You are on page 1of 711

COMPUTATIONAL MATERIALS SCIENCE

OF
POLYMERS

COMPUTATIONAL
MATERIALS SCIENCE
OF
POLYMERS
A NDREY A LEKSANDROVICH ASKADSKII

CAMBRIDGE INTERNATIONAL SCIENCE PUBLISHING

Published by
Cambridge International Science Publishing
7 Meadow Walk, Great Abington, Cambridge CB1 6AZ, UK
http://www.cisp-publishing.com
First published January 2003
A A Askadskii
Cambridge International Science Publishing

Conditions of sale
All rights reserved. No part of this publication may be reproduced or transmitted
in any form or by any means, electronic or mechanical, including photocopy,
recording, or any information storage and retrieval system, without permission
in writing from the publisher
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library

ISBN 1 898326 6 22

Production Irina Stupak


Printed by Antony Rowe Ltd, Chippenham, Wiltshire, Great Britain

About the Author


Andrey Aleksandrovich Askadskii is a Professor of Chemistry at the Institute of Organo-Element Compounds of the Russian Academy of Sciences.
He holds M.S. in Civil Engineering from the Moscow Civil Engineering Institute
(1959), M.S. in Chemistry from the Mendeleev Institute of Chemical Technology
(1962) and Ph.D. in Physics of Polymers (1968).
The main scientific interests of the author are: the development of a
physical approach to the quantitative evaluation of the physical properties
of linear and network polymers on the basis of their chemical structure;
development of computer programs for calculating the properties of polymers and low-molecular liquids and also computer synthesis of polymers with
the required properties; experimental examination of the structure of properties
of heat-resistant aromatic polymers of different grades; development of new
methods of experimental and theoretical analysis of the relaxation properties of polymer materials; production of new types of polymers; production
and examination of electrically conducting polymer materials on the basis
of heat-resistant polymers and organo-element compounds; development of
gradient polymer materials with a variable modulus of elasticity within the
limits of the same material and retaining elastic (not viscoelastic) properties at any point of the gradient material.
Prof Askadskii is the author of more than 400 scientific studies and
20 books, six of which have been published abroad.

Contents

Preface
Introduction
Chapter I. Brief information on types of polymes and their chemical structure
Chapter II. Packing of macromolecules and polymers density
II.1. Increments method and basic physical assumption
Chapter III. Temperature coefficient of volumetric expansion
Chapter IV. Glass transition temperature of polymers
IV.I. Thermomechanical and other methods of evaluation of the glass
transition temperature of polymers
IV.2. Mechanism of glass transition
IV.3. Calculation of the glass transition temperature of linear polymers
IV.4. Influence of plasticization on the glass transition temperature of polymers
IV.5. Calculation of the glass transition
Chapter V. Temperature of transition into the viscous flow state for amorphous
polymers
V.1. Estimation of temperature of transition into the viscous flow state of
polymers
V.2. Dependence of Newtonian viscosity on molecular mass of polymer in a
wide range of its change
Chapter VI. Melting point of polymers
Chapter VII. Temperature of onset of intense thermal degradation of polymers
Chapter VIII. Optical and opto-mechanical properties of polymers
VIII.1. Refractive index
VIII. 2. Stress-optical coefficient
Chapter IX. Dielectric constant of polymers and organic solvents
Chapter X. Equilibrium rubber modulus for polymer networks
X.1. Calculation of the equilibrium modulus
X.2. Heteromodular and gradient-modulus polymers
Chapter XI. Description of relaxation processes in polymers
XI.1. Stress relaxation
XI. 2. Sorption and swelling processes
Chapter XII. Solubility of polymers
XII.1. Specific cohesive energy of organic liquids and polymers. Hildebrand
solubility parameter
XII.2. Solubility criterion
XII.3. Influence of molecular mass and degree of macromolecule orientation
on solubility
Chapter XIII. Surface properties of organic liquids and polymers
XIII.1. Surface tension of organic liquids
XIII.2. Surface tension of polymers
Chapter XIV. Miscibility of polymers
Chapter XV. Influence of the end groups on the properties of polymers
Chapter XVI. Thermophysical properties of polymers
XVI.1. Heat capacity
XVI.2. Thermal diffusivity and heat conductivity

3
9
16
16
58
67
67
88
108
322
343
385
385
388
398
408
418
418
426
445
456
456
466
475
475
497
504
504
509
520
527
528
536
547
555
562
562
564

Chapter XVII. Molecular design and computer synthesis of polymers with


predermined properties
Appendix 1. Examples of solution of direct problems of polymers synthesis
Appendix 2. Examples of solving the reverse problem of polymer synthesis
Appendix 3. The example of solving the complex problem analysis of the
chemical structure of phenol formaldehyde resin
Appendix 4. Application of the approach to multicomponent copolymers
Appendix 5. Influence of strong intermolecular interaction occurring between
two dissimilar polymers on their miscibility
Appendix 6. On formation of super-molecular structure in amorphous polymers
1. Scheme of formation of the super-molecular structure
2. Calculation method of evaluation of dimensions of elements of supermolecular structure of polymers
3. Phase state of polymers as a result of formation of the super-molecular
structure by one-cavity bond hyperboloids
References
Index

567
589
602
607
621
625
645
645

653
669
689

PREFACE
Published in the journal Chemistry and Life, No. 2, 1981 was the article by
me, titled by the editor as Atom plus atom plus thousand atoms. This article
discussed the possibility of calculating some physical properties of polymers on the
basis of the chemical structure of the repeat unit (it was then possible to calculate
properties of linear polymers only). In conclusion of the article, titled A little
fantasy, it was written: Therefore, many properties of polymer can be predicted, if
nothing except the structural formula of the appropriate monomer is known. It is a
great progress: nowadays already, such calculations allow chemists to be drawn
away from heavy duty to synthesize hopeless monomers. Formerly, under empirical
selection of materials, many of such monomers had to be synthesized. Nevertheless,
calculations are to be made manually still. Moreover, when they are translated into
the machinery language, chalk and blackboard traditional for any chemical dispute
can be substituted by an electronic pencil. A chemist will draw a formula of the
suggested monomer on the screen by it, and the computer will answer immediately if
it is useful or not to synthesize it. Another opposite task seems to be much more
absorbing. If the computer is able to calculate properties by structural formulae,
apparently, it may be taught, vice versa, to calculate the formula of a suitable
monomer (or several formulae to choose) by any, even contradictory set of properties,
given to it. In this case, it will be able to substitute the chemist in his most problematic
part of work, one is able to succeed in on the basis of experience, intuition and luck.
That was a fantasy, and it could be hardly imagined that these ideas would be realized
at any time in neat future. However, events were developing very fast, especially after
appearance of high-power personal computers. Before discussing stages of this great
work, methods of the quantitative estimation of polymer physical properties must be
presented in brief performed on the basis of their chemical structure. At the present
time, there are three main approaches to this estimation. One of them, developed by
Van Krevelen [214], is based on the idea of so-called group contributions, according
to which the simplest empirical expressions of the additive type are written down, the
present group, existing in different polymeric units, making one and the same
contribution to the calculated characteristic (for example, glass transition temperature,
melting, etc.). As the author states, this is just an empirical approach, which allows
the physical properties of many of linear polymers to be calculated with high
accuracy.
Another approach, being developed for a long time by the author of this
preface in company with Yu.I. Matveev [28, 128] is semi-empirical. According to it,
equations for calculation of the physical properties are deduced on the basis of ideas
of physics of solids, and calibration of the method is performed with the help of
physical characteristics of polymeric standards, the properties of which are studied
well. Consequently, parameters of equations possess a definite physical sense (energy
of dispersion interaction, energy of strong intermolecular interaction, including
hydrogen bonds, Van-der-Walls volume, etc.). Application of this approach makes
possible estimation with enough accuracy of many physical characteristics of
polymers (about 60 up to now). Therefore, the number of polymers of various
structures is unlimited.
The third approach developed by J. Bicerano [133] has appeared recently. It is
based on the so-called coherence indexes, reduced in practice to a search for various

correlations of physical properties with many rules of obtaining coefficients of


correlation dependencies.
Discussed in the present monograph are principles of the approach, developed
by A.A. Askadskii and Yu.I. Matveev, special attention being paid particularly to
computer realization of the current calculation method for physical properties of
polymers. The first computer software has been composed by E.G. Galpern, I.V.
Stankevich and A.L. Chistyakov - investigators of quantum chemistry laboratory of
A.N. Nesmeyanov Institute of Organo-Element Compounds, RAS. Initially, computer
synthesis of polymers by this software was performed from so-called large
procurements representing residues of monomers, involved into the synthesis
reaction. In the second variant, computer synthesis was performed from smallest
procurements, from which the repeat unit of the polymer was constructed. This
broadens significantly capabilities of the software for solving both direct (calculation
of the polymer properties from its chemical structure) and reverse task (computer
synthesis of polymers with preliminarily programmed /assigned/ properties, the
ranges of which were set in the computer), because the amount of synthesized
olymers has increased sharply. Then principally new software was composed by A.F.
Klinskikh, in which chemical structure of the repeat unit was constructedfrom
atoms. Thus, the user needs just to depict chemical structure of the polymer on the
computer screen as chemist does it on the paper, and computer lays out all physical
properties of polymers, involved in the software (all about 60). This software also
provides for calculation of a sequence of properties of low-molecular weight organic
compounds, as well as, which is very important, properties of polymeric networks.
Solution of the reverse task is also provided. Of special importance is the possibility
to calculate properties of copolymers and their mixtures, to predict solubility and
compatibility of polymers, to construct dependencies of properties on temperature,
molecular mass, crystallinity degree, microtacticity (of special importance are
dependences of glass transition temperature and temperature of transition into the
viscous flow state on molecular mass).
It stands to reason that not all the problems are solved. Accuracy of the
calculation and various predictions of polymers behavior at dissolution and mixing
with each other must be increased, calculation schemes to estimate new properties of
polymers must be developed, and their computer realization must be performed, etc.
It is obvious that the present monograph possesses some drawbacks. The
authors will be thankful for any notes on the point of the book.

