961, 241.
[15] T. R. Griffiths and M. C. R. Symons,Mol. Phys., 3, 90 (1960).
[16] N. Bjerrum, Kgl. Danske Videnskab. Selskab, Mat. F y s . Medd., 7, (9),
(1926).
[17] Ref. 1, p. 26.
[18] Ref. 3, p. 12.
[19] S. Okamura, T. Higashimura, and H. Sakurada, Kogyo Kagaku Zasshi 61,
1640 (1958).
[20] T. Kagiya, Y. Sumida, and T. Nakata, Bull. Chem. SOC. Jap., 41, 2247
(1968).
[21] K. E. Russell, T r a n s . Faraday SOC.,48, 114 (1952).
[ 2 2 ] J. P. Kennedy and R. M. Thomas, International Symposium on Macro-
molecular Chemistry, Montreal, Canada, 1961.
[23] M. P. Dreyfuss and P. Dreyfuss,Polymer, 6, 93 (1965).
[24] D. J. Worsfold and S. Bywater, J. A m e r . Chem. SOC.,70, 4917 (1957).
[25]D. C. Pepper and P. J. Reilly,Proc. Roy. SOC., S e r . A , 291, 41 (1966).
[26] N. Kanoh, T. Higashimura, and S. Okamura, Makromol. Chem., 56, 65
(1962).
[27] W. Kern and V. Jaacks, J . Polym. Sci.,48, 399 (1960).
[28] Ref. 3, p. 19.
[29] V. Jaacks and N. Mathes, Makromol. Chem., 131, 295 (1970).
[30] Ref. 2, Chap. 30.
J. MACROMOL. XI.-REVS. MACROMOL. CHEM., C8(2), 235-288 (1972)

Thermodynamics of Polymerization. IV.


Thermodynamics of Equilibrium Polymerization

HIDE0 SAWADA
Filter Laboratory
Daicel Ltd.
Teppo-cho, Sakai, Osaka, Japan

I. POSSIBLE T Y P E S O F EQUILIBRIUM POLYMERI-


ZATION .................................. 238
11. SOME CASE STUDIES O F EQUILIBRIUM
POLYMERIZATION . . . .. .. . . ..........
.. . . .. 243
.. . .. .... ... . . . ...
A. Vinyl P o l y m e r i z a t i o n s . . . 244
. . .... . ... .. . . .
B. Ring-Opening P o l y m e r i z a t i o n s . 247
. . . .. . . . .. .. .
C. P o l y m e r i z a t i o n of Aldehydes . . .. 262
111. TRANSITION PHENOMENA IN EQUILIBRIUM
POLYMERIZATION . . . . . ... . . . .. . . . .... . ... . 264
IV. MOLECULAR WEIGHT DISTRIBUTION . . , . , . . , . . . . 270
A. E q u i l i b r i u m P o l y m e r i z a t i o n . . . . . . . . . . . . . . . . . 270
B. Living P o l y m e r i z a t i o n . . . . . . . . . . . . . . . . . . . . . 276
V. THERMODYNAMICS O F EQUILIBRIUM POLYMERI-
ZATION..... ............................ 280
ACKNOWLEDGMENT.. . . ... , . ... . . .. .. . .... . 285
REFERENCES ............................ 286

235
Copyright ill, 1972 by Marcel Dekker, Inc. All Rights Rejerved. Neither this work nor any part
may he reproduced o r transmitted in :my form o r by any means, eleclronic or mechanical, includ-
ing photocopying. microfilming. and recording, or by any information storage and retrieval sys-
tem, without permi\sion in writing from the publi\her.
236 SAWADA
-

I. POSSIBLE TYPES OF EQUILIBRIUM POLYMERIZATION


In many cases polymerization proceeds under conditions of equi-
librium between polymer and monomer. In some cases an initiator
enters into the polymerization equilibria. A general theory of equi-
librium polymerization was originally developed by Tobolsky in a
long series of papers [l-61. The particular approach used in the
treatment presented here is that developed by Tobolsky and Eisen-
berg [51.
Three basic types of possible equilibrium polymerizations are:

Case 1
Kl
XY + M LXMY
K2
XMY +M A XM,Y
...
Kn
XMn-,Y +M XMnY
. ..
Case I1
Kl
M & M*
K2
M* +M - M,*
L

(The asterisk represents an activated state such as a diradical o r


a zwitterion.)

Case I11
K1
M + M -M2
K2
M, + M M,

...
THERMODYNAMICS OF POLYMERIZATION. IV 237

Polymerization proceeds under conditions of equilibrium between


polymer and monomer in Cases I1 and 111. In Case I an initiator
enters into the polymerization equilibria. The experimental prob-
lem is to place monomer (or monomer plus initiator) of known con-
centration [M,,] (and XY, where present) in a sealed system. The
system is kept a t constant temperature until equilibrium is attained.
At this point the system is quenched, and the sample analyzed for
the M, P, and X Y .
Let u s consider the detailed mathematical s t e p s in the treat-
ment of Case 111. The overall kinetic scheme is then, in the absence
of transfer

...

.
where K , , K,, , . , K, are the equilibrium constants for the initia-
tion and succeeding propagation reactions respectively. All the ac-
tivities appearing in the equilibrium equations are defined in moles
per kilogram. The problem of volume contraction may, however, be
circumvented by expressing all the concentrations in moles p e r unit
m a s s of solution. The subscript e denotes that all the relevant con-
centrations have the equilibrium values.
The equilibrium concentration of n-mer is

The total concentration of polymer molecules, N, will be given by


n-1
N = ll Kk[M,Ien
k =I

The assumption will be made that the constants K,, K 3 , . .., K,, .. .
are all equal and can be represented by K,. Thus
238 SAWADA

and hence the total equilibrium concentration of polymer molecules


is

N = 2KIe1[Ml]t
n=z
= K1[M1],2(1 + K,[M,], + K12[M1],2 + * * (4)
K,[M, 1,"
- 1 - K,[M1l,
and the total equilibrium concentration of monomer segments in-
corporated in polymer, W, is given by

W =cnK,'''[M,],n
m

n=a

Consequently the number-average degree of polymerization, P is

P = W/N
1
-
1 - KI[M,I*
Let us assume that K, = K,,, = . ..
= K m for some value of j ->I+,.
It has been pointed out by Szwarc [7] that when the system attains
its state of equilibrium, the following relation is obeyed.

where [M,,] denotes the total amount of the monomer introduced into
a unit volume of the solution, [MI, the equilibrium concentration of
the monomer, [Po*] the concentration of all the polymer molecules,
[Q,*] the concentration of those polymers which have a degree of
polymerization i < j, and [R,]the amount per unit volume of the
monomer incorporated in these polymers (i < j). For a high number-
THERMODYNAMICS OF POLYMERIZATION. IV 239

average degree of polymerization the ratio ([P *I - [Q,*])/([M,,] -


[R,]- [MI,) approaches zero and then K, = l/[M],.
Tobolsky [5] introduced two physically reasonable approxima-
tions
K, = K
& = K-3 = & = . s. = Kn = .. . = K-3 (a)
K, = K, = K-3 = . * * = K, = * - .= K-3 (b)
Whether a given equilibrium falls under approximation (a) o r (b)
can sometimes be determined from chemical intuition.
The detailed mathematical steps in the treatment of Case IIIa
are as follows:
M +M -- K
M,

M, +M -K3
M,
Kg
M, +M & M,

The total concentration of polymer molecules is

N = 2 [MnIe
n 2

and hence the total equilibrium concentration of monomer segments


incorporated in polymer, W is given by
240 SAWADA

Thus the number-average degree of polymerization, P, is

When equilibrium concentrations of monomer and initiating species


a r e attained in a polymerization system, the degree of polymeriza-
tion is dependent upon one equilibrium constant only, K,, and is in-
dependent of the rate of initiation. This is in direct contrast to the
conditions in an irreversible polymerization (or in a polymerization
with some depropagation but in which equilibrium conditions are not
attained) when the degree of polymerization depends upon the rate
of initiation.
Consequently the initial monomer concentration, [M,,], is

Combining Eqs. (10) and (11) and setting P - 1/P = 1 (valid for
P >> l ) , it is possible to eliminate [MI, and obtain

From Eq. (lo), if P >> 1, we find


K, G l/[MI, (13)
The above results are equally applicable to the Case 11% where all
constants a r e identical by assumption (b). Thus we have
P ([MolK,)”2 = ([%]/[M]e)1’2 (14)

Let us consider the detailed mathematical steps in the treatment


of Case Ia.
K
XY +M XMY [XMY], = K[XYI,[MI,
K3
XMY + M -2XM,Y [XM,Y], = K3[M],[XMYle
K3
XM,Y + M & XM3Y [XM,Y], = (K,[M1,)2[XMY1,
...
THERMODYNAMICS OF POLYMERIZATION. IV 24 1

The total concentration of polymer molecule is

The total equilibrium concentration of monomer segments incorpo-


rated in polymer is given by

and hence the number-average degree of polymerization is


W 1
p =- =
N 1 - &[MI,

Consequently the initial monomer concentration is

Therefore the initial concentration of initiator is

In addition, the following derived relationship is also useful

p =- [%I - [MI, (19)


- [xyle
[ ~ y o ]

In Table 1 a r e summarized the necessary relations for various cases.


Let us now examine the occurrence of maxima in the molecular
weight-temperature curve for equilibrium polymerization [8]. Sub-
242 SAWADA

Table 1
Types of Equilibrium Polymerization [ 51

Case P [M, 1

stituting the expression for [XY], from Eq. (18) into Eq. (19), in-
serting 1/& for [MI, (which is valid P >> 11, and simplifying, we find

If P > &/K (this is true for Case Ib in which K, = K), maxima


are impossible since essentially only a propagation reaction is in-
volved; in a case in which P z K,/K a maximum height might pos-
sibly occur, but to ascertain the fact, Eq. (20)would have to be
solved for P and differentiated with respect to T. Finally, if P/& <<
1/K, and this is the most important case since here we are most
likely to observe maxima, we can neglect P/K, with respect to 1/K,
and obtain the simplified expression

Differentiating this expression with respect to temperature after


writing K = exp(AS'/R) exp(-AH'/RT) and the equivalent for K, and
setting d(P2)/dT = 0, which is an acceptable criterion for the occur-
rence of a maximum in P vs. T, we obtain

It should be stressed here that a close relationship exists between


the temperature at which the maximum in P occurs, T,, and the
transition, i.e., 'ceiling" o r "floor" temperature, T,. Based on a
previous development of expressions for transition temperatures,
THERMODYNAMICS OF POLYMERIZATION. IV 243

this relationship can be written a s

_
T m
- AS," + R In[&]
Tt
AS," + R In
[&](AH' - AH,')
AHo

If AH: is much smaller numerically than AHO, these two tempera-


tures a r e seen to be very close together. Elsenberg [a] has con-
sidered this problem in more detail, and have applied it to equilib-
rium polymerization of elemental sulfur shown in Fig. 1.
&

TEMPERATURE ( O K )
Fig. 1. Degree of polymerization vs. temperature for the equilibrium polym-
erization of sulfur [3,6].(0):Experimental points. (-) : Calculated values.

11. SOME CASE STUDIES OF EQUILIBRIUM POLYMERIZATION


The establishment of an equilibrium between a monomer and its
addition polymer is not usually an easy matter. This is because the
active centers involved in the formation of the polymers a r e usually
short-lived. To establish a reasonably mobile equilibrium between
monomer and polymer, it is necessary (a) that a certain concentra-
244 SAWADA
-

tion of active centers is continually present in the polymerization


system, and (b) that in such a system only long-chain polymer is
present at equilibrium and there is no danger of side reaction; in
addition, (c) it i s desirable that the polymerization proceeds homo-
geneously.
However, the fundamental principles of equilibrium polymeriza-
tion have been presented and applied somewhat concisely to several
cases. It might be profitable to consider some actual cases of
equilibrium polymerization. The particular cases to be discussed
a r e chosen so as to illustrate a variety of equilibrium polymeriza-
tion type.
A. Vinyl Polymerizations
In treating equilibrium polymerization of vinyl monomers, par-
ticularly methyl methacrylate, Dainton and Ivin [9] and Bywater [lo]
implicitly treated the problem as a one-constant case, and derived
the following relation by other methods:

However, Tobolsky [6]determined P in this system and treated the


problem as a two-constant case where

In fact, from the measured value of P, a value of K, can be deter-


mined which i s about 3 X 10“ (cf. K, = 3).
Equation (13) has been used with success for methacry lonitrile
[ l l ]as well as methyl methacrylate, in both of which initiation of
either polymerization o r depolymerization proceeded by a radical
mechanism. The anionic homopolymerization of tert-butyl vinyl
ketone appeared to be an equilibrium polymerization and the effect
of temperature on the equilibrium monomer concentration was in-
vestigated [12].
The studies of McCormick [13] and of Worsfold and Bywater [14,
151 illustrate applications of Eq. (13) to such systems as a-methyl-
styrene-poly(a-methylstyrene) and styrene-polystyrene. a-Methyl-
styrene in tetrahydrofuran reaches equilibrium with its polymer in
the presence of sodium naphthalene used as an anionic initiator at
all temperatures [13]. On the other hand, styrene-polystyrene sys-
tem initiated by butyllithium reaches equilibrium with the polymer
in the temperature range of 100-150°C [15]. The results are shown
graphically in Fig. 2 as a plot of log K against 1/T. The equilibrium
TH E R MODY NAM I CS OF PO LY M E R I ZAT ION. I V 245

9c

-
+
10
8.C

5
Q
I
7c

6.(

24 2 5 26

1031~

Fig. 2. F r e e energy functions plotted against re ciprocal t emper at ur es f o r


the equilibrium polymerization of s tyrene [15]. (0):Benzene. (+) Cyclo-
hexane.

concentrations were found to differ in the two solvents benzene and


cyclohexane, but the calculated free energies of the polymer-mono-
m e r systems were the same after corrections had been applied for
the different heats of solution. F r o m the free energy data the heat
and entropy of polymerization of styrene were calculated and found
to agree with the combustion data and many other calorimetric esti-
mations.
Tobolsky, Rembaum, and Eisenberg [16] investigated the equi-
librium polymerization of a-methylstyrene a t 0°C initiated by
naphthalenesodium and diphenylacethylenesodium. With both ini-
tiators the extrapolated equilibrium monomer concentration has
been found to be equal to 0.89 mole/kg. Initiation by naphthalene-
sodium was shown to be an equilibrium electron-transfer reaction
with a high equilibrium constant. In the sodium diphenylacethylene
system the initiation consists of the equilibrium step involving bond
formation between the monomer and initiator, and the equilibrium
constant has a sufficiently low value to be determined practically.
When the initiation equilibrium was taken into account, good agree-
246 SAWADA

ment was found between theoretical calculations and experimental


results, With both initiator at O"C, as well as when cesium was used
in place of sodium, the same equilibrium monomer concentration
was found. Polymerizations of substituted a-methylstyrenes were
carried out using anionic initiator [17]. Data of heats and entropies
of polymerization from equilibrium studies are given in Table 2.

Table 2
Heats and Entropies of the Polymerization of
Substituted a-Methylstyrenes

Monomer
~~

a-Methylstyrene 6.96 24.8 13


8.02 28.75 14
7.47 26.5 a
8.50 30.6 b
p-tertButyl a-methylstyrene 7.1 25.5 17
p-Diisopropenyl benzene 7 .O 23.0 17
m-Diisopropenylbenzene 8.2 25.8 17
1,3,5-Triisopropenylbenzene 5.9 17.1 17
1,2,4-Triisopropenylbenzene 4.7 13.6 17

aA. Vrancken, J. Smid, and M. Szwarc, Trans. Faraday Soc., 58,2036 (1962).
bH.Hopff and H. Liisi, Makromol. Chem., 45,169,183 (1960).

The equilibrium pressure of methyl methacrylate vapor over its


polymer has been determined by Small [18] and Ivin [19]. This pres-
s u r e is given by

In P = (AH/RT) - (AS/R) (24)

Small [18] measured the equilibrium pressure of monomer at tem-


peratures between 100 and 160°C. The attainment of equilibrium was
induced by radicals derived from the photodecomposition of benzoin,
the equilibrium partial pressure of monomer being measured by the
dew-point method. On the other hand, Ivin [19] used a kinetic method
to obtain the equilibrium pressure of monomer over polymer at tem-
peratures between 96 and 142"C, the approach to equilibrium being
induced by direct absorption of ultraviolet light. Values of AH are
AHgc = -19.5 kcal/mole [18], AH,, = -13.4 kcal/mole [19], and the
THERMODYNAMICS OF POLYMERIZATION. IV 247

corresponding entropy value ASlc = -27.8 eu [19]. Equilibrium


pressures of ethyl methacrylate over its polymer have been deter-
mined from 102 to 138.5"C [20]. The equilibrium pressures are
slightly lower, and the derived heat and entropy changes slightly
higher than for the methyl ester.
It should be stressed that the results of such studies as outlined
above are independent of the mechanism of the reaction, and although
they were derived from studies of anionic o r radical polymeriza-
tions, they apply equally well to any polymerization process involv-
ing the investigated monomer and polymer. The nature of the sol-
vent does, however, affect the results since it modifies the f r e e
energy of the initial and final states.
B. Ring-Opening Polymerizations
1. Tetrahydrofuran. Meerwein [21] observed that the maximum
possible degree of conversion to polymer at room temperature was
around 70% and therefore suggested that the propagation reaction in-
volved oxonium ions with a monomer-polymer equilibrium(Scheme 1).

3 4" 1
OdCH
24
0
CHj-CH,

\
CH,-CH,
+ rFC>o
CH,-CH,
S~O-(CH,$I<
n.1
CH2-CH,
1
CH2-CH,

Scheme 1.
Several papers have recently appeared concerning the equilibrium
polymerization of tetrahydrofuran under the influence of cationic
catalysts such as phosphorus pentafluoride [22], triphenylmethyl
hexachloroantimonate [23], trialkyloxonium salts [24,25], o r benzene-
diazonium hexafluorophosphate [26]. The polymerization process is
reversible and an equilibrium between monomer and polymer may be
established in the undiluted polymer [22-24,271 o r in solution [25,28].
The data in Table 3 show clearly that the degree of conversion of
monomer to polymer reached a constant (equilibrium) value at 25°C
independent of initiator concentration [23]. The conversion a t equi-
librium is given by

conversion = ([M,] - [M],)/([M],) (25)

Tobolsky [25] found that [MI, i s indeed a constant, namely 2.62


mole/l at 0°C provided that [M,,] is larger than 5-6 mole/l. Above
an initial monomer concentration of 5-6 mole/l, agreement between
the calculated value from Eq. (25) for [MI, = 2.62 mole/l and actual
248 SANADA

Table 3
Polymerization of Tetrahydrofuran at 25'C' [ 231

Weight of polymer
[ ph,C+SbCl,-] X l o 3 recoveredb [VlC
M (9) (dug)

1.37 6.44 3.55


2.06 6.85 2.88
2.74 6.35 2.70
3.53 6.71 2.06
4.93 6.57 1.76
8.00 6.29 1.64

'10.0 mi of THF used for each experiments.


bEach reaction left for approximately 48 hr.
cMeasured in benzene at 25.0"C.

equilibrium conversions i s within experimental error. When equi-


librium conversions at various temperatures are plotted against
temperature (Fig. 3), extrapolation gives a ceiling temperature of
84°C [26].

0 30 60 90
TEMPERATURE VC)
Fig. 3. Ceiling temperature in tetrahydrofuran polymerization 1261.
THERMODYNAMICS OF POLYMERIZATION. IV 249

The equilibrium concentrations [MI, at different temperatures


have been used to derive the heat and entropy of polymerization us-
ing the simple formula

ln[M], = (AH/RT) - (AS/R) (26)

A plot of ln[M], vs. 1/T should be a straight line of slope AH/R ac-
cording to Eq. (26). Figure 4 shows t h i s plot for data obtained in
Dreyfuss's work [26], and the value of AH obtained in this manner
is -4.58 kcal/mole and the corresponding entropy change AS is
-17.7 eu. Results of thermodynamic parameters calculated from
equilibrium concentrations of tetrahydrofuran by several workers
a r e shown in Table 4.

24-

i 16-
-

I I I \
08'
26 29 32

lo3 I T
Fig. 4. Equilibrium monomer concentration a s a function of t e m p e r a t u r e
for the equilibrium polymerization of tetrahydrofuran [ZS].

2. c-Caprolactam. The process of polymerization of c-caprolac-


tam initiated by water may be divided into three steps: ring open-
ing (K), condensation (K2),and addition (K3).Both ring opening (the
hydrolysis of caprolactam to the linear €-amino caproic acid) and
addition (the direct attachment of a molecule of c-caprolactam to an
endgroup of a polymer molecule) are the principal reactions accord-
ing to which €-caprolactam is converted, whereas the condensation
250 SAWADA

Table 4
Thermodynamic Constants of THF Polymerization

-AH -AS TC
Standard statesa (kcal/mole) ( 4 ("C) Ref.

Is 4.28 f 0.2 17.0 f 0.6 83 22


Is 4.58 17.7 84 26
Ic 2.97 9.75 80 f 3 27
Is 5.3 f 1.0 60 - 70 23
C
Ic 5.5 20.8

aMonomer and polymer states: 1, liquid; s, solution; c, condensed.


bCeiling temperature in bulk polymerization.
CB.A. Rosenberg, E. B. Ludvig, A. R. Gantmakher, and S. S. Medvedev,J. Polyrn.
Sci.,Part C,16,1917(1967).

reaction results in an interlinking of linear species by the reaction


between an amino end group and a carboxyl end group forming an
amide group and a molecule of water (Scheme 2).

co NH

K
HOOC(CH,),NH, 4- HOOC(CH, NH, 1,
HOOC(CH,), NHCO(CH,), NH2 -k H,O

n 0, 1, 2, 3, 4, - - - - - - -
Scheme 2.
THERMODYNAMICS OF POLYMERIZATION. I V 25 1

At a given temperature, for a given value of [q] and [&I, P and


[MI, a r e available from data in the literature [29]. Referring to
-
Eq. (10) and Table 1, K, was obtained from P = 1/(1 K,[M],).
[XI, was obtained from P = ([%I - [MI,)/([%] - [XI.). Finally,
K was obtained from [&] = [X],(l + K[M],P). The constants K and
K, may be found by application of the van't Hoff equation with the
result [5]

K = exp(ASo/R) exp(-AH"/RT), AS" = -6.8 eu, AH" = +2.24


kcal/mole
K,=exp(AS,"/R) exp(-AH,"/RT), AS," = -7.0 eu, AH," =-4.03
kcal/mole

corresponding to the initiation and propagation reactions, respec-


tively.
The overall reaction of c-caprolactam initiated by an alkyl
amine o r an organic acid can be written as in Scheme 3. To illus-
trate the problem, let us consider the reaction
K3
RNH, + CL LR(CL)NH,

1, 2, 3, 4, . . .
K3
R(CL),NH, t CL & R(CL),+,NH, n =

where CL represents caprolactam and (CL) represents -NHCO(CH,),-.


In this case we take K = 1(3 aa a good approximation since the same
functional groups a r e involved in the equilibria. No satisfactory equi-
librium data exist in the literature for this case. Tobolsky [4]ven-
tured to predict that this equilibrium could be described by using
the results of Case Ia by merely replacing K by KS.The first im-
portant prediction is obtained from Eq. (131, namely K, l/[M],.
Hence, in the range of high degrees of polymerization, the equi-
librium concentration of monomer is a function of temperature only
and independent of the nature of the initiator.

Scheme 3.

It should be pointed out that in theory, a s outlined above, the


presence of cyclic oligomers is not taken into consideration. These
a r e present in the equilibrium mixture to an extent of at most 5%
[30],and since the size of the rings varies just as does the length
252 SAWADA

of the open chains, they a r e assumed not to change the value of P


which is normally obtained for the straight chains only. Also, con-
densation reactions of the type M,* + M,* * M$+, are assumed not
to change the number-average molecular weight of the polymer at
equilibrium since equilibrium depends only on the initial and final
state, not on the reaction path o r reaction mechanism.
The AH, for the condensation reaction
K2
-NH, + -COOH ---L -NHCO- + H,O
was reported to lie between -1400 and -600 cal/mole. Meggy [31]
found AH, = 3560 cal/mole for the propagation reaction. A calori-
metric study of the polymerization sets the total heat of reaction
Q = 28.5-29.0 cal/deg, and Fukumoto [32] found the heat of conden-
sation between linear polymers to be -6.8 kcal/mole. The temper-
ature dependence of K, of a nylon 6 is given by the equation [33]
log K, = 1,342.5/T + 0.2033
AH, = -6,140 cal/mole
AS, = 0.9 eu
When E-caprolactam polymerizes at 257'C, about 7% of the lac-
tam monomer remains unchanged on account of the ring-chain equi-
librium. If one of the hydrogen atoms is substituted by a methyl
group, the amount of the lactam monomer unchanged increases,
that is to say, the equilibrium i s shifted to the side of ring mole-
cules, If two of the hydrogens are substituted by two methyl groups,
the polymerization does not occur. These shifts in chain-ring equi-
libria a r e explained by the change in the thermodynamical properties
caused by the existence of rotational isomers [34]. The effect of the
alkyl group substitution on the amount of the lactam ring remaining
at the equilibrium is shown in Table 5.
3. Laurolactam. The equilibrium polymerization of laurolactam
was studied between 260 and 320°C using organic acids and amines
as initiators, and it was treated as a two-constant case by Elias and
Fritz [35]. The heat of polymerization was found to be AH, = -3.1
kcal/mole, the corresponding entropy AS, = -3.75 eu, the enthalpy
and entropy for the formation of N-lauroylamino lauric acid to be
AH = -3.35 kcal/mole, and AS = -4.5 eu. The ceiling temperature
is approximately 550°C.
4. 1,3-Dioxolane. Equilibrium polymerization of 1,3-dioxolane
has been studied with cationic catalysts in methylene dichloride [36]
TH E R MODY NAM ICS OF PO LYM E R I ZAT ION. I V 253

Table 5
Chain-Ring Equilibria in the Polymerization of
Substituted eCaprolactams [ 341

Equilibrium monomer concentration Iwt%)


Substituted ~

ECaprolactam Calculated 0 bserved

ECaprolactam - 7
3-Methyl- 16 13
4-Methyl- 23 27
5-Methyl- 23 28
6-Methyl- 23 28
7-Methyl- 16 16
N-Methyl- - 100
5-Ethyl- 50 67
5-Propyl- 60 100
3,6Dimethyl- 100 100

o r in benzene [37]. The living nature of the polymerization of 1,3-


dioxolane was confirmed by the following evidence: A second por-
tion of monomer added to a polymerization mixture in which reac-
tion had ceased gave a reaction rate of the same order of magnitude
a s the first reaction and the yield of polymer and its degree of
polymerization corresponded to what was obtained when a quantity
of monomer equal to the sum of the two portions was polymerized
in one step. The degree of polymerization of the polymer and the
position of equilibrium could be adjusted to any desired value by
adjusting the final temperature of the reaction mixture, irrespective
of the temperature at which the polymerization had been carried out.
This evidence shows that termination is unimportant, and that the
polymerization of 1,3-dioxolane is a true equilibrium polymeriza-
tion [36,37].
From the measured yields of polymer the equilibrium constant
K, for the monomer-polymer equilibrium o r the equilibrium mono-
mer concentration, which is the same, was calculated by Plesch
et al. [36] and Yamashita et al. [37]. From the variation of K, with
temperature the following values of AH,, and ASs, have been obtained:

AH,, = -5.1 i 0.2 kcal/mole AS,, = 18.6 f 1.2 eu by Plesch et al.


AH,, = -3.6 f 0.6 kcal/mole AS,, = -14 f 2 eu byyamashitaetal.
254 SAWADA

Direct calorimetric measurements in the adibatic reaction calorim-


eter gave AHam= -5.2 f 0.1 kcal/mole, and depolymerization ex-
periments gave for the enthalpy of depolymerization AH,, = +5.2 f
0.6 kcal/mole [36].
A plot of AG,," against temperature gave AS,," = -18.6 f 1.2 eu
and a ceiling temperature T," = 1.5"C for the solution in the standard
state 1 mole/liter. Since AH was found to vary neither with mono-
mer concentration nor with the degree of polymerization of the poly-
mer, it is possible to assume that the observed AH,, = AH,," and,
from this and AS,,", Tf = 1.O"C. For the 1.22 M solution T, is found
to be 7°C. In Fig. 5 is shown the variation of the degree of polym-
erization with temperature. It is observed from this figure that the
curve extrapolates to P = 1 at 8"C, the ceiling temperature for the
1.22 M solution, in good agreement with the calculated value. The
shape of the curve indicates, according to Dainton and Ivin [38],
that in this reaction there is transfer to monomer. The fact that
this curve appears not to become asymptotic to the temperature
axis, although the P's range from 70 down to 7, and the close agree-
ment between the experimental and calculated values of T,", both
indicate that AG,," cannot vary much with P [36]. The ceiling tem-
perature of the bulk polymerization of 1,3-dioxolane was found to be

TEMPERATURE ( O K )
Fig. 5. Dependence of degree of polymerization on temperature for the
equilibrium polymerization of 1,3-dioxolane 1361.
THERMODYNAMICS OF POLYMERIZATION. IV 255

about 150°C [37]. This is much higher than the ceiling temperature
for the polymerization of tetrahydrofuran.
5. 1,J-Dioxepan. The first thermodynamic information on the
polymerization of this monomer, which is an equilibrium reaction,
came from the semiempirical calculation of AHgg= -4.7 kcal/mole
by Skuratov and his co-workers [39]. A brief exploratory study of
the monomer-polymer equilibrium of this and some related com-
pounds was made by Strepikheev and Volokhina [40].
1,3-Dioxepan was polymerized under vacuum in methylene di-
chloride solution by anhydrous perchloric acid. The reaction in-
volves a perfectly clear monomer-polymer equilibrium. The
oligomers and polymers a r e cyclic and a r e formed by a ring-ex-
pansion mechanism (Scheme 4). This system was studied by Plesch
and Westermann [41] who obtained AH,," = -3.5 f 0.3 kcal/mole

Scheme 4.
from the van't Hoff plot of the results; this showed that there is no
significant variation of AH,," with temperature. From the plot of
AG,," against temperature, AS,," = - 11.7 f 1.5 eu, and AS,," was
also found to be invariant with temperature. From Fig. 6 the stan-
dard ceiling temperature T," at which AG,," = 0 for the 1 M solution
(strictly, the ideal 1 M solution) is found to be +27"C. The value
calculated from AH,," and AS,," is T, = +26"C. Under these condi-
tions departures from ideal behavior seem to be unimportant.
For comparison, polymerization of 1,3-dioxepan and 1,3,6-tri-
oxocane were examined in benzene solution with BF,,Et,O [371.
From the relation between equilibrium monomer concentration and
polymerization temperature the heats and entropies of these polym-
erizations were evaluated as follows:
1,3-Dioxepan AHs8 = -3.2 f 0.5 kcal/mole
AS,, = -9.3 * 1.4 eu
1,3,6-Trioxocane AH,, = -5.3 * 0.8 kcal/mole
AS,, = -9.3 f 1.4 eu
256 SAWADA

I
I
I
12 I I I I I
200 220 240 260 280 300

TEMPERATURE (OK)

Fig. 6. Graphical determination of standard ceiling temperature of 1,3-


dioxepan [4 11.

6. Trioxane. In the cationic polymerization of trioxane and tetra-


oxane near room temperature, the equilibrium trioxane concentra-
tion is not negligible during polymerization [42,43]. Tetraoxane was
polymerized with BF,Et,O in various solvents, and the equilibrium
concentration of trioxane produced by the polymerization of tetra-
oxane and equilibrated with the growing polyoxymethylene chain
was determined [44]. The equilibrium trioxane concentrations were
0.05, 0.13, and 0.19 mole/l in benzene, ethylene dichloride, and
nitrobenzene a t 30°C, respectively, and 0.20 mole/l in ethylene di-
chloride at 50°C. The values in ethylene dichloride showed that
the approximate values of AH3 and AS, were -4.2 kcal/mole and
-9.7 eu, respectively [44].
Equilibrium pressure of gaseous monomers a r e measured spec-
troscopically over the solid polymer at different temperatures. The
heat and entropy of polymerization of gaseous trioxane to the crystal
line polymer were calculated from the formula

RT In P = AH - TAS

Thermodynamic constants of polymerization of trioxane were cal-


culated for different phase conditions of monomers and polymers [45].
Results of calculation are indicated in Table 6 with those of tetra-
oxane .
Values of the standard enthalpy and standard entropy of polym-
erization of solid trioxane can be calculated using t h e data of Bus-
THERMODYNAMICS OF POLYMERIZATION. IV 257

Table 6
Thermodynamic Constants for Polymerization
of Trioxane and Tetraoxane

Standard -AH" (298°K) -AS" (298°K) -AGO (298°K)


Monomer states (kcallmole) (eu) (kcal/mole) Ref.

Trioxane gc 15.2' 37.0 4.2 45


9.9b 25.3 1.45 45
O.lb 0 0.1 45
Ic 5.4b 11.7 0.9 45
cc 2.0b 1.6 1.5 45
cc *
1.1 0.93b -4.3 3.8
_+ 2.38 46
sc 4.2a 9.7 1.3 43
11.5b 37.2 0.42 d
gc
cc 1.71b -7.23 0.45 d
Tetraoxane cc 0.79 f 0.04' *
-0.82 0.01 1.03 48

'Data from equilibrium studies.


bData from various kinds of thermal data from different sources.
'Data from measurement of the heat capacity of monomer and polymer and heat of
polymerization.
dT. P. Melia, D. Bailey, and A. Tyson, J. Appl. Chem., 17,15 (1967).

field and Merigold [46] and that obtained from formaldehyde/tri-


oxane equilibrium experiments [47]. Thus for the reaction

,-(CH,0)37 (c,crystalline) __t 3/n ---(CH201n---

AH,," = 1.1 * 0.93 kcal/mole


AS,," = -(4.3 f 3.8) eu

Recently the free energy of polymerization of crystalline tetra-


oxane h a s been accurately determined by Nakatsuka and others [48].
Their results are compared with the corresponding thermodynamic
functions for the polymerization of crystalline trioxane in Table 6.
For both monomers both AH,," and AS,," a r e close to zero and the
standard free energy of polymerization, AG,,", is negative at all
temperatures. Neither system therefore exhibits either a ceiling
or a floor temperature and polymerization is thermodynamically
possible at all temperatures. However, in the case of trioxane, the
258 SAWADA

precision is such that the sign of both AH,," (298°K) and AS,," (298°K)
is a little uncertain. The effect of temperature on these quantities
also makes extrapolation to other temperature ranges uncertain
without accurate heat capacity data.
7. Sulfur. The first relatively complete theory of the equilibrium
polymerization of liquid sulfur was given by Gee [49,50]. This theory
was characterized by the use of two distinct treatments: one valid
below the transition temperature and one valid above the transition
temperature. Tobolsky and Eisenberg [3] later presented a unitary
theory valid above and below the transition temperature. This theory
was perfectly consistent with the results of Gee.
The conditions at equilibrium at any temperature for the polym-
erization of sulfur may be derived as

Let N and W represent the total concentration of polymer molecules


and the total concentration of monomer segments (S, units) incorpo
rated in the polymer, respectively.
THERMODYNAMICS OF POLYMERIZATION. IV 259

It is obvious that W/N = P, where P is the number-average chain


length (in terms of S, units)

P = W/N = 1/(1 - %[MI,) (10)

Thus

Equation (10) may be substituted into Eq. (29) and

P-1
[%I = PK, + Z P ( P - 1) (30)

For the temperature region in which high polymer is stable, i.e.,


above the floor temperature, P is >> 1, from Eq. (30) we see that

p 2 (
[%IK,--1
K )
It must be emphasized that in deriving Eq. (30), no assumptions
were made that restrict the validity of the formula to any tempera-
ture region; it should be applicable in the entire liquid range.
Only a knowledge of K, and & for any temperatures is required
for the determination of P and [MI,at that temperature and, con-
versely, if P and [MI,are known, K and K, can be determined. From
Eqs. (10) and (29), one can calculate the value of K and I& at a given
temperature from the experimental values of [MI, and P. This was
done for two temperatures above 159"C, and by use of the van't Hoff
equation the following relations were obtained for the equilibrium
polymerization of sulfur:

In K = -AH"/RT + AS"/R; AH" = 32.8 kcal/mole, AS" = 23.0 eu

In K,= -AH,"/RT + AS,"/R; AH," = 3.17 kcal/mole, AS," = 4.63 eu

The linearity of the plots of In K vs. 1/T and In K3vs. 1/T is strik-
ing evidence for the validity of the theory, At this point K and K,
were known at all temperatures and P and [MI, could also be eval-
uated at all temperatures. Remarkably this led to a complete pre-
diction of the P vs. T curve and the [MI, vs. T curve, as shown in
Figs. 1 and 7, including the prediction of the sharp transition. The
260 SAWADA

1 I I I
450 500 55 0

TEMPERATURE (OK)
Fig. 7. Equilibrium monomer concentration a s a function of temperature
for the equilibrium polymerization of sulfur [3l. (0):Experimental points.
(-) : Calculated values.

entropy change on polymerization is positive, as the larger rings


had more favorable entropies of ring opening than did the small
rings. The standard entropy change involved in converting cyclo-
hexane to hexene-1 is 25 eu, which agrees well with the value of
AS" = 23 eu which was obtained for the opening of the S, ring to the
S, diradical. The enthalpy change is also positive in sulfur polym-
erization because of the greater stabilization of the S-S bonds in
the ring, due to resonance and bond interactions, than in the open-
ing- chain polymer.
The Tobolsky-Eisenberg theory was based on the chemical reac
tions

K
S, + S,*, n-1) 3
7' _8 , * LS8n*1 = K3[S81[S$(n-1)l (33)

where the asterisk indicates chain molecules and the only nonpoly-
mer molecule is supposed to be the 8-membered ring. With the
help of these equations and a straightforward application of the
theory of chemical equilibria, the weight concentration of polymers
as well as the mean chain lengths were expressed in terms of K,
and I(3. It is obvious that Eq. (33) can be replaced by Eq. (34) with-
THERMODYNAMICS OF POLYMERIZATION. I V 26 1

out any change whatever in the theory o r its results:

s8p* + s8q*- s,*(p+q) cs,*(DW)l = K5[SED*I[SE,*l (34)


(p,q=1,2, 3,4,. . .)
The relation between K,, I(3, and K, is

K, = K,K, (35)

Though this theory gave an essentially correct picture of the polym-


erization of sulfur, its basis as given by Eqs. (32) and (33) o r (34)
contains the assumption that the number of atoms in all the mole-
cules can only be multiples of eight.
It is possible to write alternative equations where this assump-
tion is avoided [51]. Let u s consider the following relations

Sn* + Sm*- Sfn+m) (n , m = 1, 2, 3, 4 , . . .)

s, - s,*
Equation (38) accounts for the possibility of other than the 8-mem-
bered ring.
In order to compare the consequences of Eqs. (36)-(38) with
Eqs. (32)-(35), the following definition f o r & is introduced, in
analogy with Eq. (35):

Equation (36) still contains a simplifying assumption, namely that


Z, is the same for all values of n and m. While this is true for large
values of n and m, it is doubtful for values of n and m = 1, 2, etc.
The equations resulting from these equilibria are identical with
equations resulting from the theory based upon Eqs. (32) and (33)
after substitution:

K, = 8E1, K, = l/SE,, I(3 = & (40)

The equilibrium polymerization of selenium follows a scheme


262 SAWADA

mathematically identical to that of sulfur. The results for the vari-


ous thermodynamic quantities for selenium a r e [51]
AH" = 25 kcal/mole, ASo = 23.0 eu
AH,' = 2.27 kcal/mole, AS," = 5.47 eu
Although AH and AS are positive, the volume change on polymeriza-
tion (- AV = 6.15 ml/mole) is negative, as in a vinyl polymeriza-
tion. The floor temperature for the appearance of polymer in the
liquid state is calculated as 83°C.
C. Polymerization of Aldehydes
Anionic polymerization of chloral at low temperature was found
to be living, and an equilibrium between polymerization and depolym-
erization was reached easily provided that the degree of polymeri-
zation of polychloral was not too high. True equilibrium monomer
concentration and polymer yields were obtained only when an end-
capping reaction was carried out at the polymerization temperature
because the nonend-capped living polymer was depolymerized im-
mediately at room temperature. When the polymer chain grows
longer, polymerization is stopped not by termination but by occlu-
sion of the active end. Consequently a true equilibrium did not hold.
From the temperature dependence of the equilibrium monomer
concentration, the change in enthalpy and entropy as well as the
ceiling temperature for the polymerization of chloral in tetra-
hydrofuran were determined as -3.5 kcal/mole and -12 eu and
1l0C, respectively [52]. On the other hand, Busfield and Whalley [53]
have obtained for monomer-polymer equilibrium of chloral in pyri-
dine -8.0 kcal/mole, -28 eu, and 125°C for the change in enthalpy,
entropy, and ceiling temperature, respectively.
Chloro- and/or methyl-trisubstituted acetaldehydes and iso-
butyraldehyde were polymerized anionically in tetrahydrofuran at
low temperatures, All the polymers were living and an equilibrium
between polymerization and depolymerization was reached easily
provided that high catalyst concentrations were used. From the
temperature dependence of the equilibrium monomer concentration,
the heats and entropies of polymerization and the ceiling tempera-
tures were determined. The ceiling temperatures of the polymeri-
zation of trisubstituted acetaldehydes decreased in the order
CC1,CHO > CCl,(CH,)CHO > CCl(CH,),CHO > C(CH,),CHO
TH E R MO DY NAM ICS OF PO LY M E R I ZAT I ON. I V 263

and the heats of polymerization decreased in the reverse order,


showing that chlorine substitution for methyl groups in aldehydes
increases the stability of a polymer relative to its monomer and
that the change in entropy has a stronger influence on the order of
the ceiling temperatures of the series than the change in enthalpy [ 541.
Formaldehyde is one of the few aldehydes where an equilibrium
with the addition polymer can be established fairly readily without
the deliberate addition of catalysts. Therefore, equilibrium pres-
sures of gaseous formaldehyde over solid polyoxymethylene, p,
have been measured by several workers [55,56]. The heat and
entropy of polymerization were calculated from

R T l n p = AH - TAS
These results are compared in Table 7 with some of the more re-
cent published data [57].

Table 7
Thermodynamic Constants for Polymerization of
Formaldehyde from Equilibrium Studies

-AH
Standard states (kcaVmole)

15.9 40.7 46
12.2 30.6 55
16.3 41.8 56
14.3 - d
17.0 - e
13.1a - 57
.-
41Bb f
16.7 42.6 45
ll.lC 20.7' 45
14.gC 38.7c 45

"Data from combustion method.


bData from third law.
'Data from various kinds of thermal data from different sources.
dJ. F. Walker, Formaldehyde, 3rd ed., Reinhold, New York, 1964.
eJ. B. Thompson and W. M. D. Bryant, Polym. Preprints, 11(1), 204 (1970).
fF. S. Dainton, D. M. Evans, F. E. Hoare, and T. P. Melia,Polymer, 3,263 (1962).
264 SAWADA

111. TRANSITION PHENOMENA IN


EQUILIBRIUM POLYMERIZATION
In certain cases there exists a sharply defined “ceiling tempera-
ture’’ above which high polymer is thermodynaniically unstable with
respect to monomer; in other cases there exists a sharply defined
“floor temperature” below which polymer is unstable with respect
to monomer.
For temperatures above t h e ceiling temperature, since polymer
is thermodynamically unstable with respect to the monomer, the
equilibrium monomer concentration equals approximately the
original monomer concentration, i.e.,

[&I [MI, (41)


and therefore
P 1/(1 - KJM,,]) (4 2)

At the ceiling temperature, Tobolsky [58] obtained the approxima-


tions
P, (K,[%])-’” (43)

([&I - [MI,), = ([%I/K3)1’2 (44)

the subscript t denotes the transition (floor o r ceiling) temperature.


The steepness of the transition can be indicated by the value
dP/dT at the ceiling temperature

(3 ~
AH,”
, = RT2{2K3[%I - (K3[%13”)I - KJX,I{2
AH,O/RT*
- (K3[%I)1’2}
(45)
We add another equation which defines the sharpness of the transi-
tion as a function of [&]in the range of the critical [M,,] concentra-
tion at constant temperature

(dP/d[M,,]), [%1/2 (4 6)

If there is a sharp transition, [MI, should be close to [M,,] and P


should be large at the transition temperature. F r o m Eq. (42) we
see that the transition temperature is defined by
K,[&] = 1
THERMODYNAMICS OF POLYMERIZATION. I V 265

This will be a ceiling temperature if AH,' i s negative, and a floor


temperature if AH," is positive. And thus

[M,,] exp(AS,"/R) exp(-AH,'/RT,) = 1


or
I~[M,,] = -AS,O/R + AH,O/RT,
Therefore

Thermodynamically, Dainton and Ivin [38] conceived the ceiling


temperature to be that temperature at which the free energy of
polymerization (for long-chain polymers) passes from a negative
to a positive value a s the temperature i s raised, i.e.,

T, = AHJAS,

o r , defining AS," as the entropy change for [MI = 1 mole/l of mono-


mer,
A H,
T, = (49)
AS," + R ln[M],

which i s equivalent to Eq. (47) of Tobolsky.


It i s interesting to compare Dainton's definition of ceiling tem-
perature with that of Tobolsky. Both have the same mathematical

free monomer -
form. Dainton's definition refers to the single chemical change

monomer segment of a high polymer

and takes no account of the building process leading to a macro-


molecule. On the other hand, Tobolsky refers to the whole stepwise
building reaction and gives full consideration to the initiation pro-
cesses.
The factors affecting the sharpness of the transition should be
examined. The expression giving [dP/dT], shows clearly that the
transition is sharper the greater AH,' (the heat of polymerization)
and the smaller X,, (the number of initiating species). Increase in
AH value narrows the temperature range in which K varies from
values much larger than 1/[M,,] to those much smaller. For lower
X,,, the amount of polymerized monomer is distributed among fewer
266 SAWADA

chains, giving a greater increase in P. Similarly, the degree of


polymerization, P, as a function of [q] for constant temperature
and constant [X,,]may be calculated, The sharpness of the transi-
tion is defined by ( d P / d [ q ] ) in the vicinity of the critical concentra-
tion [M,,] and is given approximately by equation (dP/d[M,,],) = $[%I.
It should be pointed out, however, that if only one equilibrium
constant is involved, o r stated alternatively, if the constants for
initiation and propagation were identical as Dainton [38] assumed,
and as is presented in P a r t I of this review [59], the transition would
be s o diffuse that any indication of the sharpness of the transition
would not be recognizable [60]. All of the kinetic studies of Dainton's
group [38] showed that the rate of polymerization does not drop
abruptly to zero at the ceiling temperature, but that in its vicinity
the curve, giving rate as a function of temperature, becomes asym-
ptotic to the T axis. This again points to a broadening of the transi-
tion arising from a decrease in the molecular weight of the result-
ing polymer.
Equations (42) to (45) were applied numerically to a-methylsty-
rene. Using the data of Worsfold and Bywater [14] and also of
McCormick [13], the following result for I(3 (the concentration units
a r e still moles/kg and the standard state is 1 mole/kg):

K, = exp(AS,'/R) exp(-AH,"/RT)
AH,' = -7.2 kcal/mole
AS," = -27.6 eu

For an initiator concentration of 0.001 mole/kg and for two values


of [M,,], namely, 1.0 and 2.5 moles/kg, Tobolsky and Eisenberg [58]
have computed P vs. T according to Eqs. (42) and (43) as shown in
Fig. 8. In these same conditions the total weight concentration of
polymer ([M,,] - [MI,) will appear as shown on Fig. 9 as a function
of temperature. In Fig. 10 P is plotted vs. [M,,] at two different tem-
peratures, +5.6 and -7.6"C., for the same initiator concentration.
The numerical values for the transitions are:

For [M,,] = 1 (Figs. 8 and 9)


T, = 5.6"C
P, = 31.6 (Point A)
([M,,] - [MI,),= 0.0306 mole/kg (Point B)
(dP/dT), = -24.7 units/degree
(d([M,,] - [M],)/dT), = -0.0247 mole/kg/degree
THE R MODY NAM I CS OF POLYME R I ZAT I ON. I V 26 7

TEMPERATURE PC)
Fig. 8. Dependence of degree of polymerization on temperature for the
polymerization of a-methylstyrene for Mo = 1 and 2.5 moles/kg. Xo = 0.001
mole/kg [58].

For [q]= 2.5 (Figs. 8 and 9)


T, = 25.6"C
P, = 50 (Point C)
-
([%I [MI,), = 0.0490 mole/kg (Point D)
(dP/dT), = -53.1 units/degree
(d([M,,] -
[M],)/dT), = -0.0531 mole/kg/degree

For K, = 1 (T = 5.6"C, Fig. 10)


P, = 31.6 (Point A)
([M,] - [MI,), = 0.0306 mole/kg
(dP/d[w]), = +500 units/degree

For K, = 2 (T = -'7.6"C, Fig. 10)


P, = 22.4 (Point E)
([q] - [MI,), = 0.0214 mole/kg
(dP/d[w]), = 500 units/mole/kg
208 SAWADA

TEMPERATURE ('C)
Flg. 0, ( Mo- Me) VB. temperature for the polymerlzatlon of cu-methylety-
rene for Mo = 1 and 2.6 molee/kg. Xo = 0.001 mole/kg [68].

For bulk sulfur [w]= 3.90, and this transition occurs at 159"C,
a temperature at which the liquid suddenly seems to acquire a very
high viscosity. The AH,' is positive in this case, so we have a
"floor temperature." At this transition point, the following rela-
tionships hold:
T, = 432'K = 159°C
P, = 1 / ~ 1 / 3= 1.2 x 10' units (50)

([%I - [MI,),= K1/,/KS = 3.24 X lo4 mole/kg (51)


- = 1.35 x loBunits/degree (52)
3RTa

From Eqs. (50) and (61) it is clear that at the transition tempera-
ture the relative amount of polymer is very small, but the degree of
THE R MODY NAM ICS OF PO LYME R I ZAT ION I V
I -
209

Flg, 10. Dependence of degree of polymerleatlon on the lnltlal monomer


concentratlon for the polymerlaatlon of rY-methylatyrene for T = 8.6 and
-7.B"C. Xo = 0.001 mole/kg [68].

polymerization is very high. Furthermore, at the transition tem-


perature, the rate of increase of degree of polymerization with
temperature is very great, as shown by Eq. (52).
In summary, for temperatures below the floor temperature, the
following approximations can be used:

For the temperature region in which high polymer it3 stable, Le.,
above the floor temperature, P is >> 1, and the following approxima-
tions can be used:
270 SAWADA

IV. MOLECULAR WEIGHT DISTRIBUTION


A. Equilibrium Polymerization
Polymeric materials display a distribution in molecular sizes.
For example, Flory [61] developed the theory for size distribution
in condensation polymers by a statistical treatment. From a con-
sideration of the kinetics of condensation polymerization, it has
been shown that the molecular size distribution in linear condensa-
tion polymers containing equal numbers of the two cooperating func-
tional groups (e.g., OH and COOH) is given by

N, = NpX-'(l - p) (54)
where N, is the number of molecules composed of x monomer units,
N is the total number of molecules, and p is the extent of reaction,
i.e., the fraction of the functional groups which has condensed.
The distribution of Eq. (54) for the equilibrated linear polymer
can be derived [62] in a manner analogous to the derivation of the
Maxwell-Boltzmann energy distribution law:

2~~= N
x =1
(55)

2 x N x = n,,
x =1

where n,, is the total number of units and


N= - P) (57)
Both n,, and N are constant under the conditions, and Eqs. (55) and
(56) are analogous, respectively, to the conditions of conservation
of matter and energy in the Maxwell-Boltzmann derivation. A
macrostate is defined merely by the numbers of molecules of the
various size, i.e., by N,, N,, N,, etc. For a given macrostate there
are
w = N!/TN~!
microstates. Solution of Eq. (58) for the maximum value of W con-
sistent with Eqs. (55) and (56) by the usual variation method yields
Eq. (54) for the most probable macrostate. Thus, under the assump-
tion that the thermodynamic stability of a given interunit bond is
independent of the size of the molecule and its position along the
chain, the equilibrium distribution is ultimately attained by random
synthesis. The difference in entropy between a mole of a hetero-
THERMODYNAMICS OF POLYMERIZATION. I V 27 1

geneous polymer and a mole of the single species of molecular


weight equal to the number-average molecular weight for the hetero-
geneous polymer is given by the entropy of mixing expression:

= -RC(N,/N) ln(N,/N) (59)


which may be called the molar entropy of heterogeneity. For the
most probable distribution, substitution of Eq. (54) in Eq. (59) gives

= R[ln(p/l - p) - (In p)/(l - p)1 (60)


which also can be obtained directly from the Boltzmann relation
S = k In W, where W is taken to be W, in Eq. (58). For a high-
molecular-weight polymer, p is near unity and
AS, = ~ [-iIn(1 - p)]
or
AS, = R[1 + ln(P,)]
since the number-average degree of polymerization is given by
P, = 1/(1 - P) (63)
For the entropy of heterogeneity per mole of structural unit, we have
ASh’ = AS,/P, (64)
These equations express the maximum entropy of heterogeneity,
o r entropy of mixing, for a given degree of polymerization. Any
distribution other than Eq. (54) will yield a lower entropy of hetero-
geneity. The entropy per mole of polymer molecules AS,, increases
without limit as p approaches unity and P, approaches infinity; the
entropy per mole of structural units mh’, after reaching a maximum
at a very low degree of polymerization, decreases asymptotically
toward zero as P, increases.
At any given temperature there should be an equilibrium dis-
tribution i n sizes, corresponding to a condition where entropy and
heat are balanced so as to give a minimum free energy, The
logarithm of the number of configurations SZ available to a mixture
of n, solvent molecules and nI molecules of monomer, n, mole-
cules of dimer, n, molecules of x-mer, etc., is [63]
272 SAWADA

where v is the coordination number of lattice and u is the symmetry


number of the long-chain molecule (usually 2).
If equilibrium properties of polymer sizes is known, Eq. (65)
gives the number of available configurations. The energy of each
distribution with specified values of the n,'s can easily be written
in terms of the free energy change of polymerization per molecule,
f:
F = -xn,(x - 1)f (66)

The equilibrium properties of a system containing n, solvent mole-


cules and N monomeric units (which may be linked together in any
fashion into polymeric units) can be obtained in theory by evaluation
of the partition functions:

where the subscript i refers to a particular set of the n,'s corre-


sponding to a definite distribution in sizes among the polymer mole-
cules.
The macrostate corresponding to 'equilibrium" is the state for
which the free energy is a minimum and can be found by maximizing
the expression

In Sat - FI/kT (68)

with respect to the variables n,, n,, .., , n,, . ,, subject to the condi-
tions that
xxn, = N (69)
Xm1

The problem of maximizing Eq. (68) subject to Eq. (69)can be best


solved by the method of Lagrange multipliers. In other words, mul-
tiply Eq. (69) by -a/kT and add to Eq. (68) to give, after substitution,
Eq. (65). Hence, it is easy to show that the resulting distribution [64]
will be identical with the one described by Eq. (54). The same dis-
tribution results when random scission occurs to infinitely long
chains [65]. Flory [66] has derived the above molecular size distri-
bution equation by a similar procedure based on the same equations,
To describe the properties of a polymerizing system at equilib-
rium, the change in free energy of polymerization must be estimated,
Polymerization from monomer to the equilibrium polymer consisting
of a mixture (solution) of monomer, dimer, trimer, n-mer, etc., may
be considered to be the sum of two processes, the polymerization of
TH ER MODY NAM ICS OF PO LY ME R I ZAT ION. IV 273

monomer to pure homopolymers and the mixing of the pure homo-


polymers in their polymerization system, The free energies of
formation, and hence, the free energies of polymerization, of pure
gaseous monomer to form pure gaseous n-mer are expected to be
proportional to molecular size, i.e., chain length, for an homologous
series such as the polymerization of ethylene gas to the gaseous
1-alkanes [67]. This is a consequence of assuming that vibrational,
rotational, and translational motions of a molecule are independent
of one another and that partition functions describing these motions
may be written as the product of partition functions, each describ-
ing one type of motion [68].
The free energies of formation, and polymerization of gaseous
o r liquid monomer to form pure liquid n-mer homopolymer, may o r
may not be linear functions of molecular size depending upon the
degree of order present in pure liquids, The difference in entropy
between a rigid (crystalline) n-mer molecule and a flexible n-mer
molecule is approximately [69]

where k is Boltzmann’s constant, n is the degree of polymerization,


and z is the coordination number of a lattice approximating the
structure of the liquid polymer. Equation (70) is derived by count-
ing the number of ways that n links of a flexible n-mer molecule
can be arranged in n subvolumes. A consequence of Eq. (70) is the
expectation that free energies of formation and polymerization of
monomer to pure liquid n-mer homopolymer may be expected to be
linear functions of chain length n if the liquid is highly random,
approximating the disorder of pure, dense gaseous n-mer, and to
depart from linearity by approximately kT ln(n) per molecule if the
liquid is highly ordered, approaching the regularity of a crystal.
That real liquid homopolymers a r e somewhere between these two
extremes seems likely.
The two limiting cases for polymerization equilibri.um arise as
follows:
(1) Assume the free energies of formation and polymerization
of monomer to pure, crystalline (perfectly oriented) n-mer homo-
polymers are linear functions of molecular size. Mix the monomer,
dimer, trimer, n-mer, etc., to form the equilibrium polymer. This
gives a “most probable” distribution at polymerization equilibrium
[661.
274 SAWADA

(2) Assume the free energies of formation and polymerization of


monomer to pure, unoriented (random oriented) n-mer homopoly-
mers are linear functions of molecular size. Mix the monomer,
dimer, trimer, n-mer, etc., to form the equilibrium polymer. This
gives a very broad molecular weight distribution at polymerization
equilibrium [TO].

The free energies of formation and polymerization to form liquid


alkanes (or 1-alkanes) may be expected to be linear functions of
chain length. If normal alkanes are disoriented liquids, Case 2 lead-
ing to a very broad molecular weight distribution at equilibrium is
the more likely. If normal alkanes are highly ordered liquids, Case 1
leading to a “most probable” molecular weight distribution at equi-
librium may be correct in spite of the expectation that entropy of
polymerization of monomer to pure, oriented polymer might be non-
linear with chain length by the logarithm of chain length according
to the estimate of disorder of a random polymer chain (Eq. 70).
Consider the polymerization of n-gaseous monomer units at unit
fugacity to form liquid n-mer in the equilibrium polymerization mix-
ture under a total pressure P:

n-monomer (g,f = 1) n-mer (L,P)


The activity product for this reaction may be written as

where f , = fugacity of monomer, a, = activity of n-mer, Nu = mole


fraction of n-mer, y(n) = activity coefficient of n-mer, and AF,O(n) =
standard change in free energy of polymerization of gaseous mono-
mer at unit fugacity to pure, unoriented, liquid, n-mer homopolymer
at pressure P.
A bulk polymerization is a solution (mixture) of monomer, dimer,
trimer, n-mer, etc. In such a system in the liquid phase, the activity
coefficient y(n) is given by

In y(n) = In(n/P,) + 1 - (n/p,)

where P, is the number-average degree of polymerization.