INTRODUCTION
As mentioned above, the approach to estimation of the physical properties of
polymers, discussed in the monograph, is semi-empirical. When estimating the
thermal characteristics of polymers, such as glass transition temperature, melting
point, it is supposed that the repeat unit is composed of a set of anharmonic oscillators
representing atomic pairs, linked by intermolecular physical bonds. The critical
temperature of this set of anharmonic oscillators is that determines the abovementioned two transition temperatures. The thermal expansion coefficient is also
closely related to these characteristics. In the case of a characteristic as the
temperature of the onset of intensive thermal degradation, the polymeric unit is
considered as a set of anharmonic oscillators representing atomic pairs, linked by
chemical bonds. The critical temperature of such a set of oscillators characterizes the
temperature of the onset of intensive thermal degradation at the given rate of heating
(clearly at a different rate of heating, the temperature of the onset of intensive thermal
degradation will be different, i.e. kinetic effects play a significant role in this case). At
first glance, it may seem strange that thermal degradation is considered here not as a
kinetic, which is conventional, but as an original phase transition, at which, however,
the initial substance cannot be obtained from the products of thermal decomposition
by simple cooling down.
Equations for calculating other physical characteristics are based on physical
approaches, discussed in detail below, and we will not consider them in this part.
Common for all these equations is summarizing the sequence of atomic
constants, which characterize contributions to the energy of intermolecular
interaction, chemical bonds energy, Van-der-Waals volume, etc. Strictly speaking, the
present approach cannot be named additive in the common sense of the word, because
the calculated properties are not additive in relation to atoms and groups, which
compose the repeat unit of polymer.
Here additivity is applied to the characteristics which are really additive (Vander-Waals volume, molecular mass, intermolecular interaction energy, etc.). The
approach being described allows calculation of their properties of the unlimited
number of polymers and conduction of the computer synthesis of polymers with
assigned properties with the help of software created and described in the monograph
that is not possible using other existing programs.
As mentioned above, the approach discussed in the monograph is semiempirical, calibration of the method being based on the so-called polymeric standards,
the properties of which are studied in detail and common. Let us consider the essence
of calibration on an example of the equation calculating glass transition temperature
of a linear polymer, Tg:

Tg =

Vi
i

ai Vi + b j
i

where ai are atomic constants; bj are constants bound to the energy of strong
intermolecular interaction (dipole-dipole, hydrogen bonds), occurred between
polymeric chains at the sacrifice of polar groups existing in them; Vi is the Vani

der-Waals volume of the polymer repeat unit, summarized from Van-der-Waals


volumes of atoms participating in the composition of the unit.
Reduce the equation to the following view:
1

ai Vi + b j = T Vi .
i

g i

Basing on this equation, the excessive system of linear equations is composed


as follows:

1
a1V1,1 + a2 V1,2 + ... + an V1, n + 1b1 + 1b2 + ... + 1bk =
Vi ;

Tg ,1 i

Vi ;
a1V2,1 + a2 V2,2 + ... + an V2, n + 2b1 + 2b2 + ... + 2bk =

Tg ,1 i
2

.......... .................... .................... .................... .................... .................... ...........

1
Vi .
a1Vm,1 + a2 Vm,2 + ... + an Vm, n + m b1 + m b2 + ... + m bk =

Tg ,1 i

m
Then the matrix of coefficients at the unknowns of this excessive system of
equations:
'
'
 =

'P



 



'
'

 

 

'P



' Q
' Q





'P Q P

P


and the column matrix of free terms of these equations




'L


%J L


'L

 = %J L



'L

%J P L
P




are composed.
Further on, a transposed matrix is composed and multiplied by the initial
one A, as well as by the column matrix B. All this results in obtaining a

canonic system of equations. This canonic system is solved, for example, by the
Gauss method. The whole procedure of calibration is performed by standard software.
Without considering features of such regressive analysis, let us note only that
polymers, selected for calibrating the method, must possess experimental values of
analyzed physical characteristics in broadest range, and the chemical structure of
polymeric standards must be sufficiently different. Usually, an excessive system
composed of 300 equations is to be solved, which corresponds to 3040 polymers.
Next, the properties of other polymers are calculated from the coefficients obtained.
In this case, the energy of weak dispersion interaction, strong dipoledipole
interactions and hydrogen bonds, their relative part and many other physical
parameters of the system are determined.
We are coming now to a brief description of the contents of individual
chapters of the monographs.
The first chapter discusses the data of modern classification of polymers and
their chemical structure. Of the outstanding importance, induced by the features of the
chemical structure and the application field, are interpolymers, dendric and staircase
(ladder) polymers.
The second chapter discusses the approach to computerized materials
technology of polymers on the atomicolecular level, based on the method of
increments. The increments of various atoms and main groups of them are calculated.
The main physical ideas about structure of macromolecules of polymers and
parameters determining it are displayed. The method for calculating such an important
characteristic of the polymer structure, as the coefficient of molecular packing, is
given. A connection between the free volume of the polymer, the coefficient of
molecular packing and parameters of its porous structures is established. For
experimental determination of characteristics of the microporous structure of
polymers, the method of positron annihilation, the application of which indicated
structural changes in polymers in their relaxation, is used.
With consideration of weak dispersion and strong (dipoledipole and
hydrogen bonds), the third chapter gives formulae for calculating the thermal
coefficient of the volume expansion in dependence on the chemical structure of the
polymer. In this case, the type of atoms in the polymeric chain and type of the
intermolecular interaction are estimated by a limited number of corresponding
increments, numerical values of which are determined.
The fourth chapter describes in detail the thermomechanical method of
determination of the glass transition temperature and fluidity of polymers, features of
interpreting thermomechanical curves for amorphous and crystalline polymers are
analyzed, the calculation method of determination of the mechanical segment from
the chemical structure of the polymer is displayed. Two main concepts of the
mechanism of vitrification processes of polymers, relaxation and intermolecular, are
discussed. The atomistic approach which is more universal than the widespread socalled group contributions method to calculation of polymer properties from their
chemical structure, is considered. This approach was used for deriving an analytical
expression to calculate the glass transition temperature of linear and network
polymers from their chemical structure. The influence of types of linear polymers
branching and the number of units between cross-link points, type and structure of
these points, existence and type of the network defects for network polymers on the
glass transition temperature of the polymers is analyzed.
Given in the fifth chapter is the method for calculating the fluidity temperature
of amorphous polymers and the temperature range of the rubbery state of polymers

from their chemical structure, and conditions of appearance of the rubbery state in a
polymer depending on its molecular mass, as well, which is important for processing
of polymers.
The sixth chapter describes two approaches to calculating the melting point of
polymers from the chemical structure of the repeat unit. The first approach is based on
the experimental fact of closeness in parts of the empty volume in melting of a
crystalline polymer and in transition of an amorphous polymer of the same structure
from the glassy-like into the high-elastic state. The second approach is based on the
consideration of the repeat unit of a polymer as a selection of anharmonic oscillators.
Discussed in the seventh chapter is the most important characteristic of
thermal resistance of polymers initial temperature of their intensive thermal
degradation. The formula to calculate this temperature based on the chemical structure
of the polymer was deduced, and necessity to take into account the resulting products
of thermal degradation which starts with the decay of end groups in polymer
macromolecules, are indicated.
In the eighth chapter, LorenzLorentz equations are used for deriving
equations for calculation of the refractive index of polymers and copolymers from
their chemical structure. To obtain the stress-optical coefficient, empirical and semiempirical approaches are established, in which the contribution of each atom and the
type of intermolecular interaction are estimated by an appropriate increment. Using
the dependencies obtained for the stress-optical coefficient on the chemical structure
of the repeat unit of the polymer, the contribution of various atoms and polar groups
to the value of this coefficient is estimated, and a polymer with the properties unique
for the method of dynamic photo-elasticity is proposed.
The ninth chapter displays a scheme for calculating the dielectric constant of
polymers and organic liquids with respect to their chemical structure which is
important for both synthesis of polymers with the required dielectric constant and
prognosis of polymer solubility in organic liquids. Taking into account not only the
contribution of various polar groups to the dielectric constant of polymers and liquids,
but also different contributions of a polar group in the present class of liquids resulted
in the previously unobtainable agreement in the experimental and calculated values of
the dielectric constant for a broad spectrum of organic polymers and liquids.
Based on the notion of network polymers as an elastic and rotationalisomeric
subsystem and taking into account its structure as linear fragments and cross-linked
points, the tenth chapter indicates the deduction of formulae for calculating the
equilibrium rubbery modulus and molecular mass of a linear fragment between
neighboring cross-linked points. Further analysis of the resultant dependencies
allowed the formulation of conditions for obtaining a polymer with unique (unusual)
properties different modulus and gradient polymers characterized by large changes
of the equilibrium rubbery modulus within the same article. Existence of these unique
properties is confirmed experimentally for synthesized network of polyisocyanurates.
The eleventh chapter describes the derivation of analytical expressions for
relaxation memory functions, necessary for determining the stress relaxation and
creep of the polymers. In this case, the production of entropy of a relaxing system is
represented by transition of relaxants (kinetic units of a polymer of different nature)
into non-relaxants by means of their interaction or diffusion, the mechanism of
interaction of relaxants in stress relaxation being found predominant. The apparatus
created for description of relaxation events in polymers is applied in description of
sorption and swelling processes. Thus, contrary to stress relaxation, the mechanism of
relaxants diffusion is predominant in sorption.