The entropies of homologous series of liquid hydrocarbons are
proportional to molecular size [71]. The heats of combustion (and
formation) of real polymers are proportional to molecular size [72].
Thus the standard changes in free energies of polymerization to
pure liquid homopolymers may be expected to be given by
THERMODYNAMICS OF POLYMERIZATION. IV 275

AF,”(n) = (Y + pn (73)

where cy and /3 a r e constants (dependent upon pressure and tempera-


ture) and n is the number of monomer units. For the mole fraction
of n-mer

N, = exp {n In f, c y + m- l n -n- l +
- ___
RT P,
(74)

For the number-average degree of polymerization

P, - 1
P, = - (75)
In P,

For the weight-average degree of polymerization

P, = 1 + Pi In P,

The molecular weight distribution of an equilibrium polymer as


given by Eq. (74) is very broad. The difference between the treat-
ments that lead to the “most probable” and the very broad molecular
weight distributions at polymerization equilibrium arises in whether
the entropies (and, hence, the free energies of formation and polym-
erization) of an homologous series of molecules are proportional to
molecular size in the liquid state o r in the solid state. For relatively
disordered liquids, such as polymerizing, never-frozen monoolefins,
the liquid is perhaps best described as a dense gas a s far a s con-
figurational entropy is concerned. In this case a broad molecular
weight distribution at equilibrium might be expected. On the other
hand, in a polymerizing system involving ionic species in a melt,
a high degree of order is present and the system is probably best
described a s a crystalline system. In this case, a narrow molecu-
lar weight distribution, perhaps approaching a “most probable”
distribution, might be expected.
For Case 2, assuming the entropies of pure, unoriented homo-
polymers a r e linear functions of degree of polymerization, the free
energies of polymerization may be estimated empirically using ideal
gas state free energy data and measured vapor pressures and molar
volumes for the 1-alkenes. If observed molecular weight, weight-
average/number-average ratios, and molecular weight distributions
270 SAWADA

for ordinary polyethylenes a r e compared with these properties cal-


culated for equilibrium polyethylene using published thermodynamic
data for the 1-alkenes, reasonable agreement is found. This indi-
cates that polymerizations of ethylene, particularly by the “high
pressure” process, may be controlled by energetics, at least in
part, rather than by kinetics alone, A broad molecular weight dis-
tribution is predicted for an equilibrium polymer. Thus, if ethylene
polymerizations are energetics-controlled, the observed broad
molecular weight distributions of polyethylenes are not anomalous
but are to be expected for vinyl polymerizations in bulk if the reac-
tions are carried out in such a way as to approach equilibrium.
In conclusion, the Lundberg [TO] distribution mentioned above is
a result of extreme deviations from ideality, Thus the Schulz-Flory
distribution can be derived not only from the ideal solution theory at
infinite dilution, but also from the Flory-Huggins theory at finite
concentrations. In fact Harris [73] points out that the Schulz-Flory
distribution results when deviations from ideality are expressed in
terms of the second virial coefficient, provided that the latter is in-
dependent of molecular weight, However, the distribution law is not
independent of the thermodynamics of the system and, indeed, may
vary significantly at higher concentrations from the idealized ex-
pression.
Eisenberg and Tobolsky [58] found that the distribution in equi-
librium polymerization, even in the presence of an initiator, was
the random distribution. For example, the expression for the mole
fraction of x-mers obtained by dividing the expression for the ab-
solute concentration of x-mers by the total concentration of polymer
is
N,/N = n, = (&[M],)a-l(l - K,[Ml,) (77)

which is the Schulz-Flory equation with p = &[MI,. Furthermore,


-
Pa = 1/(1 &[MI,) (10)
P, = (1 + &[MI,)/(1 - KJMIJ (78)
and the heterogeneity index is
P,/P, = 1 + %[MI, = 2 - 1/P, 2 (79)

B. Living Polymerization
It was shown [74] that the polymerization of vinyl monomers may
be carried out under conditions which exclude the termination step.
The resulting polymers a r e referred to as “living” polymers, since
THERMODYNAMICS OF POLYMERIZATION. IV 277

their ends retain the activity required for further propagation of the
polymerization process. Since the molecular weight distribution of
the product yielded by such a polymerization is unusual, this dis-
tribution will be examined in some detail. Let us consider now the
molecular weight distribution of a living polymer produced in a sys-
tem in which all termination processes are excluded. The polymer-
ization which interests us is initiated by some anionic species de-
noted by MI*. The propagation of the polymerization involves the
reactions
MI* + M - Ma*

-
M,*+M ___c M,*
...
MI* + M M:+1

Let us assume the rate constants of these reactions to be inde-


pendent of j and equal to k,*. The principle of microscopic revers-
ibility demands the occurrence of the reserve reactions, namely,
decomposition of an active j-mer into a monomer and an active
(j - 1)-mer, and the unimolecular rate constant of this depropaga-
tion, assumed again to be independent of j, is denoted by k,.
Since there is no termination, the system described above must
eventually reach an equilibrium state, the equilibrium concentration
[MI, of the monomer being determined by the equation

where [MI,, etc, denote concentrations. If the initial concentration


of all the active species is denoted by C, then

g[M,*], = C
I
and g[M,*], = C
2
- [MI*], (81)
where [MI*Ie denotes the equilibrium concentration of the active
species [MI*]. For a polymerization proceeding to a high molecular
weight product, [MI*], is negligible compared with C, and we obtain

This result is obtained from Eq. (80) by putting d[M]/dt = 0. The


equilibrium concentration of the monomer is independent of the ini-
278 SAWADA

tial concentrations of the monomer and of the catalyst (i.e., of the


active species), and is determined uniquely by the temperature and
by the nature of the monomer, of the polymer, and of the solvent.
The equilibrium molecular weight distribution of ‘living” poly-
mers may be calculated from the following set of equations:

The subscript e denotes that all the relevant concentrations have the
equilibrium values. It is essential to realize that in solving the above
set of equations, one must not apply the approximate Eq. (82), but
must use the exact expression

C-’d[M]/dt = -k,[M], + kd(1 - [M,*],/C) = 0 (84)


which leads to
[MJ*Ie = [MI *],(I - [Mi *Ie/C)’-’ (85)

Thus the equilibrium molecular weight distribution of “living” poly-


mers is given by the usual function named by Flory “the most prob-
able distribution” “751, and leads to P, = M, /C and P, /P, = 2.
(M,-the total amount of polymerized monomer-is given by the
equation M, = [&] - [MI,.)
Although the equilibrium molecular weight distribution of “living”
polymers is the same as that characterizing polymers formed by
many conventional processes, an entirely different molecular weight
distribution may be formed immediately after completion of the
polymerization process. Whenever a high-molecular-weight poly-
mer is produced, the propagation rate constant, k,, must be much
greater than kd, the depropagation constant, and the half-life time
of the depropagation is long compared with the time of the experi-
ment. Furthermore, the initial concentration of monomer, [&I,
must be substantially larger than its equilibrium concentration [MI,
and [&]/C >> 1. Polymerization in such a system usually leads to a
Poisson molecular weight distribution when nearly all the polymeric
molecules have a degree of polymerization close to [&]/C. Although
such a system is not yet in its true equilibrium state, the polymeriza-
THERMODYNAMICS OF POLYMERIZATION. IV 279

tion is essentially completed, and the concentration of the monomer


differs only insignificantly from that attained at the ultimate equi-
librium. On the other hand, the higher averages, e.g., P,-the
weight-average degree of polymerization-vary in time. A s was
pointed out by Flory [76], when initiation is a t least as fast as prop-
agation, a Poisson distribution of polymer chain lengths is obtained

N, = N exp(- v)F1/(x - l)!


where v = P, - 1.
The entropy of heterogeneity for this distribution, obtained by
substituting Eq. (86) in Eq. (59), is given to a close approximation
when the degree of polymerization is large by

R
Ash = y{1 + ln(2nPJ)

The distribution of Eq. (86) covers a much narrower range than the
equilibrium distribution. Correspondingly, its entropy of hetero-
geneity is less than for the equilibrium distribution a s given by
Eq. (62). Therefore, if interchange between polymer molecules
occurs, the distribution of Eq. (86) will be broadened toward the
equilibrium distribution of Eq. (54).
Miyake and Stockmayer [77] have investigated the mathematical
problem of the reversible living polymer system without transfer
and termination reactions, concentrating particularly on the molec-
ular weight distribution. This problem is identical to that for the
kinetics of BET adsorption, and a complete analytical solution is
obtained when the monomer concentration is kept constant. The
time required for the initial narrow Poisson distribution to go over
into the final “most probable” equilibrium o r Schulz distribution is
proportional to the square of the average chain length. For polysty-
rene at room temperature, the Poisson polymer is produced in a
few seconds, but the final equilibration requires of the order of 100
years, although the unreacted monomer concentration reaches its
equilibrium value very early. Thus the equilibrium constant for
polymerization can be accurately measured long before complete
equilibrium is attained in the system.
For purposes of illustration, let us discuss the polymerization
in essentially three stages as has been indicated by Miyake and
Stockmayer “771. The material formed in the first stage has a
Poissonian character and for all practical purposes it may be
taken as monodisperse if the average chain length is large. In the
280 SAWADA

second stage the statistical character does not alter appreciably


while the free monomer concentration acquires its equilibrium
value, In the third stage there is redistribution of monomers over
the chains. The number-average chain length stays constant while
the weight-average chain length changes to the most probable value.
Under the assumption k,[&] >> kd, the first two stages are com-
pleted in a time which is negligible compared with the time involved
in the third stage. Thus, to investigate changes in the statistical
character of the polymer due to depropagation, we may effectively
assume that at t = 0 the polymer sample has a Poisson distribution
of chain lengths with a number-average chain length equal to [w]/C,
Nanda and Jain [?a] have obtained an expression for size distribution
which is applicable to the earlier stages of widening of the distribu-
tion,

V, THERMODYNAMICS OF EQUILIBRIUM POLYMERIZATION


The effects of solvents on the equilibrium constants were inter-
preted on the basis of the thermodynamic theory of solutions in Sec-
tion I of Ref. 60. Let us now examine the effect of polymer concen-
tration on the thermodynamic activity of the monomer from a ther-
modynamic point of view, The yield and the degree of polymeriza-
tion of equilibrium polymerization are independent of initiator con-
centration, the polymer content of the system o r the solvent, and
depend only on the initial monomer concentration and the tempera-
ture. This ie strictly true only if the polymer is insoluble in the
reaction medium and the activity of the monomer is equal to its
concentration, Ivin and Leonard "791 have considered this problem
in detail, and have applied it to the case of the equilibrium anionic
polymerization of a-methylstyrene in tetrahydrofuran where, at
certain temperatures, the equilibrium monomer concentration de-
creased linearly with increasing concentration of polymer, The
polymer is soluble in tetrahydrofuran and its concentration there-
fore influences the thermodynamic activity of the monomer. This
effect waa first noted experimentally by Vrancken, Smid, and
Szwarc [80] who found a fall in the equilibrium concentration from
O,? M to about 0.46 M at 0" as the polymer concentration increased
from 0 to 2.8 base-mole/l.
The free energy change for the conversion of 1 mole of liquid
monomer to 1 base-mole of amorphous polymer, AQlo, will be given
by
AQ,, 9 AU,,, - AUn (88)
TH E R MODY NAM ICS OF PO LY ME R I ZAT I ON. I V 28 1

where A& is the partial molar free energy of the monomer (per
mole) in the equilibrium mixture relative to that of the pure liquid
monomer, and ha, is the partial molar free energy of the polymer
(per base-mole) in the equilibrium mixture relative to that of the
amorphous polymer. Expressions for A&, and AGD in a three-com-
ponent system (polymer-monomer-solvent)are quoted by Flory [el]
as

1
AGJRT = ;{ln +p + (1 - + J - + m ( X J J x J - +,(XJX,)
(go)
+ (xDm$m + xpn$a)($m + $8) - xmn(XJXrn)+m~J
where CpI is the volume fraction of component i, XIis the number of
segments per molecule for component i, n is the degree of polymer-
ization, R is the gas constant, T is the temperature, and x is the
free energy parameter between any two components, the subscripts
m, 8 , and p referring to monomer, solvent, and polymer, respectively.
Putting X,/X, = vI/v,, the ratio of the molar volumes, assuming
vJv, = n, and expressing xpmand xpl in terms of the molecular
weight- independent quantities xmDand xnpthrough the relationships
[81I Xpm = Xmg(VJVm) = Ymfl and pa = X ~ ~ ( V m / V a ) we
, find
AG,,*/RT = In CPm +1 + (xmm$m + xmD@D)(+# + 6,) - x ~ D ( v ~ ~ v 8 ) ~ E ~ D

+ Xrn,9m@m (91)
where +, now represents the equilibrium volume fraction of mono-
mer in the presence of a volume fraction of polymer equal to $D,
Neglecting terms in l / n when n is large, and replacing ($a + $,) by
-
(1 $m) and ($m + $,I by (1 -
= In + m + 1 +
AG~RT +a(Xmm - XapVm/Va) + X m s ( + p - +m) (g2)
Equation (B2) is more general than that deduced by Bywater [62],
who made the approximation v, = v,. AG,, represents the free-
energy upon the polymerization of 1 mole of liquid monomer to 1
base-mole of liquid amorphous polymer of infinite chain length,
282 SAWADA

In their work on the polymerization of a-methylstyrene, Ivin


and Leonard [79] found that the monomer concentration varies
linearly with the polymer concentration. Experimentally, it is found
that the variation of 9, can be expressed by

where (I, is the value of +,,,when 9, approaches zero and B is the


slope obtained from a plot of 9, against Qq. In all cases n is quite
large.
A linear dependence of (I,,, on 9, may be obtained from Eq. (92)
as follows [83]. First, In (Im is expanded in a series; that is,

-
In Q,,,= III a + ( ( I ~ a)/a - ((I,, - aY/2a2 + (@,,, + aI3/3a3 - -.-
(94)
where 0 < @,, < 2a. If the value of constant is such that ( Q m - a)'/2a2
is negligible with respect to the first two t e r m s of Eq. (941, one can
then write

-
Then, with ( I 8 = 1 (I, -
(I, and substituting Eq. (95) for (Im, after
rearranging Eq. (92), the following expression for the variation of
@,-with @, is obtained:

-
where p = xms xsp(v,/vs). With the exception of (I, and $I,, all the
terms appearing in Eq. (93) a r e assumed to be constant at a given
temperature.
Comparing Eqs. (93) and (96), it then follows that

B = (Xmp - + Xmp - l/a) (97)

- -
Since (I,,, (Imas 9, 0, the first t e r m on the right-hand side of
Eq. (96) is equal to (I,,. From this relation the variation of the
equilibrium monomer concentration with the polymer concentration
may be explained in terms of the constant a and the interactions
between the components of the system measured by the thermody-
namic parameter p and xmp. Since 9 , is a constant for a given
temperature, a can therefore be replaced by 9 , in Eq. (95).
TH E R MOD Y NAM ICS 0 F PO LY ME R I ZAT ION. I V 283

From the above considerations and comparing Eqs. (93) and (96),
one finally obtains
-(AG,JRT) + In a + p
@ml = =a (98)
P + Xmp - l/a
Using this definition of a, Eq. (96) may be rewritten to give

9, = a + (x,,,~- P ) / @ + Xmp - l/a)QP (99)


If (AG,JRT), 0 and xmPare known, and Eq. (98) can be solved for a.
Once a is known, it is then possible to calculate the variation of
@, with 9, using Eq. (99) and to compare it with experimental re-
sults. Table 8 gives experimental values of Q m and B obtained for
the anionic polymerization of a-methylstyrene in tetrahydrofuran
a t various temperatures and the corresponding values calculated by
means of Eqs. (98) and (99). The variation of $,,,with QP is shown
in Fig. 11 where a good agreement is obtained between the experi-
mental curve and curves computed using Eqs. (98) and (99).

0.0
0.0
, 0.1 0.2 0.3

Fig. 11. Variation of bm with aP


for the equilibrium polymerization of a-
methylstyrene in THF at 20 and -20°C 183). (Q): Experimental points.
(-) : Values calculated using xmp = 0.3
N
m
P

Table 8
Thermodynamic Parameters Used for the Computation of a,,, [83]

-20 0.030 -0.044 4.08 -1.65 0.031 -0.058 0.031 -0.061


0 0.100 -0.133 2.68 -1.52 0.102 -0.165 0.103 -+.I77
+ 20 0.279 -0.299 1.34 -1.38 0.282 -0.363 0.288 -0.400

aExperimental data from Ref. 79.


bWith xmp= 0.3.
'With xmp= 0.4.
THERMOClYNAMlCS OF POLYMERIZATION. IV 285

As can be seen, the effect of monomer-polymer interactions is


not predominant, and the same assertion can be made on the solvent-
polymer interactions. On the other hand, a change in monomer-
solvent interactions can bring about a pronounced change in the
position of equilibrium for the polymerization of a given monomer.
This can be shown by comparing the polymerization of cy-methyl-
styrene in tetrahydrofuran and cyclohexane at a given temperature.
We find

@, = 0.60 - 0.61@, (in tetrahydrofuran)


and similarly
@,, = 0.20 + 0.051@, (in cyclohexane)

Here, the solvent effect on the equilibrium position is obvious.


Changing the solvent from cyclohexane to tetrahydrofuran should
increase @a threefold. From the above results it can be seen that
the monomer concentration will decrease rapidly upon the addition
of polymer in tetrahydrofuran, whereas it will have little o r no
effect in cyclohexane, except for a slight depolymerization which
will occur upon the addition of polymer. Recently it has been shown
[84] that for a given temperature, the equilibrium monomer con-
centration in p-dioxane is approximately 20% lower than the corre-
sponding monomer concentration in tetrahydrofuran, together with
the variation of @,, with @., The effect is explained in terms of a
solvent-monomer and solvent-polymer interaction parameter.
The equilibrium polymerization of dioxolane has been studied
over a wide range of initial monomer concentrations at temperatures
from 20 to 60°C in methylene chloride, benzene, and 1,4-dioxane.
In all the solvents, over a wide range of initial concentrations, the
equilibrium monomer concentration falls with an increase in the
polymer content of the system [85]. A measure of the effect of the
solvent on the equilibrium concentration of dioxolane is given by
the value of x (the apparent Flory-Huggins parameter taking account
of polymer-solvent interaction). Thus the dependence of the equi-
librium monomer concentration on the quantity of polymer in the
system (or the initial monomer concentration), and on the solvent,
can also be explained on the basis of the thermodynamic theory of
polymer solutions [85].

Acknowledgment
The author wishes to thank Prof. K. F. O’Driscoll of University
of Waterloo for h i s continuing interest and helpful advice.
286_
_ SAWADA

References

[ 11 A. V. Tobolsky, J. Polym. Sci., 25,220 (1957).


[ 21 A. V. Tobolsky, J. Polym. Sci., 31, 126 (1958).
[ 31 A. V. Tobolsky and A. Eisenberg,J. Amer. Chem. Soc., 81,780 (1959).
[4] A. V. Tobolsky and A. Eisenberg,J. Amer. Chem. Soc., 81,2302 (1959).
[5] A. V. Tobolsky and A. Eisenberg,J. Amer. Chem. SOC., 82,289 (1960).
[ 61 A. V. Tobolsky, Properties and Structures of Polymers, Wiley (Interscience), New
York, 1960, p. 266.
[ 71 M. Szwarc, Proc. Roy. Soc., Ser. A, 279,260 (1964).
[ 8 ] A. Eisenberg, Makromol. Chem., 65,122 (1963).
[ 9 ] F. S. Dainton and K. J. Ivin, Nature, 162, 705 (1948).
[ l o ] S. Bywater, Trans. Faraday Soc., 51,1267 (1955).
[11] S. Bywater, Can. J. Chem., 35,552 (1957).
[12] C. G. Overberger and A.M. Schi1ler.J. Polym. Sci., Part C, 1, 325 (1963).
[13] H. W. McCormick,J. Polym. Sci.,25,488 (1957).
[14] D. J. Worsfold and S. Bywater, J. Polyrn. Sci., 26,299 (1957).
[15] S. Bywater and D. J. Worsfold,J. Polym. Sci., 58,571 (1962).
[ 161 A. V. Tobolsky, A. Rembaum, and A. Eisenberg,J. Polyrn. Sci., 45,347 (1960).
[ 171 R. Asami, in Kobunshi No Tenbo 1970 ( Y .Iwakura, ed.), Maruzen, Tokyo, 1970,
p. 53.
1181 P. A. Small, Trans. Faraday Soc., 49,441 (1953).
[19] K. J. Ivin, Trans. Faraday Soc., 51,1273 (1955).
[20] R. E. Cook and K. J. Ivin, Trans. Faraday Soc., 53,1132 (1957).
[21] H. Meerwein, D. Delfs, and H. Morschel, Angew. Chem., 72,927 (1960).
[22] D. Sims,J. Chem. Soc., 1964,864.
[ 2 3 ] C. E. H. Bawn, R. M. Bell, and A. Ledwith,Polymer, 6,95 (1965).
[24] B. A. Rosenberg, 0. M. Chekhuta, E. B. Ludving, A. R. Gantmakher, and S. S.
Medvedev, Vysokomol. Soedin., 6,2030 (1964).
[25] D. Vofsi and A. V. Tobolsky,J. Polym. Sci.,Part A, 3, 3261 (1965).
[26] M. P. Dreyfuss and P. Dreyfuss,J. Polym. Sci., Part A-Z,4,2179 (1966).
[27] K. J. Ivin and J. Leonard,Polymer, 6,621 (1965).
[28] D. Sims, Makromol. Chem., 98,235 (1966).
[29] F. Wiloth.2. Phys. Chem., 4,66 (1955).
[ 301 P. F. Van Velden, G. M. Van Der Want, D. Heikens, Ch. A. Kruissink, P. H.
Hermans. and A. J. Staveman, Rec. Trau. Chim. Pays-Bas, 74,1376 (1956).
[31] A. B. Meggy,J. Chem. Soc., 1953,796.
(321 0. Fukumoto,J. Polym. Sci., 22,263 (1956).
[ 3.71 H. K. Reimschuessel, in Ringopening Polymerization (K. C. Frisch and S. L.
Reegen, eds.), Dekker, New York, 1969, Chap. 7.
[ 3 4 ] H. Yumoto, J. Chem. Phys., 29,1234 (1958).
1351 H A . Elias and A. Fritz, Makromol. Chem., 114,31(1968).
[ 361 P. H. Plesch and P. H. Westermann, J. Polym. Sci., Part C, 16,3837 (1968).
[37] Y. Yamashita, M. Okada, K. Suyama, and H. Kasahara, Makrornol. Chem., 114,
146 (1968).
[38] F. S. Dainton and K. J. Ivin, Quart. Reu., 12,61(1958).
[ 391 S. M. Skuratov, A. A. Strepikheev, S. M. Shtekher, and S. V. Volokhina, Dokl.
Akad. Nauk SSSR, 117,263 (1957).
THERMODYNAMICS OF POLYMERIZATION. I V 28 7

[40] A. A. Strepikheev and A. V. Volokhina, Dokl. Akud. Nuuk SSSR, 99,407 (1954).
[41] P. H. Plesch and P. H. Westermann, Polymer, 10,105 (1969).
[42] T. Miki, T. Higashimura, and S. Okamura, J. Polym. Sci., Purt A - I , 5,95 (1967).
[43] T. Miki, T. Higashimura, and S. Okamura, J. Polym. Sci., Purt A - I , 5,2997 (1967).
[44] T. Miki, T. Higashimura, and S. Okamura, J. Polym. Sci., Purt A - I , 8,157 (1970).
[45] A. A. Berlin, S. A. Vol’fson, E. F. Oleinik, and N. S. Yenikolopyan, Vysokomol.
Soedin., A12,443 (1970).
[46] W. K. Busfield and D. Merigold, Mukromol. Chem., 138,65 (1970).
[47] W. K. Busfield and D. Merigold, J. Chem. Soc., 1969, A 2975.
[48] K. Nakatsuka, H. Suga, and S. Seki, J. Polym. Sci., Purt B, 7,361 (1969).
[49] G. Gee, Trans. Furuduy Soc., 48,515 (1952).
[50] F. Fairbrother, G. Gee, and G. T. Merrall, J. Polym. Sci., 16,459 (1955).
[51] A. Eisenberg and A. V.Tobolsky,J. Polym. Sci., 46,19 (1960).
[52] I. Mita, I. Imai, and H. Kambe, Mukromol. Chem., 137,143 (1970).
[53] W. K. Busfield and E. Whalley, Trans. Furuduy Soc., 59,679 (1963).
[54] I. Mita, I. Imai, and H. Kambe, Mukromol. Chem., 137,155 (1970).
[55] F. S. Dainton, K. J. Ivin, and D. A. G. Walmsley, Trans. Furuduy SOC., 55,61
(1959).
[ 561 Y. Iwasa and T. Imoto, J. Chem. SOC. Japan, Pure Chem. Sect., 84,29 (1963).
[ 571 G. S. Parks and H. P. Mosher, J. Polym. Sci., Purt A , 1, 1979 (1963).
[ 581 A. V. Tobolsky and A. Eisenberg, J. Colloid Sci., 17,49 (1962).
[ 5 9 ] H. Sawada, J. Mucromol. Sci., C3(2), 313 (1969).
[60] A. V. Tobolsky, Polym. Preprints, 11,1,165, February 1969.
[61] P. J. Flory, J. Amer. Chem. SOC.,58, 1877 (1936).
[62] P. J. Flory, J. Amer. Chem. Soc., 64,2205 (1942).
[ 6 3 ] P. J. Flory, J. Chem. Phys., 12,114 (1944).
[ 6 4 ] A. V. Tobolsky, J. Chem. Phys., 12,402 (1944).
[65] W. Kuhn, Chem. Ber., 63,1503 (1930).
[ 6 6 ] P. J. Flory, J. Chem. Phys., 12,425 (1944).
[67] R. S. Jessup,J. Chem. Phys., 16,661 (1948).
[68] M. Born and R. Oppenheimer,Ann. Phys. (Leipzig), 84,457 (1927).
[ 691 P. J. Flory, Principles of Polymer Chemistry, Cornell Univ. Press, Ithaca, New
York, 1953, p. 502.
[70] J. L. Lundberg,J. Polym. Sci., Purt A , 2,1121 (1964).
[ 711 H. L. Finke, M. E. Gross, G. Waddington, and H. M. Huffman,J. Amer. Chem.
Soc., 76,333 (1954).
[ 721 F. S. Dainton and K. J. Ivin, in Experimental Thermochemistry, Vol. 2 (H. A.
Skinner, ed.), Wiley (Interscience), New York, 1962, p. 253.
[ 731 F. E. Harris, J. Polym. Sci., 18,351 (1955).
[ 741 M. Szwarc, Nature, 178,1168 (1956); M. Szwarc, M. Levy, and R. Milkovich,
J. Amer. Chem. Soc.. 78,3590 (1956).
[75] W. B. Brown and M. Szwarc, Trans. Furuduy SOC., 54,416 (1958).
[ 7 6 ] P. J. Flory,J. Amer. Chem. SOC., 62,1561 (1940).
[ 771 A. Miyake and W. H. Stockmayer, Mukromol. Chem., 88,90 (1965).
[ 78 1 V. S. Nanda and S. C. Jain, Eur. Polym. J., 6,151 7 (1970).
[79] K. J. Ivin and J. Leonard,Eur. Polym. J., 6,331 (1970).
[80] A. Vrancken, J. Smid, and M. Szwarc, Trans. Furuduy Soc., 58,2036 (1962).
[81] Ref. 69, p. 549.
288 SAWADA

[82] S.Bywater, Makromol. Chem., 62,120(1962).


[83] J. Leonard, Macromolecules, 6,661 (1969).
(841 J. Leonard and S. L. Malhotra,J. Polym. Sci., Part A-1, 9,1983(1971).
[85) L. I. Kuzub,M. A. Markevich, A. A. Berlin, and N. S. Enikolopyan, Vysokornol.
Soedin., 10,2007(1968).
J. MACROMOL. X I . - R E V S . MACROMOL. CHEM., C10(2),293-353(1974)

Thermodynamics of Polymerization. V.
Thermodynamics of Copolymerization. Part I
HIDE0 SAWADA
Filter Laboratory
Daicel Ltd.
Teppo-cho, Sakai, Osaka, Japan

I. THE GENERAL THEORY OF BINARY COPOLYMERIZATION . . .294


A. Heat of Copolymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
B. Entropy of Copolymerization . . . . . . . . . . . . . , . . . . . . . . . . . . . . . 301
C. Equilibrium Sequence Distribution . . . . . . . . . . . . . . . . . . . . . . . . . 304
D. Free Energy Change in Binary Copolymerization System. . . . . . . . . 310
E. Equilibrium Monomer Concentration . . . . . . . . . . . . . . . . . . . . . . . 318
F. Penultimate Unit Effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
11. DEGREE OF POLYMERIZATION AND COPOLYMER
COMPOSITION OF BINARY COPOLYMERIZATION SYSTEM . . . . . 325
A. Degree of Polymerization.. . . . . . . . . . . . . . . . . . . . . . . . . . . , . . . . 325
B. Copolymer Composition Equation. . . . . . . . . . . . . . . . . . . . . . . . . . 329
111. MULTICOMPONENT COPOLYMERIZATION . . . . . . . . . . . . . . . . . . . 344
A. Heat of Terpolymerization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
B. General Theory of Multicomponent Copolymerization . . . . . . . . . . 349
ACKNOWLEDGMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
REFERENCES ........................................... 352

293
Copyrighf 0 1974 by Marcel Dekker. Inc. All Rights Reserved. Neither this work nor any part may be
reproduced or transmitted in any form or by any means. electronic or mechanical, including photocopying,
microfilming, and recording, or by any information storage and retrieval system. without permission in
writing from the publisher.
294 SAWADA

We will discuss the main features of copolymerization, with par-


ticular emphasis on the reversibility of copolymerization. The
treatment is a general one, and heat of copolymerization, entropy
of copolymerization, and free energy change of copolymerization
a r e applicable to all types of copolymerizations. Before proceeding
to a discussion of some specific types of copolymerization, it will
be worthwhile to discuss the general aspects of thermodynamics of
copolymerization.

I. THE GENERAL THEORY OF BINARY COPOLYMERIZATION


A. Heat of Copolymerization
In the case of copolymerization involving monomers M, and M,,
there a r e four distinct propagation steps. Each of these will have a
characteristic molar heat of reaction associated with it, H,,, H,,,

WM: + M, --
H,,, and H,,, respectively.

M,-M: H,,
%M: + M,
WAM; + M,
'M MZ + M,
-
--+
--A

wM,-M:
M,-M;
M,-M;
Hl2

H,,
H21

Let us consider low conversion copolymers of high degree of


polymerization containing N, monomer units of type M,, and N,
monomer units of type M,. There will be four types of bonds to
consider: 1) those between two monomer units of type M,, 2) those
between two monomer units of type M,, 3) those between two mono-
mer units M, and M,, and 4) those between two monomer units M,
and M,. In these copolymers made from 1 mole of monomer mix-
tures there will be N,, bonds of type M,-M,, N,, bonds of type
M,-M,, N,, bonds of type M,-M,, and N,, bonds of type M,-M,.
The total number of bonds in these copolymer is No. Thus the
molar heat of copolymerization will be given by

N N N N
AH = 2 H,, + A H z 2 + -H1, + ANOH , ,
NO NO NO
A s the reactions between M: and M, and M,* and M, forming M,-M$
and M,-M:, respectively, occur with equal frequency during the
polymerization,
Nl, = N,, (2)
THERMODYNAMICS OF POLYMERIZATION. V. 295

The next step is to calculate Nll, N,,, and Nl,. Before attempting
this, the treatment may be simplified by elimination of N,, and N,,
as follows. The copolymers contain N, monomer unit of type M, and
each of these has two bonds. Each M,-M, bond contains one mono-
mer unit of type M,, whereas each M,-M, bond contains two mono-
mer units of type M,. Further, each monomer unit is shared with
two bonds, thus the number of monomer units of type M, involved in
M,-M, and M,-M, bonds is (N,, + N,,)/2, and the number in type
M,-M, bonds is N,,. Since there are N, monomer units of type M,,

N, = + Nzl
2
+ N,, = N,, + N,, (3)

Similarly, for monomer units of type M,

N, = N,, + N21
2 + N,, = N,, + N,,

Thus N,, and N,, can be expressed in terms of N,,, N,, and N,, From
EqS. (1)-(4):

Equation (5) gives the molar heat of copolymerization.