The twelfth chapter is devoted to the problem of increasing the accuracy of


prediction of polymer solubility in organic liquids. It is shown that the predictive
ability of the solubility criterion, calculated with respect to the chemical structure of
the polymer and the solvent, sharply increases with consideration for the type of
supermolecular structure of the polymer and the degree of its polymerization.
Based on the chemical structure of the matter, the thirteenth chapter gives a
calculation method for the most important property of organic liquids and polymers,
i.e. surface tension. Contrary to the additive scheme for summation of parachors
which characterizes the contribution of separate atoms to the surface tension, the
approach developed allows estimation of the contribution of polar groups and specific
intermolecular interaction to the surface tension value and connection of it with the
solubility parameter and density of cohesion energy in substances.
Invoking the idea of solubility of a single homopolymer in another one, the
fourteenth chapter suggests a criterion for estimating the compatibility of polymers
basing on the data of the chemical structure of separate components. The analysis of
application of the criterion for compatible, partially compatible or incompatible
polymers indicates its high predictive ability.
On the example of the calculation of the Van-der-Waals volume, molar
refraction, heat capacity and other properties of a number of polymers, chapter fifteen
displays the role of the chemical structure of macromolecule end groups and
importance of their calculation in the study of regularities of changes in the polymer
properties on their molecular mass.
The sixteenth chapter indicates a method for calculating the molar heat
capacity with respect to the chemical structure of polymers. The method is based on a
supposition that the contribution of each atom to heat capacity is proportional to its
Van-der-Waals volume. It is noted that the heat capacity, thermal diffusivity and heat
conductivity of polymers depend not only on their chemical structure, but also on the
physical and phase states of the polymeric body.
The seventeenth chapter describes methodological ways of solving the direct
problem of computerized determination of the physical characteristics of polymers
and low-molecular liquids with respect to their chemical structure and the reverse one
computer synthesis of polymers with the given set of properties. These problems
are solved by the methods of fragments and separate atoms. The corresponding
software which allows calculation of more than 50 chemical properties of linear and
network polymers and copolymers, and a number of the most important properties of
low molecular weight liquids, as well, is developed. Discussed is the method of
depicting diagrams of polymer properties compatibility, application of which may
significantly simplify solution of the direct and, especially, reverse problems of
computational materials sciences.
Appendices demonstrate abilities of the approach, described in the
monograph, to determine the properties of some natural polymers (the example of
solving the direct problem of polymers synthesis) with respect to their chemical
structure (Appendix 1); to search for chemical structures of polyetherketones (the
example of solving the reverse problem of polymer synthesis), the properties of which
must lie in a given range (Appendix 2); to solve a mixed problem of polymers
synthesis on the example of analyzing the chemical structure of phenoloformaldehyde
resin, when the direct problem estimation of the properties of the ideal structures
of such resin with respect to their chemical formulae and the reverse one
searching for a combination of structures with which the chemical formula of
phenoloformaldehyde resin obtained provides experimentally observed values of its

properties are solved consecutively (Appendix 3); to analyze the structure and
properties of copolymers, composed of from three to five comonomers (Appendix 4);
and the influence of a strong intermolecular interaction appearing between two
heterogeneous polymers on their compatibility is analyzed (Appendices 5 and 6).

Chapter I. Brief information on types of polymers and their


chemical structure
The very large number of existing polymers may be subdivided into three
main classes forming the basis of the presently accepted classification. The first class
contains a large group of carbochain polymers whose macromolecules have a skeleton
composed of carbon atoms. Typical representatively of the polymers of this class are
polyethylene, polypropylene, polyisobutylene, poly(methyl methacrylate), poly(vinyl
alcohol) and many other. A fragment of a macromolecule of the first of them is of the
following structure
[CH2CH2]n
The second class is represented by a similar large group of heterochain
polymers, the main chain of macromolecules of which contains heteroatoms, in
addition to carbon atoms (for example, oxygen, nitrogen, sulfur, etc.). Numerous
polyethers and polyesters, polyamides, polyurethanes, natural proteins, etc., as well as
a large group of elemento-organic polymers relate to this class of polymers. The
chemical structure of some representatives of this class of polymers is the following:
[CH2CH2O]n
Poly(ethylene oxide)
(polyether);
(CH2)2

O C
O

NH

(C H 2 ) 6

Poly(ethylene terephthalate)
(polyester);

C O
O

NH

(C H 2 ) 4

n
C
O

CH3
Si O
CH3

Polyamide;
n

Polydimethylsiloxane
(elemento-organic
polymer);

Cl
N

Polyphosphonitrile chloride
(inorganic polymer).

Cl
n
The third class of polymers is composed of high-molecular compounds with a
conjugated system of bonds. It includes various polyacetylenes, polyphenylenes,
polyoxadiazoles and many other compounds. The examples of these polymers are:
[CH=CH]n
Polyacetylene

Polyphenylene
n

N N
C

Polyoxadiazole

C
O

10

An interesting group of chelate polymers possessing various elements in their


composition, able to form coordination bonds (usually, they are depicted by arrows),
also relates to this class. The elementary unit of these polymers is often complex, for
example:
O
H3C
O

CH3
O

Zn
O
P

H3C

O
CH3

The most widely used type of material in the large group of polymeric
materials are still the materials based on the representatives of the first class of
polymers which are carbochain high-molecular compounds. The most valuable
materials could be produced from carbochain polymers, for example, synthetic
rubbers, plastics, fibers, films, etc. Historically, these polymers have been
implemented in practice first (production of phenoloformaldehyde resins, synthetic
rubber, organic glass, etc.). Many of carbochain polymers became subsequently the
classic objects for investigation and creation of a theory of the mechanical behaviour
of polymeric substances (for example, polyisobutylene, poly(methyl methacrylate),
poly-propylene, phenoloformaldehyde resin, etc.).
Subsequently, materials based on heterochain polymers polyamide and
polyester fibers, films, varnishes, coatings and other materials and articles became
widespread. This has given impetus to investigating the properties and formation of
notions, in particular, of anisotropic substances possessing extremely different
properties in different directions. A special place in the sequence of these polymers is
devoted to high-molecular elemento-organic compounds.
Finally, the representatives of the third class polymers with conjugated
system of bonds were used for the preparation of conducting materials.
Considering in general terms the chemical structure of polymers of different
classes, we have discussed the structural formula of the repeating unit in the
macromolecule. However, the existence of many such units in the macromolecule
immediately complicates the situation. Let us begin, for example, with an assumption
that each unit in the elementary act of macromolecule growth may be differently
attached to the neighbouring one; in this case, we are talking about the head-to-head,
tail-to-tail or head-to-tail addition. Various variants of the unit addition to the
propagating macromolecule are possible for asymmetric monomers of the
C H2 C H

R
type which possess R substituents on one of carbon atoms. Here, variants of head-tohead
... CH2 CH CH CH2 CH2 CH CH CH2 ...
R

and head-to-tail
...

CH2

CH CH2

CH CH2

CH

...

11

additions are possible.


Alternation of the types of addition is possible, i.e. units may be differently
attached to each other in a single macromolecule. Existence of a great number of units
in the polymeric chain and possibility of only several variants of their attachment
gives a huge number of isomers in relation to the whole macromolecule. To put it
differently, a polymer may contain (and indeed contains) not only the macromolecules
of the same chemical structure, but mixtures of a large number of macromolecules,
which, of course, makes the polymer to differ from low-molecular substances,
composed of identical molecules only.
We will not talk about a rapid increase of the number of possible isomers in
the sequence of substituted saturated hydrocarbons with the number of carbon atoms
(i.e. with propagation of the molecule); even at a small (compared with polymers)
number of them this number reaches a tremendous value. It is easy to imagine that
when the number of units becomes tens or hundreds of thousands, the number of
possible isomers becomes astronomically high [80].
Let us return to monosubstituted unsaturated hydrocarbons. When a polymeric
chain is formed during polymerization, the substituents R may dispose differently in
relation to the plane of single bonds. In one of possible cases, these substituents are
disposed irregularly in relation to the plane of single bonds; such polymers are called
irregular or atactic:
H
C
H

H
C
R

H
H
H R H
R H H
H
H
R
C
C
C
C
C
C
C
C
C
C
C
C
C
C
H
H
H
H
H
H
H
R
R
R
R
R
R
R

In other cases, synthesis may be performed in such a manner that substituents


would be disposed either by the same side of the plane of the main bonds
H
H
H
H
H
H H H
H
H H H
H
H
H
H
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
H
H
H
H
H
H
H
H
R
R
R
R
R
R
R
R
or by both sides, but with regular alternation of the substituents direction:
R
R
H
H
H
H H R
H
H R H
H
H
H
H
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
H
R
R
H
H
R
H
R
H
H
H
H
H
H
H
H
The polymers composed of the units with regular alternation of substituents
were called stereoregular. If the substituents are disposed on one side of the plane of
the main bonds, stereoregular polymers are called isotactic. If they are disposed on
both sides of the plane, the polymers are called syndiotactic.
The situation is more complicated with polymers synthesized from
disubstituted monomers. Already in the monomer, substituents may dispose on the
same (cis-isomer) or on both sides (trans-isomer) of the plane of the double bonds:
H H
H R'
C C

C C

R H

R'

12

Synthesis of macromolecules from cis-isomers leads to the formation of


erythro-diisotactic polymers
R
R
R R' R
R' R R'
R'
R'
R'
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
and trans-isomers give treo-diisotactic polymers
H
H
H
H
H
H R' H
R'
R' H R'
R'
R'
R'
R'
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
R
R
R
R
R
R
R
R
H
H
H
H
H
H
H
H
Needless to say, other more complex modifications are also possible, which
immediately cause a change of properties of polymeric materials.
The materials composed from stereoregular polymers are often easily
crystallized so that gives their physical structure and properties can be regulated.
Here we meet for the first time a modification of the properties of polymeric
materials, which is caused by practically any change in the chemical structure of
macromolecules and the physical structure of the polymeric substance. Physical
modification is often indicated by a change of the chemical structure, and sometimes
is completely defined by it.
One of the main methods of modification is the synthesis of copolymers, when
not a single but several monomers participate in the reaction. That is why the
macromolecule becomes composed from different units. These units may alternate
continuously:
ABABABABAB the alternating copolymer;
but, most often, they are arranged irregularly:
AABABBAAAB the random copolymer.
The units may also be linked in separate blocks which are the linked to each other:
AAAAABBBBB the block-copolymer.
Obviously, each block may contain a different number of units. This is
immediately shown up in the properties of the future polymeric substance. In this
case, the copolymerization process becomes regulated. Running ahead, recall that
mechanical mixtures of polymers and copolymers of the same molar composition may
often possess rather different properties, but sometimes they are practically identical.
The considered schemes of addition of units during macromolecule growth
indicate the only case of copolymerization of two types of monomers. Even if many
combinations are realized in these simplest cases, their number grows immeasurably
when three or more monomers (or types of units) are used
All the above-discussed chains of polymers represent linear formations.
However, branched macromolecular chains could be easily synthesized. For this
purpose, it is even unnecessary to introduce multifunctional compounds into the chain
composition. Branching also occurs in polymerization of unsaturated hydrocarbons
with no functional groups. If no special steps are taken, the products of
polymerization of ethylene, propylene, isobutylene and other similar compounds will
always contain some amount of chains branched from the main chain. Concerning the
products of polycondensation (see the above discussion on polyesters and
R

R'

R'

13

polyamides), introduction of a three-functional compound into the main chain always


leads to the formation of branched polymers:
...

A A A A' A A A A A A A

...

A
A
A
A
..
.
It is self-evident that the polymeric body based on the branched
macromolecules will differ in the structure and properties from a substance composed
of linear macromolecules. However, we must not hurry in concluding about the type
of physical structuring of the branched polymers. At first glance, it seems that the
presence of large branches will make obstacles to denser packing of the chains, as
well as to the crystallization process or regulation of macromolecules in general.
Indeed, this is sometimes the case. In other cases, the opposite situation is observed. It
depends upon the chemical structure of the main chain and its branches, which
determines the volume of units, interaction forces between them and neighbour
chains, etc.
Recently, special attention has been paid to the structure and properties of socalled dendric polymers, the macromolecule of which is schematically depicted in
Figure 1 [98, 212]. Below, we will discuss in more detail the influence of the types of
branchings on the properties of the resulting polymers.

Figure 1. Schematic representation of dendric polymers

Branchings may be composed in different ways. They may contain the same
units, which compose the main chain. However, grafted polymers have become
widely used; they are formed in grafting of previously obtained chains of a definite
structure to the main chain with an extremely different structure:
...

A A A A' A A A A A A A
B
B
B
B
..
.

...

14

Sometimes, such grafting is performed many times.


We can now easily pass from the branched to three-dimensional cross-linked
polymers. This requires just an increase of the concentration of multifunctional
compounds in the polymer chain. The chains could also be cross-linked by special
curing agents, i.e. by compounds containing active groups, capable of reaction with
functional groups of the main chain or the end groups. The classic example is the
curing of epoxy resins:
CH3
...

O CH2

CH CH2
O

CH3
CH3
...

O CH2

NH2
+

R
NH2

CH CH2
O

CH3
CH3
...

O CH2

CH3

CH CH2
OH

CH3
...

O CH2

CH CH2

NH
R
NH

CH3
OH
Further on, the second hydrogen atom is substituted, and a network is formed.
According to the classification described in ref. [202], there exist several main
methods of obtaining network polymers:
1) Realization of a chemical reaction between two (or more) different functional
end groups, attached to a chain of low molecular mass. As a result, a dense network
with short chains between cross-link points is formed.
2) Chemical linking of high-molecular compounds by the end groups with the
help of a low-molecular cross-linking agent. Consequently, a network with long linear
fragments between the cross-linked points is formed.
3) Formation of a network by copolymerization of two- and polyfunctional
monomers. The example of such a network is the styrenedivinylbenzene system:
...

CH2

CH CH2

CH CH2

...

...

CH2

CH CH2

CH CH2

...

4) Vulcanization of polymeric chains by involving, in the reaction, functional


groups disposed along the main chain. The reaction is performed either by the
application of a low-molecular cross-linking agent or by means of radiation and other
types of influence on the functional groups.

15

Other possible (and already realized in practice) ways of producing the


network systems should also be added.
5) Formation of networks with by means of a reaction of two (or more)
heterogeneous polymers by functional groups disposed along the chain of each
polymers (i.e. in the repeating units, but not at the ends).
6) Synthesis of polymeric networks with the help of the polycyclotrimerization
reaction. For this purpose, oligomers with end groups capable of forming cycles
during the reaction [56, 79, 101, 152] are formed. The example of such a reaction is
the trimerization of two-functional oligomers (or monomers) containing cyanate end
groups. Clearly, other ways of obtaining the polymeric networks are also possible.
Recently, a new type of polymer, called interpolymers was produced [16,
215]. The interpolymer is a system composed of two (or more) macromolecules,
heterogeneous in the chemical structure, chemically bonded to each other through the
functional groups disposed in the repeating units of the each macromolecule. A
schematic representation of the interpolymer is displayed in Figure 2.

Figure 2. Schematic representation of interpolymer.

A specific example of this system is, for example, a product of interaction


between polystyrene and polytrichlorobutadiene:
...

CH2

CH

... + ...

CH2

CCl2

...

CH CCl

CH2

...

AlCl3

...

CH CCl

CCl

CH2

CH

...

...

The formation of interpolymers gives new possibilities of modifying the structure and
properties of polymers.
Another type of two-cord system is the ladder polymer, the example of
which is polyphenylsylsesquioxane [113]:

...

Si

O
...

Si

Si

...

...

O
O

Si

Chapter II. Packing of macromolecules and polymer


density
II.1. Increments method and basic physical assumptions
After discussing briefly the chemical structure of polymers, let us pass to the
volumetric representation of macromolecules, which is necessary for understanding
the features of structure formation in polymers. These considerations will be based on
the assumptions developed by A.I. Kitaigorodsky in organic crystal chemistry [75].
According to these assumptions, every atom is presented as a sphere with
intermolecular radius R. Values of these radii are determined from the data of X-ray
structural analysis of ideal crystals of organic substances. In this case, it is assumed
that valency-unbonded atoms, entering into an intermolecular (but not chemical)
interaction, contact each other along the borders of the spheres. This is schematically
represented in Figure 3. Then, if two identical atoms are in contact, the intermolecular
radius will be determined from the relation:
R = l/2,

(II.1)

where l is the distance between mass centers of two identical valency-unbonded


atoms, which, however, are capable of intermolecular physical interaction.

Figure 3. Schematic representation of intermolecular (Van-der-Waals) interaction of two atoms

According to the same assumptions, chemical interaction between two atoms


always causes their compression, because the length of the chemical bond di is always
shorter than the sum of two intermolecular radii:
d i < R1 + R2 .