If the two types of monomer units a r e arranged completely at
random in the copolymer chain, N,, can be readily evaluated from
simple statistical considerations. Let us consider copolymers made
from 1 mole of monomer mixtures. When the molecular weight of
copolymers is high, No is given by

No = N, + N,

and hence the mole fraction of N, groups in the copolymer is

and similarly

X, = N,/No
The probability that a monomer unit of type M, will be on a given
site is XI in the case of a random copolymer. The probability that a
296 SAWADA

monomer unit of type M, will be on a neighboring site is X,, and the


probability that both will be on their respective sites simultaneously
is X,X,. The total number of bonds in the copolymers is No; there-
fore, the number of MI-M, bonds will be equal to the total number
of bonds multiplied by the probability that a bond will be of the
M,-M, type. Thus

The copolymerization parameter 51 is defined a s

51 = H I 2 + H,, - (Hll + H,,) (8)

We find that the molar heat of copolymerization is

Since 51 is independent of composition, XIX,s2 is shown to be a par-


abolic function of composition (Fig. 1). In the case H,, = Hzz, the
function is symmetric about X, = X, = 0.5 (Fig. 2). (Note that any
gain in AH on the part of a system is considered to be positive and
any loss negative, but this is not considered here for simplicity.)
Let us consider an M, monomer unit in the copolymer surrounded
by two neighbors. If the copolymer is completely random, the prob-
ability that an M, monomer unit would occupy a site adjoining an M,
monomer unit in this copolymer is simply X , . For a copolymer ex-
hibiting regular alternation o r long sequences, the probability of
occupancy will be P,, f XI. The randomness parameter JI is defined
a s ill

For complete randomness, J, = 1; for some regular alternation, J,


will be greater than unity; for long sequences, JI will be less than
Unity.
The number of MI-M, pairs, N12, may be readily expressed i n
terms of JI. Rearranging Eq. (10) we find
THERMODYNAMICS OF POLYMERIZATION. V. 297

- 0
0 0.2 0.4 0.6 0.8 1.0

XI
Fig. 1. Heat of copolymerization as a function of composition. H,, = 13 kcal/mole;
H22= 18 kcal/mole. !J = 20 kcal/mole (1);10 kcal/mole (2); 0 kcal/nlole (3);
-10 kcallmole (4);and -20 kcal/mole (5) [ 1 1.

The average number of M, monomer units which surround a given


M, monomer unit is 2P,,, so therefore this is the number of MI-M,
and M,-M, bonds per M, monomer unit on the average. The num-
ber of M, monomer units in the copolymer is NoX2, and thus the
total number of MI-M, pairs is

N,, = N,, = N,X,P,, = N,X,P,, (12)


where Pi, is the conditional probability that an M, monomer unit
selected at random will be followed by an M, monomer unit. Sub-
stitution of Eq. (12) into Eq. (5) yields [l, 21
298 SAWADA

20

-a,

-.
0
E

x
3 10

I
a

n I I I I
0 0.2 0.4 0.6 0.8 1.0

XI
Fig. 2. Heat of copolymerization as a function of composition. H,, = Hzz = 15
kcal/mole. a = 20 kcal/mole (1); 10 kcal/mole (2);0 kcal/mole (3);-10 kcal/mole
(4);and -20 kcal/mole (5) [ 11.

In the case of radical copolymerization, Alfrey and Lewis [2] ob-


tained an expression for X2P2,as
1 - [l - 4X2(1 - X2)(1 - rlr2)]1’2
XZP,I = 2(1 - rlrJ

The reactivity ratios rl and r2a r e defined as kl,/kI2 and k,,/k,,,


respectively. Substituting Eq. (11)into Eq. (13), we find

AH = HllX, + H,&, + X1X,J1S2 (15)


Since the reaction is proceeding at constant temperature and
pressure, AH is the enthalpy change of the copolymerization. How-
ever, the reaction is assumed to take at moderate pressure with
negligible change in volume. Therefore, the enthalpy and the internal
energy of copolymerization can be considered equivalent. Equa-
tion (15) is applicable to all types of copolymerizations (e.g., radical
THERMODYNAMICS OF POLYMERIZATION. V. 299

o r ionic copolymerization). For radical copolymerization at low


conversion, close agreement between experimental values and cal-
culated values from Eq. (15) was obtained, a s will be described in a
later section.
For all copolymerizations except azeotropic copolymerizations,
the comonomer feed and copolymer compositions a r e different. The
comonomer feed changes in composition a s one of the monomers
preferentially enters the copolymer. When a high conversion co-
polymer is obtained, it is a blend of copolymers made from a con-
tinuously varying charge ratio. Therefore, Eq. (15) cannot be applied
to copolymers at high conversion. In the case of high conversion
copolymers, it is necessary to divide the polymerization into an
arbitary number of successive stages, each of sufficiently short
duration that [X,]/[X,] may be considered constant [3].
By combining Eqs. (11) and (14) it is found that

and if 1 - r,r, 2 0, Eq. (16) will reduce to

$J 1 + X,X,(1 - r,r,) (17)

The calculated values of J ) a r e shown a s a function of XI in Fig. 3. It


is observed that the function is symmetric about XI = X, = 0.5.

\\
1
l l l l l l l l l l l
0 02 04 06 08 10

Fig. 3. Randomness parameter as a function of composition [ 11.

If r1r2= 1, the two types of monomer units a r e arranged com-


pletely at random in the chain, so that J, will be unity. If rlrz is less
300 SAWADA

than unity, J, will be greater than unity, and the structure deviates
from random in the direction of regular alternation. If r1r2is
greater than unity, J, will be less than unity and the self-propagation
reactions will be favored over the alternation reactions. In the case
r, = r, = 0, the monomers alternate regularly along the chain, re-
gardless of the composition of the monomer feed. The mole fraction
X, is then one-half for all monomer compositions. Hence, $ will be
greater than unity (11, = 2).
Expressing the rate constant according to the Arrhenius equation,
we have

k = A expi-E/RT} (18)

where E and A are the activation energy and the frequency factor,
respectively. We may write each of these rate constants in the form
given by:

The reactivity ratio r l is

The relation between the heat of polymerization and the activation


energy has a quantitative form [4],viz.,

-AE = CYAH 0 < a <1 (20)


This equation says that in a related series of reactions, the decrease
in the activation energy AE is a fraction a of the increase in the
heat of polymerization AH. Equation (20) can be integrated to give

where C is a constant.
Substituting Eq. (21) into Eq. (19), we find

rl = (A,,/Al,) exp((H,, - H,,)dRT}


and similarly
THERMODYNAMICS OF POLYMERIZATION. V. 301

A s a result

is a function only of the r1r2product rather than of the individual


values of r, and r, a s we will discuss in P a r t 11. This is an inter-
esting result, since the r,r, product has been shown to be an index
of the alternation tendency in copolymerization. In Part I1 we will
also discuss the quantitative data concerning the heat of radical
copolymerization .
It should be noted that configurational irregularities such as
head-to-head and tail-to-tail modes o r stereoirregular placements
of monomers are not examined, and, in addition, the formation and
possible effects of more than one phase are not considered here.
B. Entropy of Copolymerization
In the case of copolymerization involving monomers M, and M,,
there a r e four distinct propagation steps. Let S,,, S,, S,,, and S,,
represent the corresponding entropy changes, per mole, for the
reactions M,-M,, M,-M,, MI-M,, and M,-MI, respectively. It is
considered that the contribution of the growing polymer chain to
entropy changes was determined mainly by the nature of the ter-
minal unit. Thus the entropy change per mole of monomer poly-
merized is given by [5]

where

x = (S,, + S,) - (S,, + S22) (24)


In addition, there a r e some other entropies in copolymers to be
considered as follows:

(1) Residual entropy at 0°K by the glass state.


(2) Entropy due to chain configuration.
(3) Entropy due to copolymerization randomness.

O r r [3] concluded that the contribution to the entropy at 0°K by


the glassy state was negligible. The contributions from the con-
302 SAWADA

figurational entropies were discussed by Temperley [6] who decided


they were quite small.
Let us now compute the entropy of copolymerization randomness,
which is an entropy of mixing. The probability of a sequence of M,
units L units in length is given by

P =Py(1 - P,,) (25)


If the number of sequences per chain is 2n (of which half must be MI
sequences and half M, sequences), the number of MI sequences of
length L is

In such a case, let there be a, sequences of one M, unit, a, sequences


of two M, units, . . ., and ay sequences of yM, units, thus

n = a , + a , + a, + - - - + ay (27)
The number of permutations (W) open to these MI sequences is

"I!
W =-
%!a,! a,! - - - a,!
In dealing with entropy calculations, we shall be interested in
values of In W and hence be concerned with the logarithm of factor-
ials. There is a convenient approximation, useful for this situation,
known as Stirling's formula which is applicable only when n is very
large. Stirling's approximation is given by

Inn! g n l n n - n

Thus, for large values of a,


In W = n In n - a, In a,--. - . -ay In ay
The number of M, units in the copolymer is

a, + 2a, + 3a, -I. - - - + yay


The entropy of mixing per mole is given by

AS,,,,== R In W = n l n n - a, l n a , - - - aY In
- . + yay
- a a Y
a, + 2a, + 3a, +
THE R MODY NAMlCS OF POLYMERIZATION. V. 303

If y is very large, Eq. (31) may be simplified by expanding in


t e r m s of P,, o r P,,. It may be readily shown that

1
a, + 2a, + 3a, + . . + yay 2 1 - PI1

Therefore
I. Y
- pl1)+-
C
I=l
a, ~n
a,
[ n+
n ~n In P,,] (33)

These equalities become exact only when y is infinite. Appropriate


substitution into Eq. (31) then yields

AS,,, = -R [PI, In PI, + (1 - Pll) h ( 1 - PI,)] (34)

Equation (34) is only exact if the permitted number of units in


the longest sequence is infinite and PI, = P,,. However, there was
no significant difference in values of copolymerization entropy cal-
culated from Eq. (31), taking n = 100, and those calculated from
Eq. (34), assuming infinite molecular weight.
We see that the entropy from Eq. (34) has the proper limiting be-
havior: in the case of PI, = X I , AS,,, = -R(X, In XI + X, In X,),
which is the entropy of random copolymers. In the case of P,, = P,,
= 1, ASmIx= 0 in agreement with what is expected for block copoly-
mers. F o r alternating copolymers, PI, = P,, = 0; so AS,,, = 0.
Thus the total entropy change per mole involved in the copoly-
merization reaction, AS,,,, is the sum of entropy changes associated
with copolymerization reactions and the entropy of copolymerization
randomness.
AS,,, = -(S,,X, + s,zx, + XWIX,)

- R [(I - $J x,) (1 - J, x , ) + $J x, JI x,l (35)

O r r [3] and North and Richardson [7)indicated the entropy of co-


polymerization randomness to be 0.5 to 1.5 cal/deg mole of re-
peating units, therefore the contribution of this entropy will seldom
be responsible for more than 5% of total entropy change.
Theil [8]expressed the entropy of sequence length distribution
per mole in the copolymers a s
AS = -R(X,[(l - Pll) + PI, In PI,]
+ XJ(1 - p22) In (1 - P Z Z )+ p,, Pzzl) (36)
304 SAWADA

Equation (36) differs from Eq. (34) which does not recognize that
the entropy of sequence length distribution includes the contribu-
tions of the separate arrangements of both the MI and the M,
sequences.
To define mathematically the concept of copolymer Urandomness,”
Tosi 191 introduced a new quantity, which h e proposed be called
%formational entropy.” It is defined a s the ratio between the nat-
ural logarithm of the number of ways of disposing of the homose-
quences of a copolymer and the number of monomer units contained
in the copolymer; it is calculated by applying the methods of infor-
mation theory. Information theory originally arose as an application
of probability theory and statistical methods to problems connected
with the translation of messages into codes.
When applied to copolymers, information theory allows one to
calculate the information gained in knowing whether a certain posi-
tion in the chain is occupied by monomer unit MI o r by monomer
unit M,. It is apparent that some analogy between information and
entropy must exist. Information is a negative t e r m in the entropy
of a system. On the other hand, entropy measures the lack of infor-
mation; it gives us the total amount of missing information on the
ultramicroscopic structure of the system.
.
C Equilibrium Sequence Distribution
A s pointed out earlier, it is clear that equilibrium polymeriza-
tion-depolymerization may govern the molecular weight distribu-
tion at high temperatures. By the same token it is expected that if
a copolymer were treated in this manner, depolymerization-poly-
merization would determine both the molecular weight distribution
and the structural sequence distribution of the copolymer. We can
assume equilibrium polymerization-depolymerization will be
achieved at quite high temperatures and will determine the copoly-
mer structure and composition.
Alfrey and Tobolsky [ l o ] first treated the case of an infinite
copolymer molecule mathematically to determine the sequence
distribution along the chain.

= exp(AG,,/RT)

=K (37)
where K is an equilibrium constant which is expressible in terms of
vibrational partition functions f,,, fZ2, and f,, and the energy change
THERMODYNAMICS OF POLYMERIZATION. V. 305

AE,,. The partition functions f,, and f,, include a symmetry factor
of 2. The quantity AE,, is defined as 2E,, - Ell - E,,, where El, is
the energy per mole of an MI-M, bond, E,, is the energy per mole
of an M,-M, bond, and El, is the energy per mole of an MI-M, bond.
The energy Ell is approximately the heat of polymerization p e r mole
of monomer MI and similarly for E,, and El,. AG,, is 2G,, - GI,
- G,,, and GI, is the molar free energy of formationof an M,-M,
bond, etc.
The distribution in sequence lengths is found to be given by the
formula

nk = N,Pf;l(l - PI,),
n: = N,P,X;'(l - P,,)2
where n: is the number of MI sequences of length x in the copolymer
chain, n: is the number of M, sequences of length x in the copolymer
chain, and PI, and P,, a r e defined as
PI, = (N, - N,,)/N, (38)

P2, = (N, - N,J/N, (39)

Combining Eqs. (12), (38), and (39) with Eq. (37), the following equa-
tion is obtained:
P,,P,l
-- - exp(- AG/RT)
p,,p,,

A similar conclusion was drawn by Theil [8] and by IZUand


O'Driscoll [ll].
Consider the copolymerization reaction as occurring at some
constant temperature, T. For the sake of illustration, let u s con-
sider the copolymers which are 50% MI and 50% M,, and the change
in free energy for this reaction, AG, is given by
AG = AH - TAS
= -1/2 [HI, + H,, + (W/2)1 + RT{[1 - (J1/2)1 - (Jl/2)1)
+ (T/2)[S,, + S, + (Jlx)/Zl (41)*
*Instead of AF, it is more convenient theoretically to use AG, the Gibbs free energy,
and thus speak of a constant pressure process. Note that any gain in a quantity o n the
part of a system is considered to be positive and any loss negative. Since heat is evolved
and randomness for the polymer decreased relative to the monomer during the polymeri-
zation, copolymerization reactions are exothermic (negative AH) and exoentropic
(negative AS).
306 SAWADA

Interchange reactions occurring in copolymer chains will ulti-


mately lead to copolymer equilibrium; that is, an equilibrium dis-
tribution of compositional sequence lengths along the chains. This
equilibrium sequence is determined as a function of temperature by
the requirement that the change in free energy AG be a minimum
with respect to the randomness parameter J,.Thus the condition
aAG/aJ, = 0 gives [5]

9 = 2/[1 + exp{(-n/aRT) + (x/~R)}I (42)

Equation (42) determines the equilibrium J, as a function of T, pro-


vided 51 and x a r e given.
In Fig. 4, J, is plotted as a function of parameter X = 2RT/S1 for
t h e case where x = 0 and 52 > 0. Regardless of the sign of s2 terms,
J, will always be 2/(1 + ex&/2R}) at sufficiently large value of T,
-
since 3 2/(1 + exph/2R)) a s T -00. Thevalues of entropy changes

for most polymerizations a r e quite similar [12]; thus x will be very


close to zero. If x = 0, the copolymer will be random at sufficiently
elevated temperatures since 3 - 1as T -
03. When a, x a r e zero,

at any temperatures JI is unity, and hence the two types of units will
be arranged a t random along the polymer chains. When 52 > 0, at
low temperatures J, is very close to 2, and hence the polymer struc-
ture deviates from the random in the direction of regular alternation.

2.0

1.5

1.0
0 2 4 6 8 10 12

X = 2RT/Q
Fig. 4. Randomness parameter as a function of 2RT/Cl[ 51.
THERMODYNAMICS OF POLYMERIZATION. V. 307

If x > 0, there must be a temperature at which --S1/2RT + x/2R is


zero. This temperature is given by

T = S2/x (43)
At this temperature the copolymer will be random.
Let us consider the energetics of attack of radical on a double
bond. The weaker the bond formed between the attacking radical and
the carbon center of the double bond, the higher the activation
energy [13]. The more exothermic the reaction, the lower the activa-
tion energy. Therefore, the more exothermic the reaction between
MI and M,, the stronger the bond formed between M, and M,.
If there is an attractive interaction between unlike monomer units
(large electrostatic interaction of charges on the radical and mono-
mer for example), the MI-M, bond will be stronger than an M,-M,
bond o r M,-M, bond, hence a will be positive in sign, yielding regu-
lar alternation. On the other hand, Joshi [14] found experimentally
that the monomer units alternate regularly along the chain only
when S2 > 0. Conversely, if there is repulsive interaction between
unlike monomer units, -S1 will be negative and repeating units occur
in blocks.
An alternative approach [ 151to the problem is based on Boltzmann
statistics, and almost similar conclusions are reached. Consider
the copolymers with equal number of MI and M, monomer units, XI
= X, = 0.5.
If the average enthalpy change on formation of like bonds
-(Hll + H,,)/2 is greater than that of unlike bonds -(HI, + H2,)/2,
we then have the simple energy level system indicated in Fig. 5. If

I
AH'

Fig. 5. Energy relation between unlike snd like bonds. [ 5 ]

the change in entropy for the reaction is zero in thermal equilibrium,


the relative probability of occupancy of the high energy state to that
308 SAWADA

of the low energy state is given by Boltzmann statistics:

energy state/Plow energy state = exP{-AH’/RT} (44)

The probability of occupancy of high energy state is

Similarly the probability of the low energy state is

Substitution of Eqs. (45) and (46) into Eq. (44) yields

JI = 2/[1 + exp{--Sl/2RT}] (47)

where

!2 = 2AH’ = H,, + H,, - (H,, + HZ2)


F ro m Eqs. (15) and (47) we obtain:
2x,x2n
1 + exp(-n/2RT}

for X, = X, = 0.5.
If CZ is positive at low temperatures, AH will approach -(Hi, +H,,)/2,
yielding alternating copolymers. If 0 is negative at low tempera-
tures, AH will approach -(Hll + H,,)/2, yielding block copolymers.
Regardless of the sign of $2t e r m s at high temperatures, AH will
approach -(H,, + H,, + H,, + H,,)/4, yielding random copolymers.
If G? is zero, the copolymer will be random at any temperatures and
hence AH is given by
AH = -(HI1 + H,J/2 = -(HI2 + H,,)/2
= -(HI2 + HZ1 + Hi, + H22)/4
THERMODYNAMICS 0 F POLYMER I ZATl ON. V. 309

Copolymerization reactions will exhibit ceiling temperatures [16].


A t the ceiling temperature T,, AG is zero, so that
Q ln(1 + exp{-Q/2RT,$
HI, + H,, + 1 + exp{-52/2RTJ + 2RTc 1 + exp{-52/2RTJ

+ In(' + exP{-52/2RT3) - T,(S,, + S,,) = 0


(49)
1 + exp{G?/ZRTJ

Equation (49) does not lend itself readily to the calculation of T,.
However, T, is found by solving Eq. (49) when HI1, H2,,SI1,and S,,
a r e known and f2 = 0. Thus we obtain the ceiling temperature for
the copolymer where X, = X, = 0.5 and x = G? = 0. The ceiling tem-
perature for t h i s case is

T, = (H,, + H,,~/~Sll+ S,, - 2R In 2)


= [(HI, + H,, + H,, + H,,)/2]/(S1, + S,, - 2R In 2)
This is the ceiling temperature for random copolymers.
The number-average sequence length of MI monomer units is
given by the ratio of the probability of randomly selecting an MI
monomer unit, X I , to the probability of occurrence of an M,-M,
bond, X,P,,. Thus

w,,= UPl2 = 1/(1 - PI,) = m x , )


Similarly

w, = U P , , = 1/(1 - P2,) = l / ( W , )
The variation of the number-average sequence lengths of the
monomer units in a binary copolymer is shown a s a function of
temperature in Fig. 6 [8]. When 52 is negative (corresponding to a
block copolymer), both W,, and W, increase as the temperature
decreases. The effect of varying 52 on the relationships between
temperature and sequence length is shown in Fig. 7 [8]. Negative
values for Q cause the sequence length in Curves l a and l b to in-
crease a t low temperature and to remain consistently higher than
that for random copolymer which is formed at all temperatures
when 52 = 0 kcal/mole (Curve 2). When 52 is a positive quantity, the
sequence length decreases with temperature until the MI monomer
units a r e nearly isolated from one another at -100°C (Curve 3).
310 SAWADA

-100 0 100 2 00 300

TEMPERATURE("C)

Fig. 6. Number-average sequence lengths vs copolymerization temperature for a


copolymer for which XI= 0.2,52 = 2.0 kcal/mole, and x = 0.0 eu [ 81.

D. Free Energy Change in Binary Copolymerization System


In the first place, let us consider the general shape of a free
energy change vs composition curve for a binary copolymerization
system in the vicinity of either end of the binary [17]. The free
energy change of copolymerization a t some particular composition
will be given by the equation
AG = AH - TAS

To simplify the following treatments, it is assumed that JI is in-


dependent of X,. This assumption can be regarded only as a first
THERMODYNAMICS 0 F POLYMER I ZATl ON. V. 31 1

20.0

10. 0

8.0
\
2 5.0

3. 0

2.0
2

1.0
-100 0 100 2 00 300
TEMPERATURE ("C)

Fig. 7. Number-average sequence lengths of M I monomer units vs copolymerization


temperature for a copolymer for which XI = 0.4: (la) Sl = 2.0 kcal/mole, x = 0.0 eu;
( l b ) Sl = 2.0 kcal/mole, x = 2.0 eu; (2) f2 = 0.0 kcal/mole, x = 0.0 eu; (3) f2 = -2.0
kcal/mole, x = 0.0 eu [ 81.

approximation, since there i s no doubt that $ depends on X, with the


exception of random copolymers, alternating copolymers, and block
copolymers. However, in the vicinity of the end of binary, $ is in-
dependent of X, and $ = 1. Differentiating AG with respect to X, at
constant temperature, we find
312 SAWADA

In the vicinity of the end of binary, we may say that (1 - JlX,) 1.


Thus

+ RTJl In JlX, (52)


-
If J, is not zero, so that In JlX, -03 a s X, -
0, aAG/aX, will
always be negative at sufficiently small values of X,, regardless
of the sign of any term. Thus the introduction of the first comono-
mer unit into a homopolymer will always result in a decrease in
free energy of the system; hence, a pure homopolymer in the pres-
ence of a comonomer is always thermodynamically unstable. This
conclusion has important ramifications in connection with the im-
portant process of copolymerizing impurities such as oxygen from
its environment [18].
If 51 is positive and x is negative o r zero, AG a s a function of X,
will be concave downward at all compositions below the ceiling tem-
perature a s shown in Fig. 8. Let us consider a general case for
which 51 is a large positive value. It is assumed that x = 0 from
experimental data of O r r [3] as well as from his theoretical dis-
cussion. It therefore allows that x has comparatively little effect on
AG although 0 affects AG markedly. At a temperature T, well below
the ceiling temperatures of either component, the free energy change

u
vs composition curves would appear schematically a s shown in Fig. 8.

-TIAS

1.01
0 0

0
f -
Q

Fig. 8. Free energy change vs composition for a copolymer when i2 is positive and
T = T , [17].
THERMODYNAMICS OF POLYMERIZATION. V. 313

Above the ceiling temperature of both components at temperature T,,


the curves would appear a s shown in Fig. 9. This gives rise to a
ceiling temperature maximum and a modified phase diagram a s
shown in Fig. 10.

t
(3
Q

Fig. 9. Free energy change vs composition for a copolymer when i2 is positive and
T = T, [ 171.

When 51 is a large positive quantity, the introduction of two types


of monomer unit into a polymer chain raises the ceiling tempera-
ture for polymerization. Therefore the maximum ceiling tempera-
ture occurs at a polymer structure of the regular alternation. The
alternating copolymerization of sulfur dioxide and isobutene might
appear to be a typical example of maximum ceiling temperature [19].
A positive value of s1 reflects on attractive interaction between
unlike neighbors. Since this attraction would be more pronounced
in the copolymer, and because of the strong bonding resulting from
this attraction, the ceiling temperature of the copolymer relative to
both homopolymers would be increased.
It is easily shown that the maximum ceiling temperature also
occurs when 51 0. This result is consistent with the experimental
data of Richardson and North [ 203 who found the maximum in the
ceiling temperature for the copolymerization of acetaldehyde and
propionaldehyde.
314 SAWADA

I 1

T
I

I-

-
II : I
I
I
I
M1 a M2
x 2

Fig. 10. Modified phase diagram corresponding to free energy curves of Figs. 8 and 9
(at T, the free energy diagram is given by Fig. 9.): (1) monomer; (2) copolymer; T, =
ceiling temperature [ 171.

Let us now consider the case of a large negative value of Q. A t


some temperature T, below the ceiling temperature of the homo-
polymers, the curves would appear a s shown in Fig. 11. The free
energy curve for the copolymerization would exhibit minima since
f2 < 0. The most stable situation at various compositions is:
(a) 0 < X, < a, polymer is stable; (b) a < X, < b, monomer is
stable; (c) X, > b, polymer is stable.
This gives rise to a ceiling temperature minimum and results
from a repulsive interaction between unlike nearest neighbors and
hence a weakening of the copolymer structure. In the case of a re-
pulsive interaction, s2 will be negative and clustering will occur [7].
The modified phase diagram associated with this type of system is
shown in Fig. 12, with Fig. 11 being associated with TI.
There will be a repulsive interaction between unlike nearest
neighbors in the copolymerization of vinyl monomer with cyclic
monomer, and hence the cross-propagation reaction occurs with
difficulty [21]. In this copolymerization must be negative, indicat-
ing the production of block copolymers [71. Tsuda and Yamashita[21]
THERMODYNAMICS OF POLYMERIZATION. V. 315

Fig. 11. Free energy change vs composition for a copolymer when i2 is a large negative
value and T = T, [ 171.

Fig. 12. Modified phase diagram corresponding to free energy curve of Fig. 11 (at
T, the free energy curve is shown by Fig. 11): (1) monomer; (2) copolymer;
(3) two copolymers of different compositions; T, = ceiling temperature [ 171.

found the block copolymer for the copolymerization of styrene and


P-propiolactone.
If a case is considered for which 51 is l e s s negative than in the
former case, the ceiling temperature minimum is not realized.
316 SAWADA

This modified phase diagram i s shown in Fig. 13. A s the tempera-


ture is lowered, however, the AH term will begin to dominate in the
free energy expression. A s discussed above, near the ends of the
binary ACi is always negative, so the complete curve will be shown
as in Fig. 14.

Fig. 13. Modified phase diagram for a copolymer when S2 is a small negative value
(at T, the free energy curve is given by Fig. 14.):(1) monomer; (2) copolymer; (3)
two copolymers of different compositions; T, = ceiling temperature [ 171.

Fig. 14. Free energy change vs composition for a copolymer when 52 is a small nega-
tive value [ 17 1.
THERMODYNAMICS OF POLYMERlZATl ON. V. 317

Consider now a copolymer of composition a. It will have a value


of AG equal to AGa and negative. Therefore this copolymer is stable
relative to pure homopolymers Nil and M,. The same is true for
copolymer b. Consider copolymer x, however. It might have a
negative value of AG but we have to question whether o r not AG
could be lower by dissociating the copolymer. If it dissociates into
two copolymers of different compositions, the lowest stable values
of AG which can be attained by the decomposition is AG at composi-
tion y and z, respectively. These copolymers a r e in equilibrium
with one another since component M, would have the same chemical
potential in both. This is because for S l < 0, (H12 + HZl) < (Hll + H,,)
and MI-M, and M,-M, pairs are energetically favored over MI-M,
pairs. Hence at low enough temperatures, two copolymers having
different compositions will be formed.
It is important to note that compositions y and z are not neces-
sarily associated with the minima of the AG vs X, curve. The free
energy changes at composition y and z may be different from the
minimum values, since the important criterion for equilibrium is
that the chemical potential of a given component is the same in both
copolymers, not that the free energy change of each copolymer be
a minimum. Thus copolymers to the left and the right of y and z,
respectively, are stable, whereas copolymers of composition be-
tween y and z a r e unstable and will dissociate into two different
types of copolymers.
In connection with Fig. 14 it will be observed that two inflection
points exist between y and z. Let us calculate the maximum critical
temperature To where decomposition will occur and the composition
of the copolymer Xg, at this temperature. These points are shown in
Fig. 13. Above the temperature To the curve of AG vs X, is every-
where concave downward. Below this temperature two minima will
occur, as already discussed, and there will be two inflection points
where a2AG/aXz = 0.
A s the temperature is raised toward To,the minima will move
closer together in composition and as a result so will the inflection
points. At To both minima and the inflection points will coincide at
the same temperature. Thus To is the temperature where both
a2AG/aXE and a3AG/aXl equal zero at the critical composition X i [22].

a3AG/aX,3 = -RTo$(l - Z$X,O)/[Xi(l - $Xi)]' =0 (54)


318 SAWADA

Hence we find

xg = 1/2# (56)

When x = 0 o r x < 0, 0 must be negative in order for To to be


positive. A negative value of 0 indicates a repulsive interaction
between unlike components, and block copolymer will occur “71.
Thus, from Eq. (55), the larger this repulsive interaction, the
higher the temperature will be where decomposition begins. A s a
generalization we might say that virtually all systems with nega-
tive f2 and x should give two copolymers of different compositions.
Thermodynamically, of course, copolymers of composition in the
region y < X, < z a r e unstable at this temperature, and two copoly-
mers of different compositions y and z will be favored. This has not
been found experimentally. The reason for this lack of agreement
with theory is related to the fact that unless 51 is a sufficiently large
negative value, To will be so low that the kinetics of the decomposi-
tion in the copolymer will be infinitesimally slow, and thus a copoly-
mer will be generally found which is actually metastable.
More recently, Harvey and Leonard [23] have considered the com-
plete reversibility in equilibrium copolymerization in solution, tak-
ing into account nonideal behavior of the solution through the use of
Flory’s expression for the partial molar free energy of mixing of
monomers and copolymer [24]. The value of P,, is set equal to
K exp(-AG,,$RT), where K is a numerical constant found to be 1
and AG,,, is the free energy change upon the addition in solution
under conditions prevailing at equilibrium of 1 mole of monomer
M, to copolymer chains in order to form MI-M, bonds. Specific
equations for equilibrium bulk copolymerization were derived in
the presence of monomers a s the solvent of copolymers.

E, Equilibrium Monomer Concentration

Izu and O’Driscoll [ l l ] treated the system where the monomers


and the copolymer coexist. Since the equilibrium state will occur
near the ceiling temperature in the case of addition copolymeriza-
tion, it is desirable to develop a theory giving an equation for the
equilibrium monomer concentrations.
THERMODY NAMl CS OF POLYMER IZATl ON. V. 319

The possible reactions and associated rate constants are as


follows :

dM:+Ml - k‘, wM,-M:


kl,

k,2
k’i w M 1 - W
-M:
- WM,-M:
““‘w Ml k,?
+ M2-
+
k21

I
k22
.“.% + M Z X mM2-M:
This model is often referred to as the diad model. The constants
kll, k12, kl, and $, are the rate constants of the propagation steps;
k c , kfi, k z , and k, are the rate constants of the depolymerization
steps. It is assumed that the constants are independent of the last
number of the chain.
The following probability parameters were defined [25] to ex-
press the relative amount of various chain ends:

([Mil), =a (57)

([MiIn-,/[M,In)t = E (58)

( [ ~ z I n - l / [ ~ z l nt )= 77 (59)

where ([M,]), is the probability of finding M, monomer unit at the


chain end, ([Ml]n-l/[Ml]n)t is the conditional probability of finding
M, monomer unit at the penultimate unit given that the chain end is
occupied by M, monomer unit, and ([M,]n,/[M,]n)t is the conditional
probability defined analogously.
The following relationships are obtained.

([MZI), =1-a (60)

([MzIn-1/[MJn)t = 1- 6 (61)

( [ ~ 1 l n - l / [ ~ , l n ) t= 1 - 77 (62)
According to the principle of detailed balancing at equilibrium, we
find for a chain of infinite length:

k,,[M,],a - k,aE =0 (63)


320 SAWADA

klz[Mzlea - kid1 - a)(l - q) = 0 (64)


k,,[M,],(l - a) - k,a(l - E ) = 0 (65)

&z[M,Ie(1 - a) - k& - a)s = 0 (66)

where [MI], and [M,], are the monomer concentrations at equilibrium.


Since we have four independent equations, we can determine all
the variables if a condition which gives another equation between
variables is specified. F r o m Eqs. (63)-(66) we see that

where the equilibrium constants are K, = kll/kfi, 6 = kl,/kE,


K3 = k&,, and & = kzz/kE.
Assuming that there is no side-reaction in the system, we obtain
the following constraints which state that the monomer structure is
conserved in the system.

[MI], + X,{[MIo .- ([Mile + [MzIe)) = fl[MIo (70)

e xzff ~
[~zl+ 1 -0 ([Mile + [ ~ z l e ) }= f J ~ 1 0 (71)

The Monomer State + The Polymer State = The Total (72)

where [MIo is the total concentration of both monomer units includ-


ing both the monomer and polymer states, X, and X, a r e the mole
fractions of M, and M, monomer units in the copolymer, and f, and
f, a r e the mole fractions of M, and M, monomer units in the mono-
mer mixture. When the conditional probability parameters at the
chain end are equal to those on the main chain, it is possible to ex-
press the copolymer composition and diad fractions in t e r m s of E
and q.
THERMODYNAMICS OF POLYMERIZATION. V. 321

where Fll, F12,FZ2,and F,, a r e diad fractions.


A l l eight rate constants are not necessary for solving the equa-
tions; the sufficient values are K,, &, and (K&)/(Kl&) which are
expressed in t e rm s of the standard free energy changes of the ele-
mentary steps:

w = *(AG:, + AGX, - AG;, - AG;,) (81)

where AG;,, AG;,, AG;,, and AG& are the standard free energy
changes of the corresponding elementary steps.
Equation (80) is identical in form to Eq. (40),with 2w being
identified with AG,,, which represents the standard free energy of
formation of a heterogeneous bond from homogeneous bonds. Thus
we can predict theoretically the equilibrium state of an addition co-
polymerization which obeys the diad model if [ML,
x,, K,, &, and
w are given. For the case of the critical condition beyond which the
monomer mixture fails to be polymerized, we know that

[MI], + [ ~ z l e= [MI, (82)

Thus, if we specify the value of [MI,, the values of [M,], and [M,], at
the critical condition can be calculated from Eqs. (68) and (82).
Modified phase diagrams a r e calculated using Eqs. (67), (73),
and (74) from the values of [M,], and [M,],. Izu and O'Driscoll [ll],
using such an approach, have calculated the monomer and polymer
curves for some hypothetical cases and obtained the results shown
in Figs. 15, 16, and 17. In these figures a r e shown the results of
calculations for the cases where there is 1) no interaction, 2) posi-
tive interaction, and 3) negative interaction between unlike monomer
units. The monomer curves represent the condition above which the
322 SAWADA

POLYMER

I I I I
0 0.2 0.4 0.6 0.8 1.0

MOLE FRACTION MI
Fig. 15. Equilibrium copolymer and comonomer compositions computed as a func-
tion of temperature for [MI, = 1.0, AS;, = AS:z = -28 eu, AH;, = -10 kcallmole,
AH& = -16 kcal/mole, w = 0 [ 111.

350 k

h
I
I
MONOMER

Y
v

300
3
+-Q
LI
w POLYMER
a
w
I-
250 t

MOLE F R A C T I O N Mi
Fig. 16. Equilibrium copolymer and comonomer compositions computed as a func-
tion of temperature for [MI, = 1.0, AS:, = AS"z = -28 eu, AH;, = -15 kcal/mole,
AH& = -16 kcal/mole, w = -1.5 kcal/mole [ 111.
THERMODYNAMICS OF POLYMERIZATION. V. 323

3 50

Y
v

W
5 300
t-
Q
[L
W
a
I
W
k
I POLYMER I
250
0 0.2 0.4 0.6 0.8 1.0

MOLE FRACTION Mi
Fig. 17. Equilibrium copolymer and comonomer compositions computed as a
function of temperature for [MI, = 1.0, AS:, = AS& = -28 eu, AH;, = -17
kcal/mole, AH& = -16 kcal/mole, w = 1.5 k c a h o l e [ 111.

monomer state is more stable. The polymer curves represent the


copolymer composition which is expected to be formed reversibly
at the critical condition. It is observed from Figs. 16 and 17 that
either maximum o r minimum critical temperature for the mono-
mer mixture is observed when there is a positive or negative inter-
action, respectively .
F. Penultimate Unit Effect
The behavior of some comonomer systems indicates that the re-
activity of the propagating species is affected by the next-to-last o r
penultimate monomer unit. This effect is often referred to as the
penultimate unit effect. Silberberg and Simha [26] have discussed
the kinetics of reversible processes on linear lattices that would
include copolymerization with penultimate unit effect. O’Driscoll
et al. [ll, 271 have used a kinetically based approach to treat re-
ve r sible chain- growth copolymerization with penultimate unit eff ect ,
which will be discussed in a later section.
Simha and Zimmerman [28] have enumerated grouping and se-
quences in a binary copolymer for which the penultimate unit effect
is considered. There are eight conditional probabilities which de-
scribe the sequence length distribution of the chain, and those are
324 SAWADA

interrelated by five equations. The probabilities a r e of the kind Pljk,


which may be defined a s the probability that an i - j pair is followed
by a k where i, j , and k can be either 1 o r 2. There are four equa-
tions of the type
Pijl + pijz 1 = (83)
In addition, we find

Theil [29] has extended the thermodynamic treatment of the penul-


timate effect to the case of the infinite molecular weight binary co-
polymers where the penultimate effect may he operative. Together
with Eq. (84), Theil [29] has suggested that four equations expressed
in terms of four conditional probabilities define the equilibrium
state.

In cases where the nearest-neighbor approach is sufficient, the


following will be true in the present treatment. Pijk will reduce to
P,, and thus P j j l + Pijz = 1 and Pi,, + Pjj2= 1 where i, j , k = 1 o r 2.
Also & i j k will then equal A E , ~and
~ , klj,will vanish. In such a case
Eq. (85) will reduce to the equilibrium expression for the nearest
neighbor case.

~-
p,,p,, - exp{2AG:,/RT}
P1,PZl

where AG;, is the average free energy of formation of an M,-M,


and M,-M, bond. Equations (86) and (87) become identities with
zero in this more special case. Equation (88) is identical in form to
Eq. (40), with 2AG:, being identified with AG,,.
THERMODYNAMICS OF POLYMERIZATION. V. 325

11. DEGREE OF POLYMERIZATION AND COPOLYMER


COMPOSITION OF BINARY COPOLYMERIZATION SYSTEM
A. Degree of Polymerization
The equations derived by Tobolsky [30] for the various cases of
equilibrium copolymerization may well serve a s a simple means of
predicting degree of copolymerization. Tobolsky and Owen [30]
neglected the interaction between monomer units in the polymer
chain so that the equilibrium sequence distribution of these models
should always be random.
The concentrations of polymer chains containing n monomer units
of any distribution o r proportion of MI and M, is designated a s C,.

K;
M, M,* (89)

-
Equation (82) applies for n = 1 a, The total equilibrium concen-
tration of copolymer molecules N is given by
326 SAWADA

The total equilibrium concentration of monomer segments W incor-


porated in the copolymer is found to be

+...

The number-average degree of copolymerization at equilibrium is


given by

P = W/N

The amount of monomer MI entering the copolymer is governed


by
Conc MI entering copolymer
Conc M, and M, entering copolymer

Therefore
THERMODYNAMICS OF POLYMERIZATION. V. 327

Similarly

F r o m Eq. (95) and by analogy with previous results for equilib-


rium homopolymerization, a transition tsmperature (floor tempera-
ture) will be encountered where

K,[M,], + &[MzI, = 1 (103)


Below the transition temperature the equilibrium concentrations
[M,], and [M,], are equal to [M,], and [M&. The degree of polym-
erization of the copolymer is

P = 1/[1 - (Kl[MIIo + %[M,I,)I (104)

Above the transition temperature, the degree of polymerization is


large and the following approximation is valid:

K ~ C M ~+I %~ E ~ z l e = 1 (105)

From Eqs. (97) and (98) we find

([MI], - [M,Ie)/([MzI, - [MZIe) = K1[M,Ie/Kz[MzIe (106)

A t any temperature above the transition temperature, [MI], and


[M,], can be computed from Eqs. (105) and (106). By rearranging
Eq. (99) we obtain the following expression for calculating P above
the transition temperature:
328 SAWADA

Results of calculations were tested experimentally using the co-


polymerization system of sulfur and selenium [30]. The following
nomenclature will be used in the derivations:

M, = S, monomer unit, M, = Se, monomer unit, M: = S, diradical,


and
M: = Se, diradical.
Calculations were made for an initial monomer composition of
90% sulfur and 10% selenium by weight, using Eqs. (103) through
Eq. (107), and the calculated values of P a r e plotted in Fig. 18 to-
gether with the experimental results obtained by Schenk [31]. Ex-
perimental values agree well with the theoretical results.

1 I I 1
350 400 4 50 500 550

TEMPERATURE ( O K )

Fig. 18. Degree of polymerization vs copolymerization temperature for sulfur-


selenium copolymers in the case of an initial monomer composition of 90%sulfur
and 10%selenium by weight. Solid line, experimental; dashed line, calculated [ 311.

Kang and O’Driscoll [32] have studied the effect of copolymeriza-


tion reversibility on molecular weights by calculating leading
moments by the generating function technique. Results of calcula-
tions of average molecular weights from the resulting equations
were tested experimentally using the radical copolymerization sys-
tem of styrene and a-methylstyrene at 60 and 100°C by gel permea-
THE RMODY NAMl CS OF POLY ME R IZATION. V. 329

tion chromatography or membrane osmometry. The effect of co-


polymerization reversibility on molecular weights mainly occurs
'indirectly," i.e., from an increasing termination rate constant, as
a-methylstyrene feed mole percent increases. The 'direct effect,"
i.e., decrease of net rate of polymerization from the participation
of depropagation steps, will be observed in the temperature range
higher than 100°C. This implies that reversibility need only be con-
sidered for the homopropagation reaction of cu-methylstyrene in the
temperature range 60 to 100°C.
B. Copolymer Composition Equation
We will now discuss the copolymer composition equation in the
presence of depolymerization reactions. There a r e two principal
approaches to the interpretation of the composition of equilibrium
copolymerization. The first approach is to consider the kinetics
of copolymerization with depropagation. The second approach is to
relate the composition of copolymer to the probability of producing
a certain sequence of monomers, and the approach used is to con-
sider the relative probabilities of the various propagation and de-
propagation steps.
1. Kinetic Approach. The binary copolymerization is generally
considered in terms of four propagation steps involved in the forma-
tion of a chain, and these a r e irreversible:

and the copolymer composition is then given by

Equation (108) is known a s the copolymerization equation o r the co-


polymer composition equation. The copolymer composition, d[M,]/
d[M,], is the molar ratio of the two monomer units in the copolymer.
The parameters of copolymerization r, and r2 a r e termed the mono-
mer reactivity ratios, and a r e given by the ratio of the rate con-
stants of the homopolymerization (kll and h2)to the rate constants
of the cross-propagation (klz and k,,).
330 SAWADA

Ham [33] has generalized the original penultimate unit treatment


of Mertz, Alfrey, and Goldfinger [34], and has found satisfactory
agreement with the data in certain systems a t given temperature.
Ham’s approach 1331 emphasizes the kinetic nature of the chain
growth and the impossibility of particular reactions.
In contrast to Ham’s approach, Lowry [35] has treated the devia-
tions from the copolymer composition equation from the point of
view of the reversibility, and has derived copolymer composition
formulas on the basis of three different sets of assumptions (I, II,
and m)about the depropagation ability of different radical struc-
tures, and Hazel1 and Ivin [36] added a fourth set, Yamashita e t al.
[37] added a fifth set, and Wittmer [38] added a sixth set. These
assumptions a r e summarized in Table 1 [39].
Table 1
Copolymerization Mechanisms with Depropagation

Terminal structureapb
Case 111* 211* 121* 221* 222* 122* 212* 112* Ref.

I + + 35,39
I1 + 35,39
I11 + + 35,39
IV + + + + 36,39
v + + + + 31
v I + + 38

alll*denotes-M,M,M, *.
+ denotes that depropagation is assumed to occur.
The copolymer composition for Case I is given by

with (Y defined by
THERMODYNAMICS OF POLYMERIZATION. V. 33 1

The qualitative predictions of Case I a r e as follows: 1) Tempera-


ture will have a marked effect on the shape of the copolymer compo-
sition curve in the region of the ceiling temperature of M,, 2) dilu-
tion of the monomers by an inert solvent will have a marked effect
on the shape of the copolymer composition curve in the region of
the ceiling temperature of M,, and 3) the limiting mole fraction of
M, in a copolymer prepared at high temperature will be one-half.
These effects a r e shown by curves in Figs. 19 and 20 for the hypo-
thetical copolymer system assumed below: (a) rl = r, = 1 for all
temperatures, (b) ceiling temperature = 50°C, (c) AE* for k,, =
2 kcal/mole, and (d) AE* for kE = 1 2 kcal/mole.
The copolymer composition for Case I1 is given by

where a is defined by Eq. (110) and v by

U =
K[M,l + K[M,]/r, -a
K[M,I

where K is the equilibrium constant for the equilibrium expressed


in the form

-M,M,M,M,+ mM,M,M? + M,
This leads to the same qualitative predictions as before but the
limiting mole fraction of M, in the copolymer prepared at high tem-
perature is two-thirds rather than one-half.
The qualitative predictions of Case III a r e the same as those of
the previous two. The quantitative differences between the second
and third cases involve so many independent parameters that it
may not be possible t o choose between the two by copolymerization
experiments alone.
The results which support the assumption of Lowry were ob-
tained by several workers [40-421.
O'Driscoll and Gasparro [40] have studied the copolymerizations
of styrene (M,) with a-methylstyrene (M,) and methyl methacrylate
(M,) and of acrylonitrile (MI) with a-methylstyrene (M,) in the ap-
proximate regions of the ceiling temperatures of a-methylstyrene
and methyl methacrylate. In all cases the copolymerizations were
carried out over a'wide range of monomer feeds, and Lowry)s
Case I1 was found to be quite adequate in describing the results.
332 SAWADA

0 0.2 0.4 0.6 0.8 1.0

mole fraction M2 in monomer mixture

Fig. 19. Effect of temperature on copolymer composition at [MI] + [M2I = 1:


(1) 0°K;(2) 0°C; (3) 50°C (T,.); and (4) 100°C [35].

L
1.0
0,

-E,
0 0.8
a
0
U

.-C 0.6

s C
0.4
.-c
0
U

L
.4-
0.2

-0
0 n
E " 0 0.2 0.4 0.6 0.8 1.0

mole fraction M2 in monomer mixture

Fig. 20. Effect of [ M, ] + [ M, ] on copolymer composition at the ceiling temperature:


(1) [MI] + [M, ] = 10; (2) [M,] + [M,] = 1.0;and (3) [MII + [M,I = 0.1 1351.
THERMODYNAMICS OF POLYMERIZATION. V. 333

The influence of reversal of propagation reactions on copolymer


composition has been studied for the following systems [41]: 1) an-
ionic copolymerization of vinyl mesitylene (M,) and a-methylstyrene
(M,) in tetrahydrofuran, and 2) radical copolymerization of styrene
(M,) and methyl methacrylate (M,) in o-dichlorobenzene. F o r these
systems Lowry's Case I1 provides a reasonably satisfactory inter-
pretation of the variation of copolymer composition with [M,] at
constant feed composition.
Lowry's Case I was studied by the cationic copolymerization of
styrene (M,) with a-methylstyrene (M,) as the simplest mechanism
of depropagation [42]. The equilibrium constant of M, homopolym-
erization thus obtained is much smaller than the literature value.
The deviation from the literature value becomes larger the higher
the polymerization temperature. The equilibrium constant is de-
pendent on the sequence length of a-methylstyrene and becomes
larger with an increase in the sequence length.
On the other hand, it has been suggested by Johnston and Rudin
[43] that the simple copolymer equation, Eq. (108), was clearly an
adequate model for the radical copolymerization of a-methylstyrene
with styrene and with methacrylonitrile at 60"C, although the ceiling
temperature of a-methylstyrene is 61°C [43]. Depropagation effects
were not important in these copolymerizations because the mean
sequence lengths of a-methylstyrene in the copolymers were short.
Such short sequences have higher polymerization enthalpies and
hence higher ceiling temperatures than higher molecular weight
homopolymers. The radical terpolymerization of methacrylonitrile,
styrene, and a-methylstyrene was also studied at 60°C in toluene
solution [44]. Again the behavior of a-methylstyrene can be de-
scribed by a simple model without reference to ceiling temperature
o r penultimate effects because the sequence lengths of this monomer
in the terpolymer a r e short. Kang and O'Driscoll [45] interpreted
the physical significance of reactivity ratios obtained by applying
the simple copolymer equation to composition data from systems
which a r e expected to polymerize with some depropagation. Thus
it is possible to represent composition data for a reversible co-
polymerization by Eq. (108) and to put a correct physical interpre-
tation on the parameters derived from that equation.
The radical copolymerization of styrene (M,) with a-methyl-
styrene (M,) in the temperature range from 60 to 150°C can be de-
scribed with the assumption that addition steps of a-methylstyrene
may be reversible [46]. For the temperature range from 60 to 110°C
it could be demonstrated that it would be sufficient to consider only
334 SAWADA

the reversibility of the addition steps of a-methylstyrene to a radical


end with a-methylstyrene as the terminal unit (Lowry’s Case I). A t
a reaction temperatures of 150°C the addition of a-methylstyrene
to a styryl chain end is also reversible (Lowry’s Case IV).
Olefins and sulfur dioxide copolymerized in the liquid phase under
the influence of radicals to form 1 : l copolymers as olefin polysul-
fones. A good example of Case IV is provided by the system cyclo-
pentene-isobutene-SO, which gives a regular 1:1 olefin-SO, copoly-
mer [36].
For the cationic copolymerizations of 3,3-bischloromethyl oxacy-
clobutane (M,) and tetrahydrofuran (M,), Yamashita et al. [37] have
considered two models in relation to the known mechanism. Model 1,
which corresponds to Case IV in Table 1, assumes a reversible addi-
tion of M, monomer at both M, and M, ends, and Model 2, which cor-
responds to Case V in Table 1, assumes reversible addition of both
M, and M, monomers at the M, end only. In both models the rates
of propagation and depropagation reactions are assumed to be inde-
pendent of the composition of the residual polymer chain. With the
steady-state assumption applied to the concentration of living ends,
a copolymer composition equation is derived by Lowry’s procedure.
Yamashita et al. [37] came to the conclusion that Model 2 expresses
the mechanism of copolymerization rather better than Model 1. In a
copolymerization which includes a comonomer with a ceiling tem-
perature near room temperatures such as tetrahydrofuran, the de-
propagation reaction considerably influences the copolymer composi-
tion, especially at the higher feed ratios of this monomer. In the
case of the depropagation of unsaturated compounds, it is sufficient
to remove the last unit to reconstruct an active center. For hetero-
cyclic compounds, however, it is necessary to consider the removal
of the last unit and cyclization of the penultimate one. Penczek [47]
derived equations for these compounds based on this model.
The copolymer composition equation with reversibility of all
propagation reactions was derived [38, 481. The reaction scheme
is the previously described diad model. To treat the problem quan-
titatively, let us define two new parameters, q, and q, [38]:

The rate constants kE and kE can be replaced by the equilibrium


constants K, and K, according to K, = kE /kll and K, = kZ&.
The copolymer composition is given by
THERMODYNAMICS OF POLYMERIZATION. V. 335

where x, is the mole fraction of the reactive chains ending with


monomer M, with a sequential length 1.

and similarly

Knowing the constants and the monomer concentrations in Eqs. (115)


and (1161, we can calculate the numerical values for x, and y,. This
can be done graphically, but it is more practical to use a computor.
Using Eq. (1141, we can then calculate the composition of the poly-
mer formed.
F o r the special case where K, = 0, q, = 0, and q, = 0, we can write
the simpler equation

1 - x , = r,([Ml] + K,) + [M,] - , / r ( [ M l l + K,) + [M,]


2r,K, 2r1K1
(118)
This case corresponds to Case ZI in Table 1, where only the homo-
polymerization of the monomer M, is reversible.
336 SAWADA

F o r the special case where K, f 0, q, = 0, and q, = 0, we know that

x, is given by Eq. (118), and y, by Eq. (120):

This case corresponds to Case VI in Table 1, where the two homo-


polymerization reactions a r e reversible and the alternative steps
a r e irreversible,
If the equilibrium constants depend on the length of the monomer
sequence (penultimate effect), further changes must be introduced
into the equations. Wittmer [38] considered the simple case where
the sequence of two monomer units of M, does not depolymerize,
and longer sequences of M, a r e subject to equilibria with a constant
K,. And for the special case where K,, q,, and q, a r e zero, we find

d[M,l/d[M,l = (121)
1 + r,[M,l/[M,l

u corresponds to Eq. (118):

(J=-1 r,([M,l + K,) + [M,]


2 r,K,
(122)
Equation (117) can be applied to the radical copolymerization of
a-methylstyrene (M,)-methyl methacrylate (M,) at 20 to 100°C.
Between 100 and 150°C the reversibilities of the homopolymeriza-
tion step of methyl methacrylate and of the alternating steps had to
be considered, i.e., Eq. (114) in connection with Eqs. (115) and (116).
THERMODYNAMICS OF POLYMERIZATION. V. 337

Equation (122) can be applied to the system a. methylstyrene


(M,)-acrylonitrile (M,) where the sequence of two monomer units
of a-methylstyrene is stable and does not depolymerize. The re-
versibility of the polymerizations of a-methylstyrene and methyl
methacrylate can be explained by sterically induced strain in the
polymer chain [35]. In the copolymer a-methylstyrene-methyl
methacrylate this strain involves the whole polymer chains whereas
in the a-methylstyrene-acrylonitrile system the strain is broken by
the acrylonitrile sequences and is built up again by a-methylstyrene.
This explains the differences in the depolymerization tendencies of
sequences of two units of a-methylstyrene and longer sequences in
this system. The same idea might be applied to the results of John-
ston and Rudin [43].
2. Probability Approach. In contrast to the kinetic approach,
O’Driscoll and his co-workers [25, 27, 491 have treated the deriva-
tions of the copolymer composition equation using a probability ap-
proach. The copolymer composition equation for the diad model is
given by [25]

where a is the transient probability that a chain ends in a [M,] unit


and b is the transient probability that a chain ends in a [M,] unit,
and q and E represent the transient probabilities of Eqs. (52) and (53),
respectively.
The following relations must be considered:

a k , , + (1 - ~)h1
= bb,, + (1 - 17)k~l (125)

bE(1 - V)k, = a[E(k,, + kz + k,,) - (k,, + e2k;)] (126)

aq(1 - E)kc = b[~)l(k,,+ k z + k,,) - (k2, + V%z)1 (127)

Equations (124)-(127) a r e the set necessary to solve for a, b, E, and


q in t e rm s of the eight kinetic rate constants.
By letting k z = kE = k z = &, = 0, it is easy to demonstrate that
Eq. (123) reduces to the simple copolymer equation, Eq. (108). Equa-
tion (123) can also be shown to reduce to Lowry’s Case I which, in
t er m of O’Driscoll’s model [25], occurs when k z = k z = k z = 0. The
338 SAWADA

next three equations give the equivalence between O’Driscoll’s nota-


tion and Lowry’s:

k,, /kZI = [Mzlrz /[MI1 (128)

kzz/b = [MZIP (129)

q=(Y (130)

Using the notation of the right-hand side of Eqs. (128)-(130) and


Eq. (127), we find

azrz/([MIIP) - d l + rz[MzI/[MII + rz/([MIIP)I + rz[MzI/[M,I = 0


(131)
Upon rearrangement,

- a(1 + [MZIP + [M,IP/rJ + [MJP = 0 (132)

which is identical to Eq. (4) in Lowry’s paper [35].


Furthermore, Eq. (123) becomes

d[M,I/d[MJ = ~ ( +1rJ/[r2 - (arZ/P)I (133)

which is identical t o Eq. (12) in Lowry’s paper [35]. Equations (125)


and (126) also simplify to

and

The Aentity of results w-th respec- -3 composition in O’Driscoll’s


treatment and Lowry’s is consistent with expectation. Further, the
probability approach has yielded new sequence distribution equations.
The mathematical treatment of the penultimate effect in a binary
copolymerization involves the use of the sixteen r a t e constants, for
there a r e eight reversible reactions to consider:

wMIMf +M , -
kl
k, ,W M,M,M:

-M,M:+MZ - 1;,
k3
-M,M,M,*
THERMODYNAMICS OF POLYMERIZATION. V. 339

In a manner similar to that used in deriving Eq. (123), the co-


polymer composition equation for this model, often referred to a s
the triad model, is given by [25]

In this equation, four transient probabilities for antipenultimate


units a r e given by

The following relations must be considered:


340 SAWADA

The foregoing equations can be shown to reduce to Lowry's Case I1


i f we assume the following: k, = k4 = k, = k, = k,, = k12,=k14 = 0,
k, = k9, k, = kll, k, = k12, and k, = k15. Equation (136) is the most
general copolymer composition equation which has ever been de-
rived [25].
Computor simulation of typical copolymerizations with depropa -
gation by means of a Monte Carlo technique were made on hypo-
thetical cases [ 271. The composition-feed relationships for copolym-
erization of the monomer pairs having those hypothetical values of
thermodynamic and kinetic quantities shown in Table 2 were simu-
lated a t various temperatures. It should be noted that the kinetic
quantity is dependent on a kinetic mechanism whereas the thermo-
dynamic quantity is independent of it.
Cases A to D correspond to copolymerizations with monomer
pairs having the same ceiling temperature. In Cases A-1 and A-2
we assume that all the elementary steps have the same thermody-
Table 2
Typical Cases Assumed for Computation [ 271

A-1 A-2 B C D E-1 E-2 E-3

-AH;, , kcal/mole 15.0 15.0 15.0 15.0 20.0 20.0 20.0 20.0
-AHy2, kcal/mole 15.0 15.0 12.0 20.0 10.0 15.0 20.0 10.0
-AHil, kcal/mole 15.0 15.0 18.0 20.0 10.0 15.0 10.0 20.0
-AH;, ,kcal/mole 15.0 15.0 15.0 15.0 20.0 10.0 10.0 10.0
-AS", cal/(deg)(mole) 28.0 28.0 28.0 28.0 28.0 28.0 28.0 28.0
11 = k,l/kl2 1 0.1 1 1 1 1 1 1
12 = k22/klI 1 10.0 1 1 1 1 1 1
Sl = k,,/k21 1 1 1 1 1 1 1 1
THERMODYNAMICS 0 F POLYMER I ZATl ON. V. 34 1

namic quantities and that the values of rl and rz a r e different. The


results for Case A-1 in Fig. 21 show the polymer composition is
identical to that of the feed at 100 and 200°C. It is impossible to ob-
tain high polymer at 300°C.
L

-: 1.0
0
(I

3! 0.8
.-C
2 0.6

0.4
z
*-

u 0.2
E
0
0 0.2 0.4 0.6 0.8 1

mole fraction M2 in monomer mixture

Fig. 21. Results of Monte-Carlo calculations of copolymer composition curves for


Cases A-1 and A-2 of Table 2: (1) A-2, 100°C; (2) A-2,2OO0C;(3) A-2, 250°C; and
(4) ( 0 ) A-1,lOO"C; ( 0 ) A-l,20OoC [ 271.

The curve at 100°C in Fig. 21 for Case A-2 is similar to that ob-
tained by plotting Eq. (108) with the same rl and r, because very few
depropagation reactions occurred. A s the temperature rises, it be-
comes similar to Case A-1, which shows that thermodynamic control
dominates the kinetic control. Case B, shown in Fig. 22, is different
from Case A-1 only in the values of -AH;z and -AH&; the sum of
terms is 30 kcal/mole in both cases. In Fig. 22 this did not make
Case B different from Case A-1 at the lower temperature (100°C).
A t the higher temperature (200°C) the curve appears similar to a
curve with rl < 1.0 and rz > 1.0 in Eq. (108). This again shows that
thermodynamic factors controlled the reaction at the higher tem-
perature.
When both of -AHlz and -AHzl are less than the heat of homo-
polymerization (Case D, Fig. 221, the curve approaches one which
corresponds to the case where both rl and r, a r e larger than unity
in Eq. (108), and thus the copolymer formed under this condition is
similar to a block copolymer. When both -AH,, and -AH,, are
larger than the heat of homopolymerization (Case C, Fig. 22), the
curves resembles one obeying Eq. (108) with rl and r, having values
342 SAWADA

.-C
EN o.6
C

0.4
U
2
c

mole fraction M, i n monomer mixture

Fig. 22. Results of Monte-Carlo calculations of copolymer composition curves for


Cases B, C, and D of Table 2: (1) D, 300°C; (2) C, 300°C; (3) B, 100°C; C + D, 200°C;
and (4) B, 200°C [ 271.

smaller than unity. The deviations from Eq. (108) at the higher
temperature are due to the occurrence of the depropagation reac-
tion and are apparent in Figs. 21 and 22.
The general trend of the behavior of the composition-feed curve
in Cases E-1 to E-3 i s that the shape of the curves become flatter
with increasing temperature (Fig. 23).