(II.2)

This is clear from Figure 4, which schematically depicts two chemically


bonded atoms. If the intermolecular radii Ri for all atoms participating in the repeat
unit, and all lengths of chemical bonds between these atoms are known, their own
(Van-der-Waals) volume of the repeat unit could be easily calculated, and a model of
this unit (or greater fragment of the macromolecule), in which the volume of each
atom is bordered by a sphere with intermolecular radius Ri, could be composed.

17

Figure 4. Schematic representation of two chemically bonded atoms.

Figure 5. Model of polyethylene chain fragment.

Table 1 shows intermolecular radii of some widespread atoms, which compose


the majority of polymers.
Table 1
Atom
C
H
O
N
F
Cl
Br
I

Van-der-Waals radii R of different atoms


R, nm
Atom
0.180
Si
0.117
Sn
0.136
As
0.157
S
0.150
P
0.178
Pb
0.195
B
0.221
Ti

R, nm
0.210
0.210
0.200
0.180
0.190
0.220
0.165
0.200

Table 2 displays bond lengths of various combinations of atoms, also


characteristic for most of existing polymers. If these values are known, the volume of
the repeat unit of any polymer may be calculated. To conduct this, the own volume of
each atom participating in the repeat unit should be preliminarily determined. It is
calculated from the formula
3
1
Vi = R 3 hi2 (3R hi ),
4
i 3

(II.3)

where Vi is the increment of the own (Van-der-Waals) volume of the present atom;
R is the intermolecular radius of this atom; hi is the height of the sphere segment, cut
off from the present atom by a neighbor one, chemically bonded to it. The value hi is
calculated from relation

18

hi = R

R 2 + d i2 Ri2
,
2d i

(II.4)

where Ri is the intermolecular radius of a neighbor valency-bonded atom; di is the


length of the chemical bond (see Figure 4).
Table 2
Chemical bond length di for same pairs of atoms
Bond*
di, nm
Bond*
di, nm
Bond*
di, nm
CC
0.154
CF
0.134
OF
0.161
CC
0.148
CF
0.131
O=N
0.120
C=C
0.140
CCl
0.177
O=S
0.144
C=C
0.134
CCl
0.164
O=P
0.145
C=C
0.119
CBr
0.194
NP
0.165
CH
0.108
CBr
0.185
NP
0.163
CO
0.150
CI
0.221
NP
0.158
CO
0.137
CI
0.205
SS
0.210
CN
0.140
CB
0.173
SAs
0.221
CN
0.137
CSn
0.215
S=As
0.208
C=N
0.131
CAs
0.196
SiSi
0.232
C=N
0.127
CPb
0.220
PF
0.155
0.134
HO
0.108
PCl
0.201
C
N
CN
0.116
HS
0.133
PS
0.181
CS
0.176
HN
0.108
BB
0.177
CS
0.156
HB
0.108
SnCl
0.235
CSi
0.188
OS
0.176
AsCl
0.216
CSi
0.168
OSi
0.164
AsAs
0.242
* If the same pair of atoms is linked by a single bond, the longer bond corresponds to attachment of this
atom to an aliphatic carbon atom; the shorter bond corresponds to attachment of the same atom to an
aromatic carbon atom.

Increments of the volumes of various atoms and atomic groups are shown in
Table 3. Obviously, the volume of the given atom depends on its surrounding, i.e. on
the type of atoms chemically bonded to it. The greater the volume of the neighbor,
chemically bonded atom and the shorter the length of the chemical bond, the greater is
the compression of the given atom.
When increments of the volumes, Vi, of all the atoms entering into the repeat
unit of polymers are determined, the relative part of the occupied volume in the total
volume of the polymeric substance may be calculated. In the case of polymer,
calculations would be appropriate to conduct basing on molar volumes of the repeat
unit, because polymers are always polydispersional (i.e. they contain macromolecules
of various length), and also because at long lengths of the macromolecule the
influence of end groups may be neglected. Then, the own molar volume will equal
Vown = N A Vi , and the total molar volume Vtotal = M/, is density of the
i

polymeric substance; M is the molecular mass of the repeat unit; NA is the Avogadro
number. Numerous experiments and calculations show that in all cases the condition
Vown < Vtotal is fulfilled. Hence, in the first approximation, the volume of the polymeric
substance could be divided into two parts: the own (Van-der-Waals) volume of atoms,
which they occupy in a solid, and the volume of spaces determined as the difference
of Vtotal and Vown. Of interest is determination of the part

19

Table 3
Van-der-Waals volumes of atoms

20

21

22

23

24

25

26

27

28

29

of the occupied volume or, according to the terminology used in organic crystal
chemistry, the molecular packing coefficient k:

V
k = own =
Vtotal

N A Vi
i

M /

(II.5)

Clearly, the value of k for the same polymer will depend on temperature and
the physical state of the polymer, because the value of depends on them.
Calculations performed for many amorphous bulky polymers existing in the glassy
state have indicated that the first approximation of k gives its value constant and
practically independent of the chemical structure of the polymer [41]. Passing on to
polymers with a complicated chemical structure from those with a simple one causes
no significant change of the part of the occupied volume (e.g. the value of k).
Table 4 indicates the chemical structure and numerical values of coefficients
of the molecular packing of some glassy polymers. It also shows that first
approximations of the values of k for each of them are equal, indeed. To demonstrate
this experimental fact more clearly, Figure 6 displays the dependence of density of
various polymers on the relation M N A Vi . In Figure 6 it is clearly seen that all
i

30

Table 4
Values of the coefficients of molecular packing for some glassy and semi-crystalline polymers
Structural formula of the repeat unit of polymer
Van-der-Waals
Packing
volume of the
coefficient k
3
unit, cm /mol

CH3
CH2

41.6

0.678

32.6

0.682

58.5

0.684

69.1

0.680

144.3

0.679

234.7

0.679

263.1

0.680

277.5

0.688

56.4

0.685

CH3
H
CH2

C
C N
CH3

CH2

C
C O CH3
O
CH3

CH2

C
C O C2H5
O

CH3
C O

CH3

C O

O
C

O
C
O

C
O

C
O

C O
O

(CH2)8

O
C

HN

NH
C

NH
C
O

CH2

CH
CH
CH2

31

CH2CH=CHCH2
CH2

CH C

CH2

59.1
74.3

0.654
0.659

100

0.699

97.8

0.708

110.3

0.693

269.0

0.692

43.9
72.4

0.753
0.663

33.8
54.9

0.700
0.666

58.7
36.0
123.1

0.654
0.744
0.641

134.3

0.664

CH3
(CH2)5

NH C
O

CH2

CH
CH

H2C
HC

CH2
CH2

N
CH3
CH2

CF2CF2
CH2

CH
O C

CH3

O
CH2CHF
CH2

CH
O
CH3

CH2CCl2
CH2CF2
CH3
CH2

CH3

C O CH
O

CH3

CH3
CH2

C
C O C4H9
O

32

CH3
CH2

168.3

0.651

120.0

0.607

85.9

0.696

163.0

0.687

88.8

0.705

C
C O C6H13
O

CH3
Si

CH2 CH2 CF3


CH2 CH2 CH2 CH3
C N
O
CH2

CH
N

CH2
H2C

CH
N

C O
CH2

H2C
CH2

CH

111.6

0.669

CH2

Cl
CH

115.5

0.657

65.6

0.638

89.3

0.650

40.0

0.681

CH3
CH2

CH

CH3
CH2

CH
C O C2H5

O
O CH2

C
O

33

O CH CH2
CH3

69.9

0.684

172.5

0.740

70.6

0.677

21.3
126.1

0.752
0.616

118.5

0.667

53.0

0.733

150.8

0.679

103.0

0.620

76.2

0.568

46.4
144.4

0.680
0.692

227.7

0.693

C NH

O
CH2

NH

CH
C O CH3

O
CH2O
CH3
Si

CH2

CH
C O C4H9
O
F
C

CF2

Cl
CH3
CH2

CH2
C O CH
CH2
O

CH2
CH2
CH2

C2H5
Si

C2H5
CH3
Si

CH3
CH2CH2S
(CH2)2

O C

C O

O
O

O
C
O

34

154.1

0.696

157.0

0.721

30.2
46.3

0.682
0.666

99.6

0.665

262.1

0.726

C
O

SO 2

CH2CH2
CH2

CH
CH3

CH2

CH2

CH

NH C O

(CH2)4

O C NH

Figure 6. Dependence of density on

  '

the values of determined experimentally fit well the same linear dependence on the
relation of atoms mass on their volume. In accordance with Equation (II.5), the
tangent of this straight line represents the molecular packing coefficient which, in the
case of amorphous bulky systems, serves as an universal constant. If it is true, the
polymer density may be calculated from the equation

kM
,
N A Vi
i

(II.6)

35

that yields directly from Equation (II.5) under the condition kavg = const. In the case of
amorphous bulky polymers, kavg = 0.681. For silicon-containing polymers, the average
coefficient of molecular packing is 0.603.
Hence, a change of the polymer chemical structure is unable to cause a
significant effect on the part of the occupied volume in amorphous polymeric
substance, and the value of density, , itself depends on the relation of mass and the
Van-der-Walls volume of the repeat unit only.
Obviously, here we are dealing with true bulky substances of the amorphous
structure. In reality, a polymeric substance with any porosity may be formed, and the
coefficient k will have extremely different values. However, in this case, the notion of
the packing density, quantitatively estimated by the value of k, loses its usual meaning
and must be calculated for pore walls material only. We return to this problem below
when discuss parameters of the porous structure of polymers, determined by the
sorption method.
For copolymers, equation (II.6) has the form

kavg (1M 1 + 2 M 2 + ... + n M n )



N A 1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n

(II.7)

where 1, 2, , n are molar parts of the components 1, 2, , n; M1, M2, , Mn are

molecular masses of the repeat units of the same components; Vi , Vi ,

i
1 i
2

, Vi are their Van-der-Waals volumes.

i
n
In the reduced form, expression (II.7) is:
k =n

k avg

k =n

k M k

k =1

N A k Vi

k =1
i
k

(II.8)

where k, Mk, Vi are the molar part, the molecular mass, and the Van-der

i
k
Waals volume of the k-th component, respectively.
If we want to express the density of copolymer via densities 1, 2, , n of
homopolymers based on the components 1, 2, , n, expression (II.7) changes to the
following form:

1M 1 + 2 M 2 + ... + n M n
,
M
M
M
1 1 + 2 2 + ... + n n
1
2
n

(II.9)

36

(in this case, it should be taken into account that 1 + 2 + + n = 1).


In the reduced form, the expression (II.9) is the following:
k =n

k M k

= k =1
k =n

(II.10)

M
k k
k
k =1

Expressions (II.7)(II.10) may also be used for calculating the density of


miscible blends of polymers.
Let us now examine the temperature dependences of the molecular packing
coefficients of glassy polymers. Calculation of values of k at different temperatures
are performed by formulae yielding from the expression (II.5):

k (T ) =

k (T ) =

N A Vi

)] ,

(T < Tg);

(II.11)

)] ,

(T > Tg);

(II.12)

MVg 1 + G T Tg
N A Vi

MVg 1 + L T Tg

where Vg is the specific volume of the polymer at the glass transition temperature Tg;
G and L are the volume expansion coefficients of polymers below and above the
glass transition temperature, respectively.

Figure 7. Temperature dependences of the coefficients of molecular packing k for a series of polymers:
1 poly(n-butyl methacrylate), 2 poly(n-propyl methacrylate), 3 poly(ethyl methacrylate), 4
polystyrene, 5 poly(methyl methacrylate), 6 polycarbonate based on bisphenol A.

Calculations by equations (II.11) and (II.12) indicate that temperature


dependences of the molecular packing coefficients are of the form depicted in Figure
7. A remarkable property of these temperature dependences in the real equality of the
molecular packing coefficient in the first approximation for all bulky polymers at any
temperature below the glass transition point. In the second, more accurate
approximation, the molecular packing coefficient is the same for every polymer at the
glass transition temperature. This value is kg 0.667.

Table 5
Name
1
Polyethylene

Coefficients of molecular packing k for a series of crystalline polymers


Type of elementary cell
Chemical formula
2
3
Rhombic
Pseudo-monoclinic
CH2CH2
Triclinic

Polypropylene:
- isotactic
- syndiotactic

Monoclinic
Monoclinic

1,2-poly(butadiene):
- isotactic
- syndiotactic

Rhombic
Rhombohedral

1,4-trans-poly(butadiene)

Pseudo-hexagonal

1,4-cis-poly(butadiene)

Monoclinic

1,4-cis-polyisoprene

Monoclinic

Polychloroprene

Rhombic

Poly(ethylene terephthalate)

Triclinic

CH2

CH3
CH2

CH
CH

CH2
CH2CH=CHCH2
CH2CH=CHCH2
CH2 CH C CH2
CH3
CH2

O CH2
Poly(hexamethylene
terephthalate)

CH

CH2

CH C
Cl

CH2

O C

Triclinic

(CH2)6

O C

, g/cm3
4
1.000
1.014
0.965
1.013

k
5
0.736
0.746
0.710
0.745

0.936
0.910

0.693
0.674

0.963
0.960

0.692
0.690

1.020

0.733

1.010

0.726

1.000

0.725

1.657

0.893

1.455

0.776

1.131

0.652

37

38

Triclinic

O CH2

CH2

4
1.358

5
0.724

1.274

0.782

1.240
1.248

0.764
0.769

1.157

0.740

1.230

0.758

1.192

0.789

0.813

0.598

O C
O
C
O

Poly(ethylene adipate)

Triclinic

(CH2)2

O C

(CH2)4

O
Polyamide 6,6:
-isomer
-isomer

Triclinic
Triclinic

Polyamide 6,10

Triclinic

Polyamide 6

Monoclinic

(CH2)4

C
O

HN

(CH2)6

NH

HN

(CH2)6

NH

O
(CH2)8

O
C HN

(CH2)5

O
Polyamide 11

Triclinic

C HN

(CH2)10

O
Poly-4-methylpentene-1

Tetragonal

CH2

CH
CH2
CH CH3
CH3

38

1
Poly(ethylene isophthalate)

39

1
Polyvinylchloride

2
Rhombic
Monoclinic

CH2

CH

4
1.440
1.455

5
0.680
0.687

2.400
2.360

0.794
0.781

1.440

0.742

1.350

0.770

1.110

0.677

1.230

0.719

1.120

0.711

1.506
1.205

0.808
0.723

Cl
Polytetrafluoroethylene

Pseudo-hexagonal
Hexagonal

Polyvinylfluoride

Hexagonal

Poly(vinyl alcohol)

Monoclinic

CF2CF2
CH2

CH
F

CH2

CH
OH

Polyacrylonitrile

Rhombic

Poly(methyl methacrylate)
isotactic

Pseudo-rhombic

CH2

CH
C N

CH3
CH2

C
C

O CH3

O
Rhombohedral

Polyoxymethylene
Polyethylene oxide

Hexagonal
Hexagonal

CH2 CH

CH2O
CH2CH2O

39

Polystyrene

40

Rhombic

CH2 CH O
CH3

4
1.102
1.154

5
0.663
0.694

40

1
Polypropylene oxide

41

Taking into account that the specific volume at the glass transition temperature
Tg equals

Vg = 1

N A Vi
i

kg M

(II.13)

where g is the polymer density at Tg; and substituting (13) into (11) and (12), we get
k (T ) =
k (T ) =

kg

[1 + G (T Tg )] ,
kg

[1 + L (T Tg )] ,

(T < Tg);

(II.14)

(T > Tg);

(II.15)

Equations (II.14) and (II.15) can be used for obtaining relations, which
describe temperature dependences of the density of polymers in the glassy and
rubbery states. For this purpose, we substitute (II.14) and (II.15) into equation (II.6):

(T ) =

kg M

[1 + G (T Tg )]N A Vi ,

(T < Tg);

(II.16)

(T > Tg);

(II.17)

k (T ) =

kg M

[1 + L (T Tg )]N A Vi ,
i

Because, as it is seen from the further considerations, values of expansion


coefficients G and L, as well as the glass transition temperature Tg, can be
calculated from the chemical structure of the repeating polymer unit, temperature
dependences of density (T) can also be calculated from relations (II.16) and (II.17).
In conclusion, let us note that the constancy of the coefficient of molecular
packing k is true only for amorphous bulky substances composed of polymers. In the
case of crystalline polymeric substances, the situation is significantly changed. If the
coefficients of molecular packing for ideal polymeric crystals are calculated with the
help of the X-ray analysis data, one can assure himself that, in spite of amorphous
ones, the coefficients of molecular packing of crystalline polymers are extremely
different. The smallest values of k are typical of aliphatic systems with volumetric
side groups, for example, for poly-4-methylpentene-1 and poly-n-butyraldehyde. The
highest coefficients of packing are typical of 1,4-trans--polyisoprene and polychloroprene.
As an example, Table 5 shows the crystallographic values of densities and
molecular packing coefficients for a series of typical crystalline polymers. It is clear
that the values of k for them vary in a wide range. Hence, crystalline polymers display
a rather wide distribution curve of the coefficients of molecular packing (Figure 8).

42

Figure 8. Curve of distribution of the coefficients of molecular packing k for crystalline polymers.