,5 0.6
zN
s
.- 0.4

--
c
u
2
0.2
0
E
0
0 0.2 0.4 0.6 0.8 1.(

mole fraction Mzin monomer mixture

Fig. 23. Results of Monte-Carlo calculations of copolymer composition for Case E-1:
(1) 100°C; (2) 200°C; and (3) 300°C [27].
THE RMODY NAMl CS 0 F POLYMER I ZATION. V. 343

O'Driscoll and his associates [49] reported the effect of reaction


temperature on the copolymer composition curve in the radical co-
polymerization of a-methylstyrene and methyl methacrylate, which
have relatively low ceiling temperatures. The diad model was
adopted a s a simple explanation. The relative values of the eight
rate constants of the diad model a s estimated by the seven constants
are

The values of K, and & are available from the thermodynamic


data of homopolymerization by using the relation

-RT In K, = AG; = AH: - TASY (149)

Since no data a r e available on K , and &, Eq. (150) has been assumed
throughout the calculations:

AG, - AG:
O = 1/2(AG," + AG:) (150)

The compositions of the polymer is expressed by Eq. (151) a s


parameters 17 and E :

= (1 -M 1 -4 (151)

The values of q and E a r e determined by solving Eqs. (124)-(127)


simultaneously. Since these equations a r e nonlinear, we can solve
them by a numerical method. The solid lines in Fig. 24 show the
theoretical prediction obtained by this procedure. The theoretical
curves fit well to the experimental points. Since the temperature
range and its effect a r e quite large, the diad model appears to be
quite adequate for accurately describing the reversible copolymer-
ization of this morlomer pair.
344 SAWADA

0 0.2 0.4 0.6 0.8 1.0

mole fraction M2 in monomer mixture

Fig. 24. Dependence of temperature and monomer feed composition on the copolymer
composition in the radical copolymerization of MMA(M, -a - MS (M, ): ( X ) 60°C;( 0 )
114°C;( 0 ) 147°C;and (-) theoretical prediction [ 491.

III. MULTICOMPONENT COPOLYMERIZATION


A. Heat of Terpolymerization
Terpolymerization, the simultaneous polymerization of three
monomers, has become increasingly important from a technological
point of view. Since no data are available for heats of terpolymeri-
zation systems and such data are more difficult to obtain than data
of binary systems, it is desirable to develop a method for the pre-
diction of heat of terpolymerization systems from binary data [50].
The quantitative treatment of terpolymerization is quite complex
since nine propagation reactions are involved.
Reaction Heat of reaction
M: + M,
M: + M,
M: + M3
M: + M,
M,* + M,
M,* + M3 H23
M: + M I H31
M: + M, H32
M: + M3 H33
THERMODYNAMICS OF POLYMERIZATION. V. 345

The expression for the molar heat of reaction of each of the propaga-
tion reactions is shown above. An expression for the heat of ter-
polymerization a t low conversion can be obtained by a treatment
similar to that used in binary copolymerization.
The molar heat of terpolymerization is given by

N' H , ,
AH, = L NH z z+ N
+A N H l z + SN H 2 ,
X H 3 3+ A
NO NO NO NO NO

N N N
+ A H l 3 + A H 3 1 + S H Z 3+ S H 3 2
NO NO NO NO

where the number of nearest-neighbor pairs of i and j is N i , , and


the total number of pairs in the copolymer is No. The number of M,
monomer units in the copolymer is N S , , and thus the total number
of M,-M, pairs is

Nll = (NdC1)Pll (153)

where X, is the mole fraction of the monomer unit M,, and P,, is
the conditional probability that the M, monomer unit will follow the
M, monomer unit in the presence of the M, monomer unit. Similarly,
we have

where Xi is the mole fraction of M i monomer unit, and Pi, is the


conditional probability that the Mi monomer unit selected at ran-
346 SAWADA

dom will be followed by the M, monomer unit. The value of Pi, may
be calculated from rate constant k by
kij[M:I[MjI
(162)
= kij[M:l[Mjl + kii[M:l[Mil + kIk[M:IIMkI

where i, j, and k have values of 1, 2, o r 3. Combining Eqs. (153)-


(161) with Eq. (152), we obtain

AH, = XlP11H1, + X2PZ2H22 + X R P 3 3 H 3 3 + XlPl2Hl2 i-


X*P2,%1

+ X1P13H13 + X3P31H31 + X2P23H23 + X3P32H32 (163)

Ham [51] has treated the problem of terpolymer composition by


making the simple assumption that the probability of producing a
certain sequence of monomers is the same as producing the exact
reverse sequence. This relationship of product probabilities is

p12p23p31 = p13p32p21 (164)

A similar concept may be expressed as [52]

X,/Xj = Pjl/Pl, (165)

Equations (164) and (165) have undergone some debate in the litera-
ture [53-561, but they a r e anecessary consequence of the Q-e scheme
and can be derived from it.
Let us now evaluate the heat of terpolymerization for monomers
which do obey the Q-e scheme. Since we a r e dealing with monomers
which obey the Q-e scheme, we find

XIPI2 = XZPZl (166)

x1p13 = x3p31 (167)

'2'23 = '3'32 (168)

Substitution of Eqs. (166), (167), and (168) into Eq. (163) yields
THERMODYNAMICS OF POLYMERIZATION. V. 347

In the present case for an infinite molecular weight terpolymer,

P1, + P,, + P13= 1


P,, + P,, + P,, =1

P3, + P3, + P3, = 1

Thus Eq. (169) can be expressed in t e r m s of the probabilities of


cross-propagation:

Upon substitution of Eqs. (166)-(168) into Eq. (173), we find

We define QIj as

QIj = (H,j + HjJ - + Hjj) (175)

Consequently, the heat of terpolymerization is given by

=H l A + HZZX, + H33X3

+ xlp12Q2,Z + x1p13Q13 + x2p23Q23 (176)

The heat of terpolymerization for the ternary system acryloni-


trile (AN = 1)-styrene (ST = 2)-methyl methacrylate (MMA = 3)
calculated from Eq. (176) is presented in Fig. 25. (Note that any
348 SAWADA

gain in AH on the part of a system is considered to be positive and


any loss negative, but this is not considered here for simplicity.)

0 0.2 0.4 0.6 0.8 1.0

mole fraction AN
Fig. 25. Heat of terpolymerization for the acrylonitrile-methyl methacrylate-styrene
system: (mole MMA)/( mole ST)= 1 [ 501.

Values in Fig. 25 a r e obtained by using [57]

rlz = 0.18, rZ1= 1.35, r13= 0.04, rZ3= 0.46

HI, = 16.74 kcal/mole, H,, = 13.42 kcal/mole,

H,, = 16.29 kcal/mole

and a,, are shown in Table 3. Ham [58] has found that Eq. (164)can
be applied to this ternary system. Thus it is possible to predict the
heat of terpolymerization in this system from binary data.
In the case of binary copolymerization, X,, aI3,aZ3, P1,,and P,,
will vanish. In this case Eq. (176)will reduce to Eq. (177), which is
identical to Eq. (13), with a,, being identified with 51.
AH = HllXl + Hz&z + PlZ~12xl (177)
THERMODYNAMICS OF POLYMERIZATION. V. 349

Table 3

Parameters Used in Prediction of Heat of Terpolymerization (kcal/mole)

Binary system Hij + Hji Hii + Hjj Qi j


AN-MMA 28.4a 30.4a -2.00
AN-ST 34.67a 32.76a +1.91
MMA-ST 32.74b 29.74b +3.00
aH. Miyama and S. Fujimoto, J. Polym. Sci., 54,s33 (1961);M.Suzuki,
H. Miyama, and S. Fujimoto, Bull. Chem. SOC., Japan, 35,57,60(1962).
bM. Suzuki, H. Miyama, and S. Fujimoto, J. Polym. Sci., 31,212(1958).

Using the same procedure as for the heat of terpolymerization


in Eq. (176),the following equations are found in the case of multi-
component copolymerization containing n monomers which obey the
Q-e scheme.

+ X,P,,a,, + X3P3,Sz,, + * * * + X,P,,a,, + * ' *

Equation (178) gives the heat of multicomponent copolymerization


with n monomers, and this equation is only applicable to multicom-
ponent systems for which copolymerization follows the Q-e scheme.
Similar expressions hold for entropy and free energy changes with
n monomers which obey the Q-e scheme.
B. General Theory of Multicomponent Copolymerization
Let us consider a terpolymerization system which consists of the
components M,, M,, and M,. The reaction which occur during equi-
librium terpolymerization process a r e shown in Eqs. (179)-(181).
350
- SAWADA

where AGIj is the characteristic molar free energy of formation for


the nearest-neighbor pairs of i and j.
From the quasi-chemical approximation [59] based on nearest-
neighbor interactions, the following three coupled equations are ob-
tained.

--%3 - eXp(- AG,,/RT)


N22N33

N321
~- - exp(- AG,,/RT)
N33N11

Since the formation o r disappearance of an i-j bond is concomitant


with the like process for a j-i bond, it is reasonable to call N,, the
number of the nearest-neighbor pairs of i and j along the chain, and
to express it as N,, = N,, and similarly AF,, = AF,,. If Eqs. (182),
(183), and (184) a r e multiplied number by number, it can be shown
that

Values of N,, a r e given by Eqs. (153)-(156), (159), and (160).


Substitution of these equations into Eq. (185) yields

pC2p23p31
= exp[-(AG,, + AG23 + AG3,)/2RT1
pllp22p33
THE RMODY NAMl CS OF POLYMERIZATION. V. 35 1

where the enthalpy changes of formation for ii, j j , i j , and j i bonds


a r e denoted by Htt, H j j , Hi,, and Hji, respectively, and the entropy
changes of such bonds by S II,S j J , SiJ, and SJI.
Equation (186) may also be extended to a general treatment of
multicomponent copolymerization. The general form of multicom-
ponent copolymerization with n monomers is obtained as follows:

i. 1

i. 1

O’Driscoll et al. [60] suggested that the reversible, multicom-


ponent copolymerization can be solved by the probability approach.
The resulting equations were compared with Monte-Carlo simula-
tions and were also tested experimentally using the terpolymerizing
system acrylonitrile-methyl methacrylate- a-methylstyrene over a
wide temperature and concentration range, The effect of reaction
temperature and total monomer feed concentration on the terpoly-
mer composition curve in the radical terpolymerization of acryloni-
trile, methyl methacrylate and a-methylstyrene was studied.

Acknowledgment
The author wishes to thank Prof. K. F. O’Driscoll for many
valuable criticisms of this manuscript and for providing reprints of
papers and an advance copy of h i s manuscript.
352 SAWADA

References

H. Sawada, J. Polym. Sci., A , 2,3095(1964).


T. Alfrey and C. Lewis, Ibid., 4,221 (1949).
R. Orr, Polymer, 2,74 (1961).
M.G.Evans, J. Gergely, and E. C. Seaman, J. Polym. Sci., 3,866(1948).
H.Sawada, Ibid., A, 3,2483 (1965).
H. N. V.Temperley, J. Res. Nut. Bur. Stand., 56,55(1956).
A. M.North and D. Richardson, Polymer, 5,227 (1964).
M.H.Theil, Macromolecules, 2,137 (1969).
R. Corradini and C. Tosi, Eur. Polym. J., 4,227 (1968).
T. Alfrey and A. V. Tobolsky, J. Polym. Sci.,38,133(1959).
M. Izu and K. F. O’Driscoll, Polym. J., 1,27 (1970).
J. C. Bevington, Radical Polymerization, Academic, New York, 1961,p. 13.
M.G. Evans, in Fibres from Synthetic Polymers (R. Hill, ed.), Elsevier,
New York, 1953,Chap. 3.
R.M. Joshi, Makromol. Chem., 66,114(1963).
H. Sawada, J. Polym. Sci.,B, 1,659 (1963).
F . S.Dainton and K. J. Ivin, Trans. Faraday Soc., 46,346(1950).
H. Sawada, J. Polym. Sci., A-1, 5, 1383 (1967).
J. C. Bevington, Radical Polymerization, Academic, New York, 1961,p. 166.
R. E. Cook, K. J. Ivin, and J. H. C’Donnell, Trans, Faraday Soc., 61,1887(
(1965).
A. M. North and D. Richardson, Polymer, 6,333(1965).
T. Tsuda and Y. Yamashita, Makromol. Chem., 86,304(1965).
H. Eyring, D. Henderson, B. J. Stover, and E. M. Eyring, Statistical Mechanics
and Dynamics, Wiley, New York, 1964,p. 394.
P. E. Harvey and J. Leonard, Macromolecules, 5,698 (1972).
P. J. Flory, Principles o f Polymer Chemistry, Cornell Univ. Press, Ithaca, New
York, 1953,p. 549.
J. Howell, M.Izu, and K. F. O’Driscoll, J. Polym. Sci.,A-1, 8,699 (1970).
A. Silberberg and R. Simha, Biopolymers, 6,479(1968).
M.Izu and K. F. O’Driscoll, J. Polym. Sci.,A-1, 8,1675(1970).
R. Simha and J. M. Zimmerman, J. Theoret Biol., 2,87 (1962).
M.H. Theil, J. Polym. Sci.,C, 31,1 (1970).
A. V. Tobolsky and G. D. T. Owen, Ibid., 59,329 (1962).
J. Schenk, Physica, 23,325 (1957).
B. K.Kang and K. F. O’Driscoll, To Be Published.
G. E. Ham, J. Polym. Sci.,45,169,177,183(1960).
E. Mertz, T. Alfrey, and G. Goldfinger, Ibid., 1,75(1946).
G. G. Lowry, Ibid., 42,463(1960).
J. E. Hazel1 and K. J. Ivin, Trans. Faraday Soc., 58,176 (1961).
Y. Yamashita, H. Kasahara, K. Suyama, and M. Okada, Makromol. Chem.,
117,242(1968).
P. Wittmer, in Multi Component Polymer Systems (N. A. J. Platzer, ed.),
American Chemical Society, Washington, D. C., 1970,p. 140.
K. J. Ivin, Pure Appl. Chem., 4,271 (1962).
THERMODYNAMICS OF POLYMERIZATION. V. 353

K. F. O’Driscoll and F. P. Gasparro, J. Macromol. Sci.-Chem., A1(4), 653


(1967).
K. J. Ivin and R. H. Spensley, Ibid., A1(4), 653 (1967).
Y. Inaki, S. Nozakura, and S. Murahashi, Ibid., A6(2), 313 (1972).
H. K. Johnston and A. Rudin, Macromolecules, 4,661, (1971); J. Paint Tech-
nol., 42,435 (197O);Ibid., 42,429 (1970).
A. Rudin, S. S. M. Chiang, H. K. Johnston, and P. D. Paulin, Can. J. Chem.,
50,1757, (1972).
B. K. Kang and K. F. O’Driscoll, J. Macromol. Sci.-Chern., A7(6), 1197
(1973).
J. P. Fischer, Macromol. Chem., 155, 211 (1972).
S. Penczek, Bull. Acad. Pol. Sci., Ser. Sci. Chim., 20,437 (1972).
A. A. Durgaryan, Vysokomol. Soedin., 8,790 (1966).
M. Izu and K. F. O’Driscoll, J. Polym. Sci., A-I, 8,1687 (1970).
H. Sawada, J. Macromol. Sci.-Chem., A7(4), 905 (1973).
G. E. Ham, J. Polym. Sci., A , 2,2735 (1964).
G. E. Ham, J. Macromol. Chem., A1(2), 403 (1966).
F. R. Mayo, J. Polym. Sci., A, 2,4207 (1964).
K. F. O’Driscoll, Ibid., B, 3,305 (1965).
P. W. Tidwell and G. A. Mortirner, Zbid.. B, 4,527 (1966).
K. F. O’Driscoll, P. W. Tidwell, and G. A. Mortirner, Ibid., B, 5,575 (1967).
H. Miyama and S. Fujimoto, Ibid., 54, s33 (1961); M. Suzuki, H. Miyama,
and S. Fujirnoto, Bull. Chem. SOC. Japan, 35, 57, 60 (1962).
G. E. Ham, in Copolymerization (G. E. Ham, ed.), Wiley, New York, 1964,
p. 42.
E. A. Guggenheirn, €’roc. Roy SOC.,A, 183,213 (1944); Trans. Faraday Soc.,
44,1007 (1948).
B. K. Kang, K. F. O’Driscoll and J. A. Howell, J. Polym. Sci., Polym., Chem.
Ed., 10,2349 (1972).
J. MACROMOL SCI.-REVS. MACROMOL. CHEM., C11(2), 257-297 (1974)

Thermodynamics of Polymerization.VI.
Thermodynamics of Copplymerization. Part 2

HIDEO SAWADA
Filter Laboratory
DaicelLtd.
Teppo-cho, Sakai, Osaka, Japan

I. RADICAL COPOLYMERIZATION 258


A. Heat of Copolymerization 258
B. Ceiling Temperature 263
C. Q-e Scheme 267
D. Substituent Effect 269
E. Effect of Polymerization Temperature 272
F. Effect of Solvent 276
II. IONIC COPOLYMERIZATION 278
A. Energetic Characteristics 278
B. Reactivity 283
C. Effect of Polymerization Temperature 284
III. OTHER COPOLYMERIZATIONS 286
A. Ring-Opening Copolymerization 286
B. Miscellaneous Copolymerizations 288
ACKNOWLEDGMENT 294
REFERENCES 294

257

Copyright O 1974 by Marcel Dekker, Inc. All Rights Reserved. Neither Ihis work nor any part may be
reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying,
microfilming, and recording, or by any information storage and retrieval system, without permission in
writing from the publisher.
258 H. SAWADA

The discussion in Part 1 [l] is largely concerned with thermo-


dynamics, and is quite general without any specification as to
whether copolymerization occurs by radical or ionic propagation.
Consider now some of the specific types of copolymerization. Thus,
much discussion will concern the free energy of the transition state
which is closely related to the mechanism and the reactivity of co-
polymerization.

I. RADICAL COPOLYMERIZATION
A. Heat of Copolymerization
As described in the previous paper, the total heat of copolymeri-
zation is given by [2]

1 - [1 - 4X2(1 - X 2 )(l -
zr : — " 221

+ X,HU + X2H22 (1)

where the X's a r e the mole fractions of each monomer in copolymer,


the r ' s are the reactivity ratios of each monomer, and the suffixes
1 and 2 denote the values of each monomer. If the values of the X's
and r ' s a r e known, measurements of the heat of copolymerization
and the heat of homopolymerization of each monomer give the value
of H12 + H 21 . However, H12 and H21, the heats of reaction for the alter-
nation reactions, are not experimentally measurable. From the fact
that the values of H12 + H21 are almost constant (Table 1), the appli-
cation of Eq. (1) to the methyl methacrylate-styrene system seems
reasonable [3]. As shown in Fig. 1, the values of AH in Table 1
agree with the full line expressing Eq. (1).
In the case of the styrene-diethyl fumarate system [4], the r e -
markable difference between r t and r 2 greatly affects the calcula-
tion of Xj and X2, and this makes the factor

1 - [1 - 4X2(1 - X 2 )(l - rtr2)>^


2(1 - r t r 2 )

in Eq. (1) zero. Thus the following equation can be obtained by a s -


suming negligible heats of cross-propagation:

AH = XXHU + X2H22 (2)


THERMODYNAMICS OF POLYMERIZATION. VI 259

• Table 1
The Values of AH and H I2 + H2I in the Methyl
Methacrylate-Styrene System at 24°Ca

AH HI2 + H2I
x, (kcal/mole) (kcal/mole)

1.00 13.19 -

0.79 14.25 32.09

0.65 15.35 33.25

0.57 15.60 32.84

0.50 15.57 31.74

0.46 15.92 32.39

0.35 16.39 32.89

0.21 16.79 34.01

0 16.55 -

av 32.74
a
Reprinted from Ref. 3 by permission of John Wiley & Sons, Inc.

Figure 2 shows the comparison between AH values derived from


Eq. (2) and values obtained by direct thermochemical experiments
[4]. It is observed that the experimental points follow the curve de-
rived from Eq. (2). It is seen that Eq. (1) must be used in the acry-
lonitrile-styrene [5] and methyl methacrylate-styrene [3] systems
in which the difference between reactivity ratios of each monomer
is small. On the other hand, in acrylonitrile-methyl methacrylate
[5, 6], vinyl acetate-methyl methacrylate [3], styrene-vinyl acetate
[3], vinyl acetate-acrylonitrile [5], and styrene-diethyl fumarate
systems [4], in which this difference is remarkable, it is not neces-
sary to use Eq. (1), but it is sufficient to use Eq. (2).
Tong and Kenyon [7] have derived Eq. (3), which is the special
case of Eq. (1) where one of the monomers does not homopolymerize
by a radical mechanism:

AH = 2X(AHCOP - H u ) + H u (3)
260 H. SAWADA

0-8 0-6 0-4 02

X,(MMA)

Fig. 1. Comparison between heat of copolymerization in methyl methacrylate-styrene


system calculated from Eq. (1) and experimental values: (o) experimental value; (—) cal-
culated value. Data from Ref. 3.

10 0-8 0-6 0-4 0-2 0

Xt (styrene)

Fig. 2. Comparison between heat of copolymerization in styrene-diethyl fumarate system


calculated from Eq. (2) and experimental values: (o) experimental value; (—) calculated
value. Reprinted by permission of John Wiley & Sons from Ref. 4.
THERMODYNAMICS OF POLYMERIZATION. VI 261

where —AH = observed heat of polymerization per mole of monomer


unit, X = mole fraction of 1,2-disubstituted monomer which does not
homopolymerize in the copolymer, — H u = heat of homopolymeriza-
tion per mole of vinyl monomer, and — AHC0P = heat of copolymeriza-
tion per mole of mixed monomer in 1:1 ratio. Equation (3) requires
AH to be linear with respect to X and equal to AHC0P at X = 0.5. The
polymerization reaction can be interpreted as proceeding in two
stages: The first stage is mainly the copolymerization of two com-
ponents in a 1:1 ratio, and the second stage is homopolymerization
of excess vinyl monomer. A summary of the heats of 1:1 copolym-
erization systems where one of the monomers is considered to be
difficult to homopolymerize is given in Table 2.
Joshi [8] has made an experimental study of heats of alternating
copolymerization systems. The heats of cross-propagation in these
copolymerizations were found to be high, exceeding in many systems
the heat of homopolymerization of either of the two monomers.
Ivin et al. [9] studied heats of copolymerization of solid olefin
polysulfones from liquid SO2 and liquid olefins. These reactions are
radical-catalyzed and will only proceed if the temperature is below
a ceiling temperature Tc given by T c = AH2/AS2, where AH2 and AS2
are the heat and entropy changes, respectively, which accompany
the formation of polymer from a given reaction mixture. The value
of AH2 will differ from that of AHX to a small extent depending on
1) heat of mixing of the monomers AH3 and 2) the heat of solution of
the polymer AH4. The relation between these quantities is shown in
the following cycle, where M denotes olefin and N denotes SO2, pres-
ent in excess, and the AH values are per mole of olefin [9]:
AH3
N(liquid) + M(liquid) >~ mixture (liquid)

AH, I I AH2
polymer (+ N + M) £_ polymer(solution)
(solid)
AH2 = AH2 + AH3 - AH4 (4)

The values of AH2, AH3, and AH4 will depend on the initial relative
amounts of M:N and on the percentage conversion to polymer. The
values of AH3 have been determined for some olefins using an ini-
tial ratio M:N = 1:10, which is the same composition used for the
determination of AH2 by direct calorimetry [10]. Precise values of
262 H. SAWADA

Table 2
Heats of Copolymerization

Standard -AH b
Monomers statesa (kcal/mole) Refs.

Sulfur dioxide:
-Cyclohexene ss 9 c
-Propylene sc 10.1 10
-Butene-1 ss 10.6 10
lc 10.0 9
-cis-Butene-2 ss 10.07 10
Ic 8.93 9
-trans-Butene-2 ss 9.35 10
Is 8.43 9
-Isobutene lc 7.41 9
lc 9.4 15
-Hexene-1 £S 10.4 10
-Hexadecene ss 10.0 10
ss 9.6 d
-Cyclopentene ss 10.8 10
Maleic anhydride:
-Vinyl acetate ss 20.2 7
-Isopropenyl acetate ss 17.8 7
•Styrene ss 19.3 8
-a-Methylstyrene ss 17.3 8
-Allyl chloride ss 17.7 8
-Vinyl n-butyl ether ss 21.5 8
Diethyl fumarate:
-Vinyl acetate ss 18.6 7
-Styrene lc 16.4 6
Diethyl malate:
-Vinyl acetate ss 20.0 7
Maleimide:
-Styrene ss 20.9 8
-a-Methylstyrene ss 17.2 8
Fumaryl chloride:
-a-Methylstyrene ss 17.1 8
-Styrene ss 19.1 8
a
Monomer and polymer states: 1, liquid; s, solution; c, condensed
b
AH represents the heat liberated by the reaction of 0.5 mole of each monomer to
form one mole of mixed monomer units.
cj. E. Hazell and K. J. Ivin, Trans. Faraday Soc, 58, 342 (1962).
d
F. S. Dainton, K. J. Ivin and D. R. Sheard, Trans. Faraday Soc, 52, 414 (1956).
THERMODYNAMICS OF POLYMERIZATION. VI 263

AHj were derived from the heats of combustion of the polysulfones.


Thus from Eq. (4) we can find AH4 in these cases. It was impossible
to measure directly the heat of copolymerization of SO2 with iso-
butene by the fusion calorimeter at 26.9°C because the ceiling tem-
perature for this polymerization is at about 5°C. However, isobutene
polysulfone is insoluble in the reaction medium so that we may set
AH4 equal to zero and derive AH2. The values are summarized in
Table 3 [9].
Table 3
Heats in kcal per Base-Mole Polymer or per Mole Olefina
Olefin -AH t -AH 2 -AH 3 -AH 4

Butene-1 20.06 ± 0.52 21.2 ± 0.1 1.43 0.3 +0.6 b

Isobutene 14.82 + 0.40 15.65 ± 0.4b 0.82 0

cis-Butene-2 17.85 ± 0.36 20.15 ± 0.1 0.64 -1.65 ± 0.5 b

trans-Butene-2 16.86 ± 0.35 18.7 +0.1 0.66 -1.2 + 0.5b


a
Reprinted from Ref. 9 by permission of the Faraday Division of the Chemical Society.
b
Values underlined are derived by means of Eq. (4).

Correlations between the heats of copolymerization of SO2 with


olefins and other heats of addition to olefins are evident. In Table 4
are listed the heats of hydrogenation of gaseous and liquid olefins,
heats of bromination, and heats of self-polymerization [10]. It can
be seen that cyclopentene has the lowest heat of hydrogenation of
any of the listed olefins but the highest heat of copolymerization
with SO2. The reason for this high heat of copolymerization might
be that the presence of the five-membered rings in some way en-
ables the cyclopentene polysulfone to take up energetically more
favorable conformations which are denied to other polysulfones [ll].
B. Ceiling Temperature
The ceiling temperature effect was first discovered in the forma-
tion of olefin polysulfones [12] through its thermodynamic explana-
tion was not realized for many years [13].
In a comprehensive study of ceiling temperature of olefin poly-
sulfones, Cook, Dainton, and Ivin [14] attempted to determine the
ceiling temperatures T c (at which the free energy change for the
reaction is zero) for the formation of polysulfones from a wide
264 H. SAWADA

Table 4
Heat of Addition to Olefinsa

Selfpolymerization
+S0 2 +H2 (calculated per one
at 26.9°C at 25°C mole of olefin)
Olefin (1) (g) (1) (g) (1)

Propylene 20.2 29.72 29.8 20.6 20.05

Butene-1 21.2 30.12 30.29 20.6 20.0

Hexene-1 20.7 30.00 30.2 20.5 19.8

Hexadecene-1 19.9 30.00 - -

trans-Butene-2 18.7 27,48 27.76 18.1 17.0

cis-Butene-2 20.2 28.48 28.91 19.1 17.9

Cyclopentene 21.65 26.33 26.44 -


a
Reprinted from Ref. 10 by permission of the Faraday Division of the Chemical
Society. Heats of reaction expressed in kilocalories per mole.

variety of olefins and allyl compounds under standard concentration


conditions ([M][s] = 27 mole 2 /liter 2 , M = unsaturated compound, S =
SO 2 ). The r e s u l t s a r e discussed in t e r m s of s t r u c t u r a l effects on the
heat and entropy changes and on the monomer activity coefficients
and polymer solubilities. Some of ceiling t e m p e r a t u r e s for poly-
sulfone formation a r e shown in Table 5 [14].
When the poly(olefin sulfone) is insoluble in the reaction mixture,
the expected relation between T c and the activities of olefin and SO2
i s readily deduced from the following cycle [15]:
AG,,
C4H8(liquid) + SO2(liquid) (l/n)[C 4 H 8 SO 2 ] n (solid)

/ (filtration)
mixture (liquid) mixture(liquid) + polymer(solid)
AG = 0
atT c
The condition for equilibrium at the ceiling temperature is
AGIC = AGm + AG3 = RT C In a m + RT C In a s
THERMODYNAMICS OF POLYMERIZATION. VI 265

Table 5
Ceiling Temperatures of Polysulfone Formation3

Olefin Tcb (°C)

Straight chain 1-olefins


Ethylene >135
Propylene 90
1-Butene 64
1-Pentene 63
1-Hexene 60
1-Hexadecene 69
Branched 1-olefins
Isobutene 4.5
2-Methyl-l-pentene -34
3-Methyl-l-pentene 36
3-Methyl-l-butene 14
2-OIefins and cycloolefins
trans-2-Butene 38
cis-2-Butene 46
Cyclopentene 102.5
Cyclohexene 24
Allyl compounds
Allyl alcohol 76
Allyl ethyl ether 68
Allyl formate 45
Allyl acetate 45
Allylacetic acid 66±4
a
Reprinted from Ref. 14 by permission of John Wiley & Sons, Inc.
b
Values are corrected to [M] [S] = 27 mole2/liter2.

where am and a3 represent the activities of olefin and SO2, respec-


tively.
At equilibrium
AGlc = AH1C - TcASlc
and therefore
AH1C - TcASlc = RTC In amas
where AHlc and ASIC are the enthalpy and entropy changes for the
conversion of 1 mole of pure liquid olefin and 1 mole of pure liquid
SO2 to 1 base-mole of insoluble 1:1 copolymer. It can then be shown
that the ceiling temperature T c for the copolymer is given by [15]
266 H. SAWADA

AHle = AH lc
(5)
ASlc + R l n a m a 3 AS l c -+Rlnf

where fm and fs are the mole fraction activity coefficients of olefins


and SO2, respectively, and xm and x3 are the mole fractions of two
monomers in the monomer mixture, respectively.
Ceiling temperatures have been measured over the complete
composition range for isobutene [15] and for 3-methyl butene-1 [16].
Both of these olefins are completely miscible with SO2 in all pro-
portions, and poly(olefin sulfones) are insoluble in all reaction mix-
tures. Figure 3 shows the results obtained by Brady and O'Donnell
[16], in which plots of ln(frofsxmx3) against 1/T consist of two arms
in each case, one arm corresponding to SO2-rich mixtures and the
other to olefin-rich mixtures. It is probably the swelling of the poly-
mer in SO2-rich mixtures that results in the formation of crystalline
polymer (for a further discussion, see Ref. 16).