II.2. Relationship between free volume of polymers, coefficient of


molecular packing and porous structure
Before we start discussing the relationship between the above-mentioned
physical characteristics, the term of the free volume must be discussed in brief.
There are three definitions of the free volume:
1)
The free volume represents the difference between the true molar
volume of the substance, VM, and its Van-der-Waals molar volume N A Vi :
V = VM N A Vi = M / N A Vi .
i

(II.18)

The value of V obtained in this way is often called the empty volume.
Clearly, the empty volume depends on temperature, because the molar volume also
depends on it: VM = M/. Substituting this relation into equations (II.16) and (II.17),
we obtain:
1 + G T Tg

V (T ) = N A Vi
1 , (T < Tg);
(II.19)
kg
i

1 + L T Tg

V (T ) = N A Vi
1 , (T > Tg);
(II.20)
kg
i

Relations (II.19) and (II.20) describe the temperature dependences of the


empty volume.
2)
The free volume represents the difference between the volumes of the
substance at the absolute zero and at the assigned temperature; to put it differently, the
free volume represents an excessive volume occurring as a result of thermal
expansion of the substance. This definition of the free volume is most valuable.
Moreover, the present free volume is subdivided into the free volume of fluctuation
and the expansion volume.
3)
The free volume represents the difference between the volume of
polymeric substance at the assigned temperature and the volume of the ideal crystal

43

composed of a polymer of the same chemical structure. This definition of the free
volume is used extremely seldom.
Let us now pass to analysis of the relationship between the free volume of
polymers, the coefficient of molecular packing and the porous structure.
The porous structure mostly defines their properties. That is why the methods
of estimation of the porous structure of polymers and its connection with such
characteristics as the coefficient of molecular packing and the free volume of polymer
must be discussed in detail. The case is that the size of micropores depends on the
method of its estimation. Clearly, interpretation of their nature and the relationship of
the characteristics of the microporous structure with the properties of polymers
significantly depends on the method of their determination.
The properties of many bulky and film polymers significantly depend on the
density of packing of macromolecules, and for such systems as sorbents, ionites, etc.,
used in gel-chromatography and production of ion exchangers, the volume of pores is
very important, together with their size distribution, specific surface.
Let us present the definition, given in ref. [68]: Pores are emptinesses or
cavities in solids usually connected with each other. They possess various and
different form and size, determined significantly by nature and the way of obtaining
absorbents.
Usually, the characteristics of a microporous structure are judged by
experimental data on equilibrium adsorption, capillary condensation of vapor and
mercury pressing in (mercury porosimetry) [121]. Recently, the positron annihilation
method has been used [3, 48, 110, 123, 134, 140, 155, 164, 187, 211]. This method
helps in determining the characteristics of the microporous structure, when the size of
pores is commensurable with the molecule size. Such micropores are inaccessible for
sorbate molecules and especially for mercury when mercury porosimetry is used.
Polymers and materials prepared from them possess the feature (in contrast to
mineral sorbents) that they swell during sorption of vapors of organic liquids.
Consequently, their structure changes and usual methods of calculation give no
possibility of estimating the true porous structure of the initial material. It stands to
reason that vapors of organic liquids, in which polymer does not swell, can be used in
sorption experiments. Then the parameters of the porous structure of the initial
material can be determined, but these cases are quite rare [107].
Before passing to comparison of parameters of the porous structure with the
free volume of the polymer, it should be noted that parameters of the porous structure
for the same polymer could be significantly different due to conditions of its synthesis
and further processing. For example, a film or fibers may be obtained from various
solvents [81], as well as from a solventprecipitant mixture [97], and will display a
different microporous structure and properties. The same can be said about materials
obtained by pressing and injection molding and with the help of hydrostatic extrusion
as well. Therewith, macropores may also be formed and their total volume may be
quite high. If special synthesis methods are used, materials based on polymer
networks may be obtained, which possess a large specific surface and extremely large
pore radii [115]. Clearly, such macropores are not defined by the packing density of
macromolecules. They may be formed by loose packing of formations larger than
macromolecules or may be caused by conduction of a chemical process of the
network formation under special conditions [167].
Several more general comments should be made. Besides macropores, as
mentioned above, micropores are present in a polymeric substance, the size of which
is commensurable with the size of sorbate molecules. Clearly, in this case, sorbate

44

molecules cannot penetrate into these micropores (it is assumed that for sorbate
molecules to penetrate into pores, the volume of the latter must be several times
greater than that of penetrating molecules). Since sorbate molecules may be different,
i.e. may possess different sizes, parameters of the porous structure determined from
the sorption data will depend on types and sizes of molecules of sorbed substances.
That is why such terms as porosity to nitrogen, porosity to benzene, etc. have been
introduced. Of interest is that the sorption method of determination of the porous
structure of polymeric substances cannot be used in the case when a substance
contains quite large macropores. This is associated with the fact that under conditions
of polymolecular adsorption, when many molecular layers are formed on walls of
macropores, their fusion becomes difficult, i.e. capillary condensation is absent. Then,
the total volume of pores calculated by the amount of sorbate penetrated into the
polymeric substance will be smaller than the true volume of macropores.
Starting the analysis of relationship between the physical characteristics of the
polymeric substance and its microporous structure, let us introduce some definitions
and designations:
Ssp is the specific surface of micropores,
W0 is the total volume of pores,
W0max is the maximal volume of pores accessible for sorbate molecules of any
size (per gram of the substance),
VF is the free volume (in the present case, the volume of expansion),
VE is the empty volume (see above),
VT is the specific volume of the polymeric substance at given temperature,
VW is the Van-der-Waals volume (per gram of the substance),
Vid.cr. is the specific volume of the ideal crystal or bulky amorphous polymer (a
bulky amorphous polymer is the one in which no sorbate molecule can penetrate into
its pores).
Let us write down some relations connecting these characteristics:
VF = VT V0 ;
(II.21)
VE = VT VW.
(II.22)
Next, let connect these characteristics with the coefficient of molecular
packing k (see above):
k = VW/VT; 1 k = VE/VT.
(II.23)
As mentioned above, there are so-called non-porous sorbents (for example,
crystalline substances), into which no molecules of sorbate can penetrate without
swelling. Clearly, that for such substances W0max = 0. At the same time, as seen from
the data in Table 5, coefficients of molecular packing of crystals fall within the range
from 0.64 to 0.89. Taking into account that the coefficient of molecular packing, by
definition, represents a part of the occupied (Van-der-Waals) volume, it can be said
that the part of empty (but inaccessible) volume is 1 k = 0.110.36. This empty
volume is inaccessible for even small sorbate molecules to penetrate in; let mark it as
Vinacc.. Then the volume of the ideal crystal (or bulky amorphous polymer, Vblk) can be
written down as
Vid.cr. = VW + Vinacc.; Vblk = VW + Vinacc..
(II.24)
The volume of the real polymeric substance (which contains micropores
accessible for a sorbate) will be summed up from three parts:
VT = VW + Vinacc. + W0max.
(II.25)
Then
W0max = VT Vid.cr.; W0max = VT Vblk.
(II.26)

45

The coefficient of molecular packing in the bulky part of the polymer will be
determined from the relation
VW
k=
.
(II.27)
VT W0max
In the case of estimation of the density of macromolecule packing for the real
polymeric substance containing micropores accessible for sorbate molecules, the
coefficient of molecular packing, k, should be calculated by the relation
VW
k=
,
(II.28)
VT W0
where W0 is the total volume of micropores (per gram of the substance), determined
on the basis of sorption measurements.
The value of W0max that represents the difference between the specific volume
of the substance at the given temperature and volume of the true bulky substance is
conceptually identical to the porosity factor P = 1/s 1/t, where s is the apparent
density; t is the true density. Therewith, s represents the density of the substance at
the current temperature, affected by the pores existing in it. It is best to measure the
apparent density of substances with the proper geometrical shape, because when using
no solvents s can be found by dividing the substance weight by its volume. If the
apparent density of substances with the improper shape is measured, the pycnometric
or dilatometric method can be used. The difficulty is in selection of a liquid that does
not wet the surface of the substance and does not penetrate deep into it. The true
density t represents density of the bulky part of the substance containing no pores. It
is best to measure the density of the ideal crystal, because it can be calculated on the
basis of crystalline lattice parameters. In the case of amorphous and partly crystalline
substances, the method of gradient tubes may be used applying liquids penetrating
well into pores. However, it should be taken into account that a mixture of two liquids
is used for creation of the density gradient in the tube, each of which may possess
different wettability and penetrability into pores. The picture is then distorted, and the
determined density is not true.
The relations shown above can be estimated unambiguously if a polymer
swells in the sorbate, used for estimation of the porous structure of the polymer. If the
experiment indicates that W0 is greater than W0max, this indicates that the volume of
vapors absorbed by the polymer is greater than the volume of pores existing in it, i.e.
the polymer swells during sorption.
Let us now consider the experimental and calculated data on determination of
the parameters of the polymer structure and coefficients of their molecular packing.
These data are shown in Table 6. For ideal polyethylene crystallites, VE = Vinacc. and
W0max = 0. The coefficient of molecular packing is quite high. For semi-crystalline
polyethylene, the empty volume, VE, is greater than in the case of the ideal crystal
and, therewith, a part of it is accessible for penetration of small sorbate molecules.
However, the total volume of pores determined by methanol sorption equals 0.01
cm3/g. The molecular packing coefficient for the bulky part of such polyethylene is
significantly lower than for the ideal crystal.
Polymers in the rubbery state (polyisobutylene, for example) also possess
comparatively low values of free volumes and are practically non-porous sorbents
(VE = Vinacc.).
Contrary to this, polymers produced by polycondensation or polymerization in
solution display immensely high values of W0max. In this synthesis method, pores are
formed due to elimination of the solvent, distributed in the volume of the synthesized

46

polymer. This is observed from the fact that the same polymers produced by
polymerization in the melt are practically non-porous, and values of VE for them are
very small, and W0max = 0.
Table 6
Parameters of porous structure and coefficients of molecular packing of a series of polymers
VE,
W0max,
W0,
Vinacc.,
K
Polymer
cm3/g
cm3/g cm3/g
cm3/g
Polyethylene (100% crystallinity)
0.26
~0
~0
0.26
0.736

CH2CH2
Polyethylene (crystallinity < 100%)

0.35

0.08

0.01

0.27

0.675

0.36

~0

~0

0.36

0.678

1.28
0.22

1.06
~0

0.22
0.22

0.687
0.687

0.82
0.24

0.58
~0

0.31
~0

0.24
0.24

0.688
0.688

CH2CH2
Polyisobutylene

CH2C(CH3)2
Polymethylidenphthalide

CH2 C
O
C
O
Polymerization in dimethylformamide solution
Polymerization in melt
Polyarylate F-1
C

C O

O
C
O
C
O

Polycondensation in chlorinated bisphenol solution


pressed at 360C and under 312.5 MPa pressure

Pores formed during synthesis may be closed in polymer pressing under high
pressure, and the porous polymer then becomes non-porous. Therewith, in all cases,
W0 is smaller than W0max that indicates the absence of swelling.
For all polymers, values of Vinacc. are close to these characteristics for the
density of crystallized samples. Of special attention is the fact that independently of
the production method, the molecular packing coefficient for amorphous and semicrystalline polymers in their bulky part is the same and close to the average value
kavg = 0.681, which was discussed above. For a crystalline sample, the value of k is
significantly higher.
There is one more interesting point to discuss, associated with molecular
packing, namely, the change of the system volume during polymerization, i.e. at
transition from monomer to polymer.
It is well known that transition from a monomeric liquid to a solid glassy
polymer is accompanied by a significant contraction, i.e. volume decrease [76]. The
specific volume of the polymer Vp is always smaller than that of monomer Vm, and
their difference V = Vp Vm < 0. One of the reasons for contraction is substitution of
longer intermolecular bonds existing in liquid monomers by shorter chemical bonds

47

formed between monomer molecules in the polymer. Therewith, the own Van-derWaals volumes of atoms decrease owing to their compressing (see above).
Nevertheless, this is not the only reason of contraction. It follows from consideration
of the experimentally determined specific volumes that there is another reason for
contraction, which is more dense packing of polymeric chains compared with the
packing of monomeric molecules. This is indicated by the fact that the packing
coefficients of polymers are always greater than those of their monomers (kp > km).
Let the total contraction, Vtotal, be presented as a sum of two values: V1,
which is the contraction stipulated by substitution of intermolecular bonds by
chemical ones, and V2, which is the contraction involved by more dense packing of
chains,
Vtotal = V1 + V2,

(II.29)

and each of the summands estimated.


To do this, values of the specific volume of a polymer should be calculated on
the assumption that it displays the packing coefficient, the same as the monomer km,
i.e.

Vi

i
p
N
Vp = A
,
(II.30)
M
km

where Vi is the Van-der-Waals volume of atoms in the repeat unit of the

i
p
polymer; M is the molecular mass of the unit. Values of Vp for some polymers,
calculated in this way, are shown in Table 7. They are always greater than
experimentally measured values of specific volumes of the polymer, Vp.
The difference between Vp and Vm is
V1 = Vp Vm,

(II.31)

and the remaining part of the contraction is calculated by the formula


V2 = Vtotal V1.

(II.32)

Relative parts of contraction are determined from the relations:

1 = V1/Vtotal;
2 = V2/Vtotal.

(II.33)
(II.34)

The data shown in Table 7 indicate that in all the cases the smaller part of
contraction depends upon opening of double bonds, and the greater part on dense
packing of polymer chains. Therewith, the chemical structure of a monomer and an
appropriate polymer significantly affects the values of 1 and 2.

49

Polymer (monomer)
1

CH3
CH2

V1, cm3/g
6
0.080

V2, cm3/g
7
0.133

1, %
8
37.6

2, %
9
62.4

C
C O CH3
O
CH3

CH2

1.102

0.890

1.031

0.212

0.071

0.141

33.5

66.5

1.109

0.928

1.045

0.181

0.064

0.117

35.4

64.6

1.046

0.815

0.951

0.231

0.095

0.136

41.1

58.9

1.082

0.873

1.000

0.209

0.082

0.127

39.2

60.8

C
C O C2H5
O
CH3

CH2

C
C O C3H7
O

CH2

CH
C O CH3
O

CH2

CH
C O C2H5
O

48

Table 7
Changes in volume of the system as a result of polymerization
Vm, cm3/g
Vn, cm3/g
Vtotal, cm3/g
Vn, cm3/g
2
3
4
5
1.068
0.855
0.968
0.213

50

CH2

CH

2
1.098

3
0.952

4
1.036

5
0.146

6
0.062

7
0.084

8
42.5

9
57.5

1.073

0.841

0.976

0.232

0.097

0.135

41.8

58.2

1.104

0.942

1.028

0.162

0.076

0.086

46.9

53.1

C O C4H9
CH2

O
CH
O

CH2

C CH3
O
CH

49

50

In the set of polyacrylates and polymethacrylates 2 grows first with the


volume of the side substituent and then decreases. Decrease of the intensity of the
effect of the dense packing of chains, apparently, depends upon steric hindrances.
Hence, it follows from the above-said that the notions of porosity and packing
density are inadequate. Porosity reflects almost always cavities greater than the
molecular size, i.e. quite large ones. As for the packing density of macromolecules
themselves, it may be judged by considering the non-porous part of the sample only.
As noted above, application of positron annihilation methods is preferable for
analyzing the microporous structure of polymers [3, 48, 110, 123, 134, 140, 155, 164,
187, 211]. With the help of these methods, qualitative and quantitative information
about the characteristics of submicropores (215 ) in polymers may be obtained.
Let us discuss the results of studying annihilation of positrons in two
polymers, which are good models of the limiting characteristics of the packing density
of macromolecular chains. One of them is polyimide characterized by a highly
regular, quasi-crystalline structure, and the second is poly(1-trimethylsilyl-1-propyne)
(PTMSP) which, on the contrary, is characterized by a low coefficient of molecular
packing.
Consider structural changes in PTMSP, which appear during its long exposure
at room temperature after synthesis.
For comparison, we also display the data on annihilation of positrons for a
series of other model polymers. The chemical structures of all above-mentioned
systems are shown below.
Poly(1-trimethylsilyl-1-propyne)
CH3
C

Si

H3C

CH3

CH3
n

Polyisoprene
CH2

C CH CH2
CH3

Polydimethylsiloxane
CH3
O

Si
CH3

Polystyrene
CH2

CH

n
Polytetraflouroethylene
[CF2CF2]n

51

Polyimide
O

O
n

Observation of the annihilation of positrons in PTMSP was performed with


the help of a method of detection of the lifetime spectra of positrons (measurements
were made by S.A. Tishin; data not published). Measurements were performed by a
thermostabilized spectrometer, which realizes the traditional fastslow scheme of
detection, with a temporal photomultiplier selected and optimized due to an original
method [111].
Processing of experimental spectra was performed with the help of wellknown software Resolution and Positron FIT.
Table 8 shows the results of separation of parameters of a long-living
component at three-component decomposition of positron lifetime spectra for
PTMSP, polyimide, polystyrene, polydimethylsiloxane and polytetrafluoroethylene.
Clearly, PTMSP possesses an anomalous long lifetime for an ortho-positronium atom,
to annihilation of which by a pick-offdecay the origin of a long-living component of
the lifetime spectrum in polymers is bound [3, 48, 110, 123, 134, 140, 155, 164, 187,
211]. Hitherto, the maximal lifetime of the long-living component, D, was observed
in polydimethylsiloxane and teflon in solid polymers [123, 164]. Comparison with the
results of measurements in model polymers (see Table 8) indicates that neither the
presence of an unsaturated bond, nor the presence of a side group or silicon atom
separately is the explanation of so high D for PTMSP.
Table 8
Parameters of the longest component of positron lifetime spectrum for a series of polymers and
rated values of radius R and volume V of micropores
R,
V, 3 E, eV
Sample
D + 0.03, ns ID 0.25, % R0,
PTMSP
5.78
38.4
6.76
5.10
416.5
0.41
Polytetrafluoroethylene
4.27
21.6
6.05
4.39
265.8
0.51
Polydimethylsiloxane
3.23
41.3
5.45
3.79
170.9
0.63
Polyimide
2.77
38.1
5.14
3.48
132.1
0.71
Polystyrene (atactic)
2.05
40.5
4.56
2.90
76.9
0.90

Two suggestions about the reasons of anomalous long average lifetime of


positrons in PTMSP can be made.
First, molecular structure of the repeat unit allows a supposition that a high
concentration of bulky, low-mobile side groups creates a porous structure with the
pore size of about Van-der-Waals volume of SiC3H9 side fragment.
Secondly, the size of pores may be associated with a long relaxation time of
synthesized PTMSP at room temperature. It