Fig. 3. Ceiling temperature equation plots for the copolymerization of 3-methyl butene-1
(A) and isobutene (B) with sulfur dioxide showing the discrepancies between sulfur dioxide-
rich mixtures (lower arm) and olefin-rich mixtures in each case. Reprinted by permission
of Microforms,International Marketing Corp. from Ref. 16.

Irradiation of mixtures of butene-1 and SO2 in the gas phase can


produce poly(butene-l-sulfone). The rapid decrease in polymeriza-
tion rate with increasing temperature is good evidence for a ceiling
THERMODYNAMICS OF POLYMERIZATION. VI 267

temperature effect. It was later recognized by Brown and O'Donnell


[17] that the ceiling temperature is lowered by about 60cC compared
with the liquid phase.
C. Q-e Scheme
Alfrey and Price [18, 19] proposed that the rate constant for a
radical-monomer reaction, e.g., for the reaction of M^ radical
with M2 monomer, is given by
k12 = P,Q2 exp(—e,e2) (6)
where Px is a measure of the general reactivity of the growing radi-
cal, Q2 is a measure of the general reactivity of the monomer, and
e1 and e2 are measures of the polarities of the radical and monomer,
respectively. The P and Q terms primarily define the resonance
effects in the radical and monomer. If the same e values applies to
both a monomer and its radical (that is, et defines the polarities of
Mt and M^, while e2 defines the polarities of M2 and M2-), then ac-
cording to Eq. (6), rj and r2 for the system are found to be

rx =-§iexP{-e1(e1 - e2)} (7)

r 2 =-^exp{-e 2 (e 2 - e i )} (8)

Schwan and Price [20] have developed a modified copolymeriza-


tion equation by utilizing the parameters q and e in place of Q and e.
Their equations for r1} r2, and r x r 2 are:
rx = exp{-( qi - q2)/RT}exp{-7.23 x 10aoe1(e1 - e2)/RT} (9)
r 2 = exp{-(q2 - qi )/RT}exp{-7.23 x 102oe2(e2 - eJ/RT} (10)
r x r 2 = exp{-7.23 x I020(e1 - e2)2/RT> (11)
The advantage of this modification is that the parameter q and e are
temperature independent (whereas Qand e depend upon temperature),
and consequently the above equations will hold for any temperature
range.
Evaluation of the activation energy for the reactions of radical
additions to vinyl compounds was studied as an extension of the
theory employing two Morse functions for the initial and the final
268 H. SAWADA

systems [21]. The Evans-Polanyi and Semenov rules, the Hammett-


type rule, and the Q-e scheme in the radical copolymerization were
derived by using this approach and have been discussed in detail [21].
As has already been discussed [22], the activation energy of the
propagation reaction can be expressed by
E = Eo - k(Rf - Ra - Rm) (12)
where Eo is the activation energy of propagation reaction involving
polyethylene radical and monomeric ethylene, and the factor k has
a value less than unity, and Rf, Ra, and Rm are the respective reso-
nance energies of the initial radical, the monomer, and the radical
formed after addition of the monomer.
Thus,

E t (corresponding to k n ) = Eo - k(Rtj - Ra - Rm}


and, similarly,

E2 (corresponding to k12) = Eo - k(R,2 - Ra - Rmi)


Therefore,
Ex - E2 = k[(Rmi - Rfi) - (Rma - R,2)]

If the preexponential factors for both reactions are the same, we


have
r, = i i i = exp{k[(Rm2 - Rf2) - (Rmj - Rfl)]/RT} (13)
K 12

This gives [23]

Q = exp[-(Rm - Rf)/RT] (14)


Several authors [24-27] have attempted to give a theoretical basis
to the Q-e scheme, even though Alfrey and Price [18] pointed out
that this was strictly an empirical approach. Indeed, in many cases
calculated Q and e values do not agree with literature data or, con-
versely, relative reactivity ratios calculated from published Q and
e values do not agree with experimental data.
r THERMODYNAMICS OF POLYMERIZATION. VI 269

D. Substituent Effect
Hammett [28] has proposed that the effect of meta- or para-sub-
stituents of side-chain reactions of benzene can be expressed as
log(k/k0) = po (15)
where kp and k are the rate or equilibrium constants for the reaction
of the unsubstituted and substituted compound, a is a parameter hav-
ing a single value for each substituent, and p is a constant for any
particular reaction.
The parameters a and p are interpreted as measures, respec-
tively, of the ability of the substituent to withdraw electrons or
donate them to the site of reaction and the effect of such electron-
availability on the reaction.
Since Walling et al. [29] reported that the reactivities of nuclear-
substituted styrenes toward a polymer radical were correlated with
Hammett's a constants, similar relationships were observed on
phenyl vinyl sulfides [30] and vinyl benzoates [31].
Imoto et al. [32] carried out the mutual copolymerization of p-sub-
stituted styrenes, and found that

log(k/k0) = pa + R (16)
where R is a parameter of the resonance stabilization in the transi-
tion state. This can be written in terms of the resonance substituent
constant ER and the effect of such resonance availability y on the
reaction so that
logft/ko) = po + yER (17)
which is proposed by Yamamoto and Otsu [33].
When the polar effect is significant, as often occurs with ionic re-
action, y is equal to zero and Eq. (17) reduces simply to Eq. (15).
However, when only resonance effects play a significant role, as in
certain radical reactions, the value of p becomes zero and Eq. (17)
reduces to [34]:
logtk/k,,) = r E H (18)
On the other hand, Bamford et al. [35] proposed that rate con-
stants for the reactions between polymer radicals and a number of
270 H.SAWADA

substrates, including monomers, are given by


log k = log kT + aa + /3 (19)
where a and j3 are constants characteristic of a given monomer, a
is the algebraic sum of Hammett's a constants for the substituents
in the terminal monomer unit of the radical, and kT is the rate con-
stant for chain transfer to toluene (taken as a nonpolar measure).
Equations (16), (17), and (6) can be written in the form of the
following equations, respectively:
log k = log k,, + pa + yER (20)
log k = log kg + per + R (21)
log k12 = log Px - exe2 + log Q2 (22)
Comparison of Eqs. (20)-(22) and Eq. (19) indicates that in each case
the first term of the right-hand side stands for the general reactivity
of the attacking radical, the second term relates to the polar effects
of both attacking radicals and reactants, and the third term relates
to the resonance factor of the reactants. The factor e of the scheme
increases almost linearly with Hammett's a constant [20, 36]. Cam-
marata et al. [37] have found that with substituted styrenes as the
monomer, the parameters E R , R, and log Q2 are correlated by the
square of Hammett's a constant. Thus
log k = acr + bff2 + c (23)
where a, b, and c are constants.
Chernobai et al. [38] derived a relation between Hammett's a
constant, activation energies, and thermodynamic characteristics
in polymerization of vinyl monomers comprising substituted sty-
renes. Equations were derived to relate a with activation energies,
propagation constants, and heat of polymerization.
Let us consider the reactions of series of alphatic compounds
where the reaction sites are not in conjugation with the substituents.
Since resonance effects are only of minor importance, steric effects
are often as important as polar effect. The value of log(l/r 1 ) =
Iog(k12/kn) is considered to represent the activity of monomer M2
relative to the reactive center of the reference monomer Mx. Thus
it was found that [39]
log(l/r1)=p*CT* + 6Es (24)
THERMODYNAMICS OF POLYMERIZATION. VI 271

where a * is Taft's polar substituent constant, Es is Taft's steric


constant, and p* and 6 are the reaction constants. In the copolymer-
ization of styrene and a-substituted methyl acrylates, the reactivities
of the esters toward the styryl radical were influenced by both the
steric and polar nature of the a-substituent and followed Eq. (24)
[40].
Chikanishi and Tsuruta [41] found a linear relationship between
log^l/rj) and Taft's steric constant in the case of the relative re-
activities of methyl a-alkyl acrylate toward attack of styryl radical.
Thus
= 6ES (25)

In the case of 2-vinyl-4-substituted-l,3-dioxolanes (M2) with acry-


lonitrile (Mj), a linear relationship was obtained by plots of logU/rJ
with Taft's steric constant, but not with the a * value in spite of the
remote distance of the substituent from the reaction center [42].
Yamashita et al. [42] suggested that the substituent exerted a steric
effect on the double bond in the transition state. Fairly good linear
relations were also obtained when the relative reactivities of a-,
cis-/3-, and trans-/3-alkylstyrenes toward the polyacrylonitrile
radical are plotted against values of E s in Eq. (25) for alkyl groups
[43].
On the other hand, Otsu et al. [44] reported that the relative re-
activity of alkyl esters of methacrylic acid was influenced only by
the polar effect and not by Es. Values of p* thus obtained are usu-
ally small and depend on the polar character of the attacking radi-
cals. Since ester alkyl groups in alkyl methacrylates are well away
from the reactive double bond, steric effects are negligible.
The reactivities of 1,1-disubstituted monomers such as a-sub-
stituted acrylonitrile and acrylic esters toward polystyrl radical
were given by [45]
log rt = A log Qx + O.83CTP (26)

where A log Q, and crP are the resonance and polar substituent con-
stants of a substituents, respectively.
Consequently, it seems reasonable to assume that log(k/ko) can
be expressed as a sum of polar, resonance, and steric factors. Thus
log(k/ko) = F P + F R + F s (27)
where F P , F K , and F s represent the polar factor, the resonance fac-
272 H. SAWADA

tor, and the steric factor, respectively. For p-substituted styrenes,


we have
F P + FR » F s = 0
For a-substituted acrylates where the substituents are not in con-
jugation with the reactive center, we have
F p + F s » FR s 0
For a-substituted acrylates where the substituents are located on
the reactive center, we have
F s » F p + FK = 0
The treatment in this section is at present best considered as an
empirical approach to placing monomer reactivity on a quantitative
basis. Any theoretical approach has not yet been quantitatively
. successful.
E. Effect of Polymerization Temperature
Lewis and co-workers [46] reported copolymerizations of styrene
in which activation energies of reactivity ratios were slight and there
was essentially no difference in the activation entropies of the com-
peting propagation reactions. Only in the reaction of the diethyl
fumarate radical with styrene made the difference in entropies of
activation clearly differ from zero by more than experimental error.
Goldfinger and Steidlitz [47] and Johnston and Rudin [48] also found
a preponderance of negative activation energies.
Since a reactivity ratio is the ratio of two rate constants, it may
be expressed as [46]

r, = exp|""» R "»+ - " " 7 T n " * | (28)

where S u t, Sjjt, H u t, and H,,t are, respectively, the entropies and


heats of activation for reaction of a radical ending in monomer i
with monomers i and j .
For the binary copolymerization system, rxr2 is given by
r x r 2 = exp{[(H12* + H21*) - (Hu* + H22*)]/RT - [(Su* + S21*)
-(Snt+Sj)]/R}
= exp{ni2VRT - XUVR} (29)
= exp{AG12VRT}
THERMODYNAMICS OF POLYMERIZATION. VI 273

where

oj * (H12* + Hgl*) - (Hxx* + H 22 t)

Xl2 * = (Sxa* + Sax*) - (S xl * + S22»)

AG12* = flja* — Tx12*

Therefore from Eq. (29) we find that

9 to r r
i 2 _ _ ^ ' (30)

Whether or not rxr2 increases or decreases as the temperature in-


creases depends upon the sign of O12t for the copolymerization sys-
tem.
Taking the logarithm of both sides of Eq. (30),

i / g (31)

According to Eq. (31), a plot of In r ^ vs 1/T should give a straight


line with slope equal to n i 2 t/R, and the intercept of In r x r 2 at 1/T = 0
is equal to —x12VR. In Fig. 4 In r^., values are shown as a function
of 1/T for styrene-methyl methacrylate.
If x12t > 0 and fl12t > 0, or \ l2 * < 0 and S212t < 0, there must be
a temperature at which AG12t is zero. This temperature will be
given by

At this temperature, r x r 2 = 1 and therefore the copolymer is com-


pletely random.
The effect of temperature on the monomer reactivity ratio has
been examined experimentally for styrene-methyl methacrylate [49],
aerylonitrile-methyl methacrylate [50], and methacrylonitrile-
styrene [51]. A list of the values of n i2 * and X12* of these binary
copolymerization systems is given in Table 6. It is observed from
this table that all of the values of J212t are negative, indicating that
r x r 2 increases as the temperature increases. It will be noted that
all values of X12* a r e nearly equal to zero, indicating that the values
of entropy change for most polymerizations are quite similar. Thus
the completely random copolymer appears to fail for these copolym-
274 H. SAWADA

Fig. 4. Least-square fit of literature values of In ifo vs 1/T for the copolymerization of
styrene and methyl methacrylate: In r ^ = 0.17 — 521/T. Data from Ref. 49 and H. Mark,
B. Immergut, E. H. Immergut, L. J. Yong, and K. I. Boynon, in Polymer Handbook
(J. Brandup and E. H. Immergut, eds.), Wiley-Interscience, New York, 1966, p. 11-142.

erization systems. For these systems, however, the tendency toward


random copolymerization is enhanced by an increase in temperature.
A plot of In rx vs l/T affords a measure of the difference in en-
tropies and enthalpies of activation. O'Driscoll [52] assumed that
the entropies of activation for the two competing propagation reac-
tions are approximately equal, and thus

In r, = - H u t - (33)
THERMODYNAMICS OF POLYMERIZATION. VI 275

Table 6
Values of fi, 2 * and X12 * for Binary Copolymerization Systems

X»*
Binary systems (kcal/mole) (eu) Refs.

STa/MMAb -1.03 0.33 49

-0.79 - e

AN=/MMA -1.38 0.46 50


MANd/ST -1.08 -0.85 51
a
ST = Styrene.
b
MMA = Methyl methacrylate.
C
AN = Acrylonitrile.
d
MAN = Methacrylonitrile.
e
S. Russo, B. M. Gallo, and G. Bonta, Chim. Ind., 54, 521 (1972).

The temperature dependence of the reactivity ratio is

djnri_ H»t - H n t (34)


d(l/T) " ~ R •
From Eqs. (33) and (34) we find

&T1- (35)

The term (—RT In r1) = (H11t — H u ^ i s the apparent activation energy


for r. Since the value of r itself appears in its activation energy,
the only reactivity ratios which will exhibit significant temperature
dependence will be those which are very large or very small.
Equation (35) predicts that a reactivity ratio greater than unity
will decrease with increasing temperature and vice versa. The
value of r will always approach unity with increasing temperature.
Therefore, the tendency toward random copolymerization is en-
hanced by an increase in temperature. On the other hand, the ten-
dency toward alternating copolymerization is reduced by an in-
crease in temperature. For the styrene-methyl methacrylate sys-
tem, the reactivity ratios decrease with decreasing temperature,
and this effect leads to a greater degree of alternation in copoly-
mers produced at lower temperatures [53],
276 H. SAWADA

Yamada and Yanagita [54] have proposed the product relation of


the copolymerization reactivity ratios:
Txr/=a or (k n /k 12 )(k 22 /k 21 ) s = a (36)
where a and/3 are the thermodynamic parameters that will be de-
fined in the following description.
In terms of transition state theory, the reactivity ratios can be
expressed by
r, = exp[(Snt - S 12 t)/R]exp[-(H u t - H12*)/RT]
r2 = exp[(S22t - S21t)/R]exp[-(H22* - H21*)/RT]
and substituting into Eq. (36), we find

= exp p j
- H 21 *)1
RT J

If Eq. (36) is independent of temperature, there is the relation

The parameters /3 and a can be defined as


P = -(H u * - H12*)/(H22t - H21t) • (38)
a = exp{([S11t - S12* ] + /3[S22t - S21t ])/R} (39)
The radical copolymerization systems were classified into two
groups corresponding to the negative and the positive sign of /3 by
Yamada and Yanagita [54]. They proposed two types of mechanisms
which influence interactions between a polymeric chain end and the
comonomer.
F. Effect of Solvent
The Alfrey and Price scheme [18, 19] predicts that the dielectric
constant of the solvent affects the monomer reactivity ratios. How-
ever, it is well known that the solvent effect is not as significant in
radical copolymerization [55]. For example, Lewis et al. [46] and
THERMODYNAMICS OF POLYMERIZATION. VI 277

Price and Walsh [56] indicated that solvents are without effect on
the monomer reactivity ratios of the radical copolymerization of
styrene and methyl methacrylate.
Solvents which are capable of forming hydrogen bonds with mono-
mers, e.g., acrylic and methacrylic acids and numerous monomers
containing nitrogen, have a strong influence on these copolymeriza-
tions. Joshi [57] first pointed out that the copolymerization anom-
alies (e.g., deviation from the ideal composition equation) are better
explained by the association of the monomeric state through hydro-
gen bonding and possibly even by dipole-dipole interactions of lesser
magnitude which stabilize the monomer state and lower the heat of
polymerization.
Kerber [58] correlated changes in the reactivity ratios of styrene
and acrylic acid with associations of the solvent with acrylic acid
through hydrogen bond.
For acrylamide [59] and methyl acrylamide [60], which are non-
ionizable monomers, the existence of the following equilibrium was
assumed:
©
CH,=CH—C—NH, - t »*- CH,=
2
CH—C=NH,
2
=p=^=
II le
o o°
CH2=CH—C=NH
OH
The strong influence of the solvent on the copolymerization of acryl-
amide should be correlated with the possibility of an amide-enol
equilibrium because such an equilibrium could be influenced by sol-
vents of different polarities and dielectric constants.
For the N,N-dimethylacrylamide [61], the enolic form cannot be
present because no amidic hydrogen is available:

/CH 3 © .CH3
CH2=CH—C—NQ -« >- CH 2 =CH-C=N('
CH
o' 3 eo CH 3

Thus, in this case no influence of the solvent on copolymer composi-


tions was found.
Minsk et al. [62] indicated quite conclusively that both polar
effects and hydrogen bonding affect the copolymerization of acryl-
amide and styrene.
Ito and Otsu [63] found that the copolymerization parameters
278 H. SAWADA

changed significantly with the polarity of solvents, and the relative


reactivity of methyl methacrylate toward the polystyryl radical was
correlated with the E T values of solvents, which are empirical values
of the proton-donating ability of solvents.
The reactivity of vinylpyridine depends to a large extent on the
acidity of the solvent used [64]. This effect was tested using sty-
rene—2-vinylpyridine, styrene—4-methyl-2-vinylpyridine pairs.
With a reduction in solvent acidity, the tendency to alternation de-
creases as a consequence of the lower polarization of vinylpyridine
in the H-complex and a slight difference in the polarities of the
styrene and vinylpyridine bonds.
The influence of solvent and temperature has been analyzed for
the methacrylic acid-methyl methacrylate system. According to
type of solvent used, methacrylic acid can exist either as an indivi-
dual molecule (I), or be associated with itself (II), with the solvent
(III), or with the carboxyl groups of the copolymer formed (IV).
Forms II and IV mainly appear in petrol and benzene, and are more
reactive than Form III which appears in isopropanol and tetrahydro-
furan [65].
In principle, solvents should have an effect on radical copolymer
composition only in special cases such as the copolymerization of
acid or basic monomers or ionizable monomers where the solvents
modify the free energy of the initial and final state.

II. IONIC COPOLYMERIZATION


A. Energetic Characteristics
If we consider the propagation step in the case of ionic copolym-
erization, it is clear that, from the point of view of polarity, there
is no possible difference between the attacking ions. Therefore, it
is reasonable to assume that the competition between two monomers
toward the attacking carbonium ion or carbanion will be independent
of the nature of the monomer last added to the chain. The propaga-
tion rate constant k u is identical with k 2l , and k12 is the same as k22:

Ki = k2i . . (40)
k22 = k12 (41)

Therefore

k n A 1 2 = k 21 /k 22 (42)
THERMODYNAMICS OF POLYMERIZATION. VI 279

and the product of the reactivity ratios is unity:

rxr2 = 1 (43)

Consequently, the equation derived for free radical propagation


becomes [66]

d[Mj/d[M 2 ] = r i [Mj/[M 2 ] (44)

Equation (44) will hold only for the case of monomers of similar
polarities [67].
Equations (40) and (41) have been denied by some experimental
results of Okamura et al. [68], who showed that the polymer cation
derived from a less reactive monomer had a larger reactivity. How-
ever, the same conclusion was reached in anionic copolymerization
by Kuntz and O'Driscoll [67] who assumed the more general equa-
tions

k u = xk21 (45)

k12 = xk22 (46)

It is to be noted that Eqs. (40) and (41) are special cases of Eqs. (45)
and (46) where x has a value of unity.
O'Driscoll [69] developed the kinetics of copolymerization of dis-
similar monomers. It was shown that the ratio of monomers in the
initial copolymer d t M j M M j is directly proportional to the square
of the initial monomer ratio ([Mj^Mj) 2 when the monomers are
dissimilar:

d[Mj/d[M 2 ] = (kx /k 2 )(k u /k22)([Mj/[M2])2 (47)

The above derivation assumes that the initial homopropagation is


fast relative to the initiation, and that the crossover reaction is a
negligible term at the beginning of the polymerization. The conclu-
sion is drawn that exponent values of 1.0 and 2.0 represent the ex-
tremes of behavior for monomers of very similar and very dissimi-
lar polarities. The system styrene-methyl methacrylate was con-
sidered as an example of Eq. (47), while styrene-isoprene served as
an example of Eq. (44).
The reactivity ratios of several vinyl monomers such as styrene,
acrylonitrile, methacrylonitrile, and methyl methacrylate have been
determined during their copolymerization initiated with phenyl mag-
280 H.SAWADA

nesium bromide [70]. With couples of monomers of very different


electronegativities, such as the systems styrene-methyl methacryl-
ate and acrylonitrile-methyl methacrylate, their copolymerization
equation will be given by Eq. (47). An increase of temperature gen-
erally increases a cross-propagation reaction, and consequently
Eq. (44) will hold for the system acrylonitrile-methyl methacrylate
[70].
In terms of absolute reaction theory, the propagation rate con-
stant k u for the polymer chain (P t ), which ends in monomer i adding
monomer j (Mj), is given by
k,, = (RT/Nh)exp{-AFtyRT}
FBj*)/RT} (48)

where F , ^ refers to the free energy of the activated complex formed


between P, and Mj.
Substituting the four forms of Eq. (48) into Eq. (43), we have

•Til —*•>>•> r
•> (49)
Figure 5 illustrates schematically the specific case of anionic
copolymerization of butadiene-styrene [67]. The uppermost curve
represents the addition of styrene to a butadiene chain end; the one
beneath it represents the addition of butadiene to a butadiene chain
end. The difference between the peaks of these two curves is the
same as the difference between the peaks of the two lowest curves
TRANSITION STATE
,12
O
cr
UJ

LU
LU
LU MONOMER POLYMER
21
cr

REACTION COORDINATE

Fig. 5. Illustration of possible transition state energy levels. Reprinted by permission of


John Wiley & Sons from Ref. 67.
THERMODYNAMICS OF POLYMERIZATION. VI 281

which represent styrene adding to a styrene end and butadiene add-


ing to a styrene end. Since the enthalpy change AH* of monomers
of similar polarities should be the same, we may approximate to
Sia*-S11t=S22t-S21* (50)
where S12$, S u t, S22t, and S21t are the corresponding entropy changes.
From Eq. (50) it is seen that there is a large change in entropy in-
volved in the transition state during the addition of a diene monomer
to a chain end as compared with addition of a styrene unit. This
equation holds for the propagation step in anionic polymerizations
in nonpolar solvents [67].
Let us now consider thermodynamics of the transition state more
closely when r x r 2 < 1. Using the four versions of Eq. (48) where
i,j = 1, 2, we find [71]

RT In r.r, = - ( F u * + F22t) + (F12* + F21*) < 0

F 12 t + F 2 1 t < F u t + F 2 2 t (51)
' H a * + Hia* - T(S21* + SX2t) < H u * + HJ - T(S u t + Sj)
(52)
In an earlier section we found that the observed values of entropy
of polymerization are generally in the narrow range between 25 and
30 eu [72], and that the entropy of activation differs little from en-
tropy of polymerization [73]. The entropies of activation are quite
similar as pointed out by some experimental results of Lewis et al.
[46]. Thus

H l2 J +H 2 X t<H 1 1 *+H M * (53)


and hence

Equation (53) is the usual case in free radical copolymerization and


is responsible for the observed alternation tendency.
Assuming a linear free energy relationship [74], AF,, can be ex-
pressed in terms of a free energy of activation AF,,$:
AF,, = -aAF,jJ + /3 (54)
where a and /3 are constants. From Eq. (54) we see that, when
r x r 2 < 1,

H u + H22 < H12 + H2l (55)

In this case we find

fi = (H12 + H21) - (H u + H22) > 0

This case was discussed earlier where the treatment of Sawada was
presented [75]. For the case where r t r 2 ^ 1,

. H18t + H21t % H u t + H22t (56)

and similarly,

H12 + H21 <F H u + H22 (57)

and 0 < 0.
Wall [76] suggested that the electronegativity of the monomer, e,,
could be distinguished from that of the chain end, e ^ . Thus

r x r 2 = exp{(e, - e,)(e1* - e,*)} (58)

It is reasonable to consider that the difference in carbonium or car-


banion electronegativities may be quite small, and thus [71 ]

e,*-ej*=0 (59)

In ionic copolymerization, monomer reactivity ratios are signifi-


cantly affected by solvent and catalyst, and systematic interpretation
is difficult even in the simplest case. It is possible only to specu-
late on the reasons for the apparent anomalities. Since the mecha-
nism of ionic propagation is more complex than free radical pro-
pagation, e,* would be expected to be a function of the solvent and
counterion. Thus, in ionic copolymerization, other factors than the
stabilization energy, especially the electrostatic interaction between
monomer and ion, should be taken into consideration to explain the
reactivity [77]. As a result, it is important that the reactivity ratio
products change from approximately unity to values much greater
than unity as the solvent is changed [71]. Good examples of this come
from the experimental work of Overberger and his co-workers [78].
THERMODYNAMICS OF POLYMERIZATION. VI 283

If there is a linear relationship between the ground state free


energy and that of the transition state, we have
F l f t = (1 + a iJ )Fp 1 + (1 + /3u)FMj (60)
where a and p are constants. Thus

k u = P.Q, expC-e^e,) = -§£ exp{-(a n F P i + ftjF^/RT} (61)

If no synergistic effects are encountered (i.e., an = ctu and /3U =


i = ft), we find [71]
r.r, = expi-^F^ - ftF^/RT} (62)
Therefore, r t r 2 can be expected to be precisely unity only for quite
similar monomers. If their contributions to the transition state are
slightly different, r,r 2 will be smaller than unity. If there are syner-
gistic effects, r x r 2 may be much greater or less than unity [71].
The dependence of relative reactivity of cationic copolymeriza-
tion on solvents was explained in terms of the difference in the
ability of complex formation between monomer and growing car-
bonium ion [79, 80]. The solvent effect in anionic copolymerization
was also studied by O'Driscoll [81].
B. Reactivity
Walling et al. [29] correlated monomer reactivity of substituted
styrenes with structure in radical copolymerization systems using
Hammett's op concept. The same concept was extended for cationic
copolymerizations.
Overberger et al. [82] studied the effect of para-substitution in
the cationic copolymerization of a-methylstyrene with several para-
substituted styrenes. It was found that the logarithm of the monomer
reactivities gave a good fit to the straight line when plotted vs Ham-
mett's a-values for the various substituents with the exception of
p-methoxystyrene. Tsuruta [83] and Brown and Okamoto [84] have
suggested that linearity was improved by the use of a* values. A
similar treatment of the data obtained from the three para-substi-
tuted a-methylstyrenes studied by Dunphy and Marvel [85] re-
sulted in a straight line. Similarly, Tobolsky and Boudreau [86]
found that 1) the electrophilic substituent constants, a*, of Brown
are superior to Hammett's a values in correlating reactivity ratios
in ionic copolymerizations of substituted styrenes, and 2) the reac-
284 H. SAWADA

tion constant is positive for anionic, nucleophilic copolymerizations


of substituted styrenes. Mizote et al. [87] proposed Eq. (63) in their
study of steric effects of p-methyl group on monomer reactivity in
cationic copolymerization of styrene (Mx) and p-methylstyrene (M2).
F s = log r 2 - log(l/r 1 ) (63)
The relative reactivities of vinyl aromatic monomers in coordi-
nated anionic copolymerization were interpreted in terms of polar
and steric effects of the substituents [88]. The relative reactivity,
a, of each monomer was defined as the ratio between its overall
rate constant kmon and the correspondent constant of styrene at the
same temperature:
a =
^mon/Kstyr

For halogen- or p-n-alkyl-substituted styrenes and for some other


monomers, a linear relation may be considered to hold between
log a and the a constant values as shown in Fig. 6. On the other
hand, a few other monomers (p-isopropylstyrene and m-methyl-
substituted styrenes) have a lower value than those which would be
predicted from the above linear relation. The behavior of these
last monomers could be due to steric effect.
The most interesting result is the negative slope of the straight
line of Fig. 6(p= —1.0), indicating that the reaction is favored in the
case of monomers having an electron-releasing substituent. The
greater the electronic density on the double bond, the more reactive
the monomer is, as if an electrophilic attack on the double bond was
kinetically determinant. This might be at first sight appear as a
typical feature of cationic polymerization in which a positive grow-
ing ion attacks the monomeric double bond, but there is another in-
terpretation in the mechanism proposed for a "coordination anionic
polymerization" by Natta [89].
Szwarc et al. found that Hammett's relation is obeyed by the rate
constants of anionic copolymerization of p-substituted styrenes with
living polystyrene [90]. For anionic polymerization p = 5.0, while
for radical polymerization p = 0.5 and for coordination polymeriza-
tion p = —0.95.
C. Effect of Polymerization Temperature
Limited data for some radical copolymerization systems show
that Arrhenius plots of log r against 1/T passes a zero, or nearly
zero, intercept. Extrapolation of the plots in cationic copolymeriza-
THERMODYNAMICS OF POLYMERIZATION. VI 285

P- -0-95
(c)O
0

cn (k) (0
O O O (m)
O

-0-5 -

1 i 1 1 1 i

-03 -0 2 -0-1 0-1 02 0-3

Fig. 6. Plots of relative reactivities (k m o n /k s t ) of substituted styrenes with regard to the


styrene. Effect of substituents:' (a) p-0CH3; (b) p-CH3; (c) p-C2Hs; (d) H; (e) p-F; (f) 0-C4H4;
(g) p-Br; (h) p-Cl; (i) m-F; (j) m-Cl; (k) 3,4-(CH3)2; (1) p-iC3H7; (m) m-CH3; and (n) 3,5-
(CH3)2. Reprinted by permission of Huthig & Wepf Verlag from Ref. 88.

tion of isobutylene and styrene gives intercepts, the logarithms of


which are considerably greater than zero [91]. This can be inter-
preted to mean that the elementary growth steps of an ion under-
going cationic copolymerization possess much larger differences
in entropies of activation (S u t — S12i) of their transition state com-
plexes than do those of growing free radicals.
From the plot of log r t vs 1/T, the linear relation was obtained
in the cationic copolymerization of acenaphthylene Mx and n-butyl
vinyl ether M2 [92]. Differences in the energy and the entropy of
activation evaluated from the slope and the intercept are

- H12t = 2.6 kcal/mole and - S12t = 6.5 eu

Since the products of rx and r2 are approximately unity, the corre-


sponding differences must be nearly equal to the values mentioned
above, but of opposite signs:

- H21* = -2.6 kcal/mole and S t - = - 6 . 5 eu

In the propagation steps, the energy of activation, H u , is higher by


2.6 kcal/mole than H^t, but the entropy of activation, S u t, is more
286 H. SAWADA

positive than S^*. Assuming that H12$ = U21$, we obtain H u t — H.,.,* =


5.2 kcal/mole. This value corresponds to a difference in overall
activation energy for the two homopolymerizations. Thus the differ-
ence in activation energies of initiation and termination reactions
are almost negligible [92].
The reactivity ratios in CEVE (2-chloroethyl vinyl ether) [Mj—
a-methylstyrene [M2] copolymerizations have also shown an appre-
ciable temperature coefficient [93]. The logarithm of r : is approxi-
mately linear with 1/T, corresponding to a difference in activation
energy for the two addition reactions to the CEVE cation, H u t — H12t
= 2.5 kcal/mole. For addition to a-methylstyrene carbonium, r 2
has a very low value at 25°C due to depropagation since a-methyl-
styrene has its ceiling temperature in this region. However, we
should not expect depropagation to be important at low tempera-
tures, and here we find that the temperature dependence of l/r 2 is
similar to that of Rowing to the relation of r l r 2 = l,i.e.,H 22 t — H^t,
is also about 2 kcal/mole.
Masuda and Higashimura [94] suggested that in the ionic copolym-
erization of monomers with different structures, fit and x* in Eq.
(29) are usually different from zero and have negative values. On
the other hand, in the ionic copolymerization between homolog, Ot
and x* are almost zero; an example of such a copolymerization
would be the a-methylstyrene—p-chlorostyrene system where the
difference in entropy of activation is very small [95].
In the cationic copolymerization of styrene and p-tert-butylsty-
rene, AHxt = H u t - H12t, AS^ = S u t — S12*, AH.,* = AH22t — H21$, and
AS2$ = S22t — S21t were evaluated [95]. The following relations were
derived between AH^ and AStt:

H,,* =aSit$ +b
AH,* = H u * - H,,* = a(S u t - S,,*) = aAS,*
where a and b are constants for a homologous series. Therefore,
these copolymerizations must have proceeded through the same
mechanism regardless of the type of solvent and catalyst used [96],

m . OTHER COPOLYMERIZATIONS
A. Ring-Opening Copolymerization
Yamashita et al. [97, 98] indicated that the basicity of cyclic
ethers, pKb, is a factor in the cationic copolymerization of cyclic
ethers; the plot of log(l/r t ) vs pKb for monomer M2 gave a straight
THERMODYNAMICS OF POLYMERIZATION. VI 287

line in the copolymerization of 1,3-dioxolane [99]. The linear plot of


l / r r vs pKb means that the proton affinity or basicity toward the
Lewis acid is a measure of carbonium ion affinity. Hence the
nucleophilic attack of cyclic ethers on the cationic active chain
end is assumed to be one of the driving forces in this copolymeriza-
tion. The slope of the plot shows the selectivity of the chain end.
Aoki et al. [100] and independently Tanaka [101] investigated the
effects of both ring strain and basicity of a-monosubstituted cyclic
ethers on their reactivity. The relative reactivities of cyclic ethers
to the poly[3,3-bis(chloromethyl) oxetane] cation are empirically
expressed by [100]

logUAi) = -0.086ARS - 0.31(pKb) + 0.57


where ARS and pKb are characteristic of the ring strain and basicity
of a-monosubstituted cyclic ethers arid are calculated from the dif-
ference in free energy of polymerization and in basicity between Mt
and M2, respectively. The pKb of the substituted cyclic ethers are
easily obtained from the equation pKb = (pKb)x + 2.0u*. The relative
reactivities of the cyclic ethers whose pKb values are not available
may therefore be predicted from this equation whenever the a *
value of the substituent is obtainable from the literature. A linear
relationship can also be observed between the a * value and the free
energy of polymerization of the cycloparaffins calculated by Dainton
et al. [102].
Kagiya et al. [103] pointed out that in cationic copolymerization
of cyclic ethers the relative reactivity can be expressed as a func-
tion of two energetic values of the nucleophilic coordination power
of monomer and the radical bond-dissociation energy of the broken
bond. Thus

+ 6'ADt) (65)
where AAFp is the increase in the free energy change between two
monomers, AD, is the radical dissociation energies of broken bonds,
and cpy and 6' are constants.
The anionic coordination copolymerization of a number of mono-
substituted ethylene oxide derivatives with propylene oxide as the
standard was studied in the presence of a 1:1 diethyl zinc:water
catalyst. The value of logU/rJ is given by [104]
+ 0.267Es - 0.145 (66)
288 H. SAWADA

where ax is the polar constant and E s is the steric constant of the


substituent on the a-oxide. The relative reactivities of the o-oxides
during anionic copolymerization in the presence of sodium phenolate
as catalyst were also found to depend on the polar as well as on the
steric influence of the substituents [105], and thus

logU/rj) = 2.09O"! + 0.154Es + 0.11 (67)


Yamashita et al. [106] studied the cationic copolymerization of
dioxolane and trioxane, mainly by examining the microstructure of
the copolymers by NMR. The incorporation of dioxolane in the co-
polymers decreased with increasing conversion and polymerization
temperature and with increasing dilution by solvent. This may be
attributed to the lower equilibrium concentration of formaldehyde
for trioxane polymerization and the higher equilibrium monomer
concentration found for the polymerization of dioxolane. The re-
sults showed that dioxolane is kinetically more reactive but thermo-
. dynamically less polymerizable owing to a remarkable ease of de-
polymerization of poly(dioxolane).
Yamashita et al. [107] also studied the cationic copolymerization
of dioxolane and tetroxane. The very reactive tetroxane was con-
sumed at an early stage, whereas dioxolane was consumed slowly
until an equilibrium conversion was reached. Thus the incorpora-
tion of the less reactive dioxolane increased with polymerization
time. After constant composition was reached, there appeared to be
no more polymerization of the remaining dioxolane. Yamashita et al.
[107] concluded that tetroxane is far more reactive than dioxolane
from both the kinetic and thermodynamic points of view. However,
one should note that the thermodynamics of these copolymerizations
are very complicated because of the number of side reactions pos-
sible, such as polymer transfer (trans-acetalization) and the back-
biting reactions.
B. Miscellaneous Copolymerizations
The effects of monomer composition on the ceiling temperature
are shown in Fig. 7 for the copolymerization of acetaldehyde and
propionaldehyde [108]. The maximum ceiling temperature occurs
at a polymer composition of 74 mole % acetaldehyde.
The copolymerization of o-phthalaldehyde and styrene was studied
at 0 and -78°C by Aso et al. [109]. At -78°C, o-phthalaldehyde pref-
erentially polymerized to yield living cyclopolymers until an equi-
librium concentration with o-phthalaldehyde monomer was reached.
Thus styrene propagated from the living terminal rather slowly. The
THERMODYNAMICS OF POLYMERIZATION. VI 289

-30

-35 -

25 75 100

MOLE •/. ACETALDEHYDE


Fig. 7. Ceiling temperature as a function of monomer feed composition. Reprinted by
permission of the publishers, IPC Science and Technology Press from Ref. 108.

block structure of the copolymer was confirmed by chemical and


spectroscopic methods. In the copoly me rization at 0°C, the o-
phthalaldehyde unit in copolymer consisted both of cyclized and un-
cyclized units. This copolymer seemed to contain short o-phthal-
aldehyde sequences. Since this polymerization temperature is much
higher than the ceiling temperature (—43CC) of the homopolymeriza-
tion of o-phthalaldehyde, depropagation of the o-phthalaldehyde unit
should be extensive, and the presence of long sequences of o-phthal-
aldehyde units is improbable [109].
The presence of a donor-acceptor charge transfer complex during
the copolymerization of a strong donor monomer (D), such as sty-
rene, with an acceptor monomer (A), such as methyl methacrylate
or acrylonitrile, in the presence of a metal halide or other Lewis
acid (MX) has been verified by UV spectroscopy [110]. It has been
proposed [ i l l ] that the homopolymerization of the [D + ... :A MX]
complex is responsible for the formation of an alternating copoly-
mer.
The concentration of the complex is dependent upon the monomer,
metal halide, and solvent concentrations, the A/MX ratio, and the
reaction temperature.

(1) A + MX --, ^- A MX

(2) D + A MX * ^ [D+ "A MX]

(3) x[D+ "A MX] »- -(DA)X— + xMX


290 H. SAWADA

At low temperature, the equilibrium in Process (2) should shift to


the right, resulting in a higher complex concentration. Increasing
the temperature may shift the equilibria in Processes (1) and (2) to
the left, resulting in a decrease in the concentration of the complex
and an increase in the concentration of free donor monomer as well
as acceptor monomer [112].
Condensation copolymerization proceeds in the same way as in
the formation of the simple homopolymers, but the chains now con-
sist of mixed units. It is often assumed that at first the reagents
condense less randomly according to the relative reactivities of the
groups, and that the resulting low or medium molecular weight
products rearrange as the reaction proceeds, finally giving a ran-
domly arranged copolymer.
To treat the problem quantitatively, we shall now introduce a new
parameter v [113]:

v = [P12 - (P I2 ) rlad ]/[(P 12 ) BM - - ( P 1 2 ) r a J (68)

where P 12 represents the number of Mj-M2 pairs in the actual co-


polymer; (P12)rMd represents the number of Mx-M2 pairs which
would exist in a copolymer of the same composition if the copoly-
mer were random; and (P 12 ) mai represents the maximum number
of Mi-M., pairs which would exist if complete order occurred.
(P 12 ) raad is given by

(P 12 ) rmd = 2X1X2N0 (69)

where Xx is the mole fraction of M t in the copolymer, X2 is the mole


fraction of M2 in the copolymer, and No is the total number bonds.
In the case of the alternating copolymer, where X1 = X2 = 1/2,

(Pi 2 )»« = No (70)


From Eq. (69) for this copolymer

(P i a ) r i a d = N o /2 (71)

and hence from Eqs. (68), (70), and (71) we obtain

i>=2P12/N0-l (72)

Then the number of Jv^-M., pairs can be expressed

P 12 = N0(l + v)/2 (73)


THERMODYNAMICS OF POLYMERIZATION. VI 291

Out of a total number of bonds No, the fraction of bonds of the


-Mj type is
(74)

Consequently, the fraction of bonds which are like bond is

fu + f22 = 1 - fia = (1 - v)/2 (75)


Assume that the enthalpy change on bond formation is the same
in an interchange reaction and in a polycondensation; then the en-
thalpy changes on formation of M^^ M2M2, M2Mlt and bonds
are denoted by H n , H22, H21, and Hl2, respectively, and the numbers
of such bonds by P u , P22, P 21 , and P 12 . For a system with an equal
number of IV^ and M2 monomer units, P u = P22. If the average en-
ergy of like bonds (Hu + H22)/2 is greater than that of unlike bonds
(H12 + H 2I )/2, we have the simple energy level system indicated in
Fig. 8. It should be noted that any gain H on the part of a system is
considered to be positive and any loss negative. [Since heat is evolved
during the polymerization, polycondensation reactions are exo-
thermic (Hn < 0, H u < 0).]

AH

Fig. 8. Two level systems with probabilities P,, P2. Reprinted by permission of John Wiley
& Sons from Ref. 113.

If the change in entropy for the reaction is zero, in thermal equi-


librium, the relative probability of occupancy of the upper energy
state Px to that of the low energy state P 2 , is given simply by Boltz-
mann statistics
P 1 /P 2 = exp(-AH/kT)
where k is Boltzmann's constant.
292 H. SAWADA

The probability of occupancy of the upper energy state is

Pi = (P» + P22)/NO = fu + f22


Similarly the probability of the low energy state is
=
*2 "l2 /"o ~ 12

From Eqs. (74) and (75),


(1 - v)/{l + v)= exp(-AH/kT) (76)
where
AH = (Hlt + H22)/2 - (H12 + H21)/2
and so
v = {1 - exp(-AH/kT)}/{l + exp(-AH/kT)} (77)
In Fig. 9, v from Eq. (77) is plotted vs temperature. If AH is
positive, v will be positive and the monomer units will tend to alter-
nate in the polymer. If AH is negative, v will be negative and the
monomer units will tend to cluster in 'blocks" in the copolymer.
If AH is positive at 0°K, v is predicted to be equal to —1, and at
high temperatures to approach zero. Therefore, in any case, it is
observed that randomness becomes more predominant as the tem-
perature is elevated.
In general, melt polymerization methods yield condensation co-
polymers with random distributions. In contrast, the method of
interfacial polycondensation yields condensation copolymers rang-
ing in order from alternating to block arrangements, since no re-
distribution of groups will take place because of the low polymeriza-
tion temperature (as shown in Fig. 9).
In the case of equal numbers of Mt and M2 monomer units and
equal probabilities for the formation of M ^ and M2M2, and M1M2
and M-JMJ bonds, the molar heat of condensation copolymerization
is given by [114]

H = (H u + H22 + H12 + H 21 )/4 - fAH/2 (78)


where
AH = {(H u + H22) - (H12 + H 21 )}/2
THERMODYNAMICS OF POLYMERIZATION. VI 293

• 1-0

T.'K

-10

Fig. 9. 7-Parameter vs temperature. Reprinted by permission of John Wiley & Sons from
Ref. 113.

H u , H22, H12, and H21 are the enthalpy changes per mole of elementary
reactions.
Substituting Eq. (77) into Eq. (78), we obtain

H = (H u +H22 + H12 + H21)/4 - AH{1 - exp(-AH/RT)}/


2{1 + exp(-AH/RT)}
= Ho - vAH/2 (79)

where

Ho = +H
22

In Fig._10, H from Eq. (79) is plotted vs temperature. _


If AH is positive at 0°K, H is predicted to be equal to (H1Z + H2i)/2
and at high temperatures, to approach_Ho. In this case v is 1, indi-
cating an alternating copolymer. If AH is negative at a temperature
of 0°K, v will be —1 and clustering will occur. Therefore, at 0°K,
H is predicted to be equal to (H u + H22)/2 and, at high temperature,
to approach Ho. If a copolymer is cooled rapidly from a high to a
low temperature, a metastable state may be produced in which a
nonequilibrium disorder is ^frozen" in the structure. In such a
case H has a large value, even at low temperatures.
294 H. SAWADA

Fig. lO^Molar heatof. condensation copolymerization as a function of temperature. A =


(Hu + H22)/2; B = (HI2 + H2,)/2. Reprinted by permission of John Wiley & Sons from
Ref. 114.

Acknowledgment
The author wishes to thank Prof. K. F . O'Driscoll for many valu-
able criticisms of this manuscript.
References

[1] H. Sawada, J. Macromol. Sci.—Revs. Macromol. Chem., CIO, 293 (1974).


[2] T. Alfrey, / Polym. Set, 4, 221 (1949).
[3] M. Suzuki, H. Miyama, and S. Fujimoto, Ibid., 31, 212 (1958);./. Chem. Soc.
Japan, 79, 607 (1958).
[4] K. Horie, I. Mita, and H. Kambe, J. Polym. Sci., A-l, 7, 2561 (1969).
[5] H. Miyama, and S. Fujimoto,Ibid., 54 s32 (1961);M. Suzuki, H. Miyama, and
S. Fujimoto, Bull. Chem. Soc, Japan, 35, 60(1962).
[6] M. Suzuki, H. Miyama, and S. Fujimoto, Ibid., 35, 57 (1962).
[7] L. K. J. Tong and W. 0. Kenyon, J. Amer. Chem. Soc, 71,1925 (1949).
[8] R. M. Joshi,Makromol. Chem., 66,114 (1963).
[9] K. J. Ivin, W. A. Keith, and H. Mackle, Trans. Faraday Soc, 55, 262 (1959).
[10] F. S. Dainton, J. Diaper, K. J. Ivin, and D. R. Sheard, Ibid., 53,1269 (1957).
[ 11 ] C. T. Mortimer, Reaction Heats and Bond Strengths, Pergamon, New York,
1962, p. 95.
[12] R. D. Snow and F. E. Ftey, Ind. Eng. Chem., 30,176 (1938); J. Amer. Chem.
Soc, 65, 2417(1943).
THERMODYNAMICS OF POLYMERIZATION. VI 295

[13] F. S. Dainton and K. J. Ivin,Nature, 162, 705 (1948);Proc. Roy. Soc, A, 212,
207 (1952).
[14] R. E. Cook, F. S. Dainton, and K. J. Ivin, J. Polym. Sci., 26, 351 (1957).
[15] R. E. Cook, K. J. Ivin, and J. H. O'Donnell, Trans. Faraday Soc, 61,1887
(1965).
[16] B. H. G. Brady and J. H. O'Donnell, Eur. Polym. J., 4, 537 (1968); Trans.
Faraday Soc, 64, 29 (1968).
[17] J. R. Brown and J. H. O'Donnell, J. Polym. Sci., 10,1997 (1972).
[18] T. Alfrey, Jr. and C. C. Price, Ibid., 2, 101 (1947).
[19] T. Alfrey, Jr. and L. J. Young, in Copolymerization (G. E. Ham, ed.), Wiley-
Interscience, New York, 1964, Chap. 11.
[20] T. C. Schwan and C. C. Price, J. Polym. Sci., 40, 457 (1959).
[21] T. Kagiya and Y. Sumida, Polym. J., 1,137 (1970).
[22] H. Sawada, J. Macromol. Sci-Revs. Macromol. Sci., C3, 370 (1969).
[23] M. G. Evans, J. Gergely, and E. C. Seaman, J. Polym. Sci, 3, 866 (1948).
[24] M. Chorton and A. J. Capata, Ibid., A, 2,1321 (1964).
[25] M.C.Shen,/6irf.,B, 1,11(1963).
[26] N. Kawabata, T. Tsuruta, and J. Furukawa,Makromol. Chem., 51, 80 (1962).
[27] T. Yonezawa, K. Hayashi, C. Nagata, S. Okamura, and K. Fukui, J. Polym.
Sci., 14, 312 (1954); K. Hayashi, T. Yonezawa, C. Nagata, S. Okamura, and
K. Fukui, Ibid., 20, 537 (1956); G. S. Levinson,Ibid., 60,43 (1962).
[28] L. P. Hammett, J. Amer. Chem. Soc, 59,96 (1937);Pftysica/ Organic
Chemistry, McGraw-Hill, New York, 1940.
[29] C. Walling, E. R. Briggs, K. Wolfstern, and F. R. Mayo, J. Amer. Chem. Soc,
70,1537 (1948).
[30] K. Tsuda, S. Kobayashi, and T. Otsu, J. Polym. Sci.,A-l, 6, 41 (1968); J.
Macromol. Sci.-Chem., Al, 1025 (1967).
[31] M. Kinoshita, T. Irie, and M. Imoto,Makromol. Chem., 110, 47 (1967).
[32] M. Imoto, M. Kinoshita, and M. Nishigaki,/&;</., 94, 238 (1966);M. Kinoshita
and M. Imoto, J. Chem. Soc Japan, Ind. Chem. Sect., 68, 2454 (1965).
[33] T. Yamamoto and T. Otsu, Org. Syn. Chem. Japan, 23, 643 (1965);T.
Yamamoto, T. Otsu, and M. Imoto, J. Chem. Soc. Japan, Ind. Chem. Sect.,
69,990(1966).
[ 34 ] T. Otsu, in Progress in Polymer Science Japan, Vol. 1 (M. Imoto and S. Onogi,
eds.), Kodansha, Tokyo, 1971, p. 34.
[35] C. Bamford, A. D. Jenkins, and R. Johnston, Trans. Faraday Soc, 55,418 (1959);
59, 530 (1963).
[36] J. Furukawa andT. Tsuruta, J. Polym. Sci, 36,275 (1959).
[37] A. Cammarata and S. J. Yan, Ibid., A-1,8,1303 (1970).
[38] A. V. Chernobaiand Z. K. Zelichenko, Vyskomol. Soedin.,A, 11(7), 1631 (1969).
[ 39 ] R. W. Taft, Jr., in Steric Effects in Organic Chemistry (M. S. Newman, ed.), Wiley,
New York, 1956, p. 556.
[40] G. G. Cameron and G. D. Kerr,Eur. Polym. J.,3,1 (1969).
[41] K. Chikanishi and T. Tsuruta,Makromol. Chem., 73, 231 (1964); 81, 211 (1965).
[42] N. Yamashita, S. Kunitani, S. Komiyama,M. Yoshihara, and T. Maeshima, J.
Polym. Sci,B, 10, 675 (1972).
[43] K. Yasufuku, S. Hirose, S. Nozakura, and S. Murahashi, Bull. Chem. Soc. Japan,
40, 2139 (1967).
[44] T. Otsu,T. Ito, and M. Imoto, J. Polym. Sci.,B, 3,113 (1965).
296 H. SAWADA

[45] B. Yamada and T. Otsu, J. Macromol. Sci.-Chem., A3,1551 (1969).


[46 ] F. M. Lewis, C. Walling, W. Cummings, E. R. Briggs, and F. R. Mayo, J. Amer.
Chem. Soc, 70,1519 (1948).
[47] G. Goldfinger and M. Steidlitz, J. Polym. Sci, 3, 786 (1948).
[48] H. K. Johnston and A. Rudin, J. Paint Technol, 42,435 (1970).
[49] V. E. Meyer, J. Polym. Sci.,A-l, 5,1289 (1967).
[50] R. M. Joshi and S. L. Kapur, J. Sci. Ind. Res., 16B, 379 (1957).
[51 ] A. Rudin and R. G. Yule, J. Polym. Sci., A-l, 9, 3009 (1971).
[52] K. F. O'Driscoll, J. Macromol. Sci.-Chem., A3, 307 (1969).
[53] A. D. Jenkins and M. G. Rayner, Eur. Polym. J., 8, 221 (1972).
[54] A. Yamada and M. Yanagita, J. Macromol. Sci.-Chem., A4,1841 (1970).
[ 55 ] C. Walling, Free Radical in Solution, Wiley, New York, 1957, p. 35.
[56] C. C. Price and J. G. Walsh, J. Polym. Sci., 6, 239 (1961).
[57] R. M. Joshi,Ibid., 60, 556 (1962); J. Sci. Ind. Res., 18B, 279 (1959).
[58] R. Kerber.il/aftromo/. Chem., 96, 30 (1966).
[59] G. Saini, A. Leoni, and S. Franco,Ibid., 144, 235 (1971).
[60] G. Saini, A. Leoni, and S. Franco,Ibid., 147,213 (1971).
[61 ] G. Saini, A. Leoni, and S. Franco, Ibid:, 146,165 (1971).
[ 61 ] L. M. Minsk, C. Kotlarchik, and R. S. Darlak, J. Polym. Sci., Polym. Chem. Ed.,
11,353(1973).
[63] T. Ito and T. Otsu, J. Macromol. Sci.-Chem., A3,197 (1969).
[64] A. V. Ryabov, Y. D. Semchikov, L. A. Smirnova, N. N. Slavitskaya, N. L. Khvatova,
and V. N. Kashayeva, Vysokomol. Soedin., A, 13,1414 (1971).
[65] V. G. Markert and H. Pennewiss, Angew. Makromol. Chem., 11, 53 (1970).
[66] Y. Landler,/ Polym. Sci., 8, 63 (1952).
[67] K.F. O'Driscoll and I. Kuntz,/6jtt, 61,19 (1962).
[68] S. Okamura, N. Kanoh, and T. Higashimura,Makromol. Chem., 47, 35 (1961).
[69] K. F. O'Driscoll,,/. Polym. Sci., 57, 721 (1962).
[70] F. Dawans and G. Smets,Makromol. Chem., 59,163 (1963).
[71] K. F. O'Driscoll, T. Higashimura, and S. Okamura, Ibid., 85,178 (1965).
[ 72 ] H. Sawada, J. Macromol. Sci.-Revs. Macromol. Chem., C3 322 (1969).
[73] R. M. Joshi, in Vinyl Polymerization, Part 1 (G. E. Ham, ed.), Dekker, New York,
1967, p. 500.
[ 74 ] P. R. Wells, Linear Free Energy Relationships, Academic, New York 1968, p. 1.
[75] H. Sawada, J. Polym. Sci., A-l, 5,1383 (1967).
[76] L. A. Wall.TSitt, 2, 542 (1947).
[77] N. Kawabata, T. Tsuruta, and J. Furukawa,Bull. Chem. Soc. Japan, 36, 905 (1963).
[78] C. G. Overberger, L. Arond, and D. Tanner, J. Amer. Chem. Soc, 73, 5541 (1951);
C. G. Overberger, R. Ehrg, and O. Tanner, Ibid., 76, 772 (1954); C. G. Overberger
and V. G. Kamath,Ibid., 85, 446 (1963).
[79] T. Masuda and T. Higashimura, J. Macromol. Sci.-Chem., A5(3), 549 (1971).
[80] K. F. O'Driscoll, T. Yonezawa, and T. Hogashimura,Ibid., 1,17 (1966).
[81] K. F. O'Driscoll, J. Polym. Sci., A, 3,2223 (1965).
[82] C. G. Overberger, L. H. Arond, D. Tanner, J. J. Taylor, and T, Alfrey, Jr., J. Amer.
Chem. Soc, 74,4848 (1952).
[83] T. Tsuruta, J. Japan. Chem. (Kagaku no Ryoiki), 8, 209 (1954);BuH. Inst. Chem.
Res. Kyoto Univ., 32,149 (1954).
THERMODYNAMICS OF POLYMERIZATION. VI 297

[84] Y. Okamoto and H. C. Brown, J. Org. Chem., 22,485 (1957).


[85] J.F.Dunphy and S. C. Marvel, J. Polym. Sci., 47,1 (1960).
[86] A. V. Tobolsky and R. J. Boudreau, Ibid., 51, s53 (1961).
[87 ] A. Mizote, T. Tanaka, T. Higashimura, and S. Okamura, Ibid., 3, 2567 (196&).
[88] G. Natta, F. Danusso, and D. Sianesi, Makromol. Chem., 30, 238 (1959).
[89] G. Natta,Experientia Suppi, 7, 21 (1957).
[90] M. Shima, D. N. Bhattacharyya, J. Smid, and M. Szwarc, J. Amer. Chem. Soc,
85,1306 (1963).
[91] J. Rehner, Jr., R. L. Zapp, and VV. J. Sparks, J. Polym. Sci., 11, 21 (1953).
[92] M. Imoto and K. Saotome./fciU, 31, 208 (1958).
[93] G. R. Brown and D. C. Pepper, J. Chem. Soc, 1963, 5930.
[94] T. Masuda and T. Higashimura,Polym. J., 2, 29 (1971).
[95] C. G. Overberger, L. H. Arond, and J. J. Taylor, J. Amer. Chem. Soc, 73, 5541
(1951).
[96] J. Furukawa, T. Taniguchi, and E. Kobayashi, Presented at the 21st Symposium
on Polymer Science, Society of High Polymers, Japan, 3A11,1972.
[97] S. Iwatsuki, N. Takigawa, M. Okada, Y. Yamashita, and Y. Ishii, J. Polym. Sci.,
B, 2, 549 (1964).
[98] Y. Yamashita, T. Tsuda, M. Okada, and S. Iwatsuki,Ibid., A-l, 4, 2121 (1966).
[99] M. Okada, N. Takikawa, S. Iwatsuki, Y. Yamashita, and Y. Ishii,Makromol. Chem.,
82,16 (1965).
[100] S. Aoki, Y. Harita, Y. Tanaka, H. Handai, and T. Otsu, J. Polym. Sci., A-l, 6,
2585 (1968).
[ 101 ] Y. Tanaka, J. Macromol. Sci.-Chem., Al(6), 1059 (1967)
[102] F. S. Dainton, T. R. E. Devlin, and P. A. Small, Trans. Faraday Soc, 51,1710
(1955).
[103] T. Kagiya, Y. Sumida, and T. Inoue,Polym. J., 1, 312 (1970).
[ 104 ] V. A. Ponomarenko, A. M. Khomutov, S. I. I'lchenko, A. V. Ignatenko, and N. M.
Khomutova, Vysokomol. Soedin.,A, 13,1551 (1971).
[105] V. A. Ponomarenko, A. M. Khomutov, S. I. I'lchenko, A. V. Ignatenko, Ibid., A13,
1546 (1971).
[ 106] Y. Yamashita, T. Asakura, M. Okada, and K. Ito, Makromol. Chem., 129,
1 (1969).
[107] Y. Yamashita, T. Inoue, G. Hattori, and K. Ito,/6iU, 151, 91 (1972).
[108] A. M. North and D. Richardson,Polymer, 5, 333 (1964).
[ 109] C. Aso, S. Tagami, and T. Kunitake, J. Polym. Sci., A-l, 8,1323 (1970).
[110] T. Ikegami and H. Hiral, Ibid., A-l, 8,195, 463 (1970).
[Ill] N. G. Gaylord and A. Takahashi, J6/d.,£, 6, 743 (1968).
[112] N. G. Gaylord and B. Patnaik,/&!</.,£, 8,411(1970).
[113] H. Sawada,/6iU,B, 1,659(1963).
[114] H. Sawada,/6W.,B, 2, 507 (1964).

You might also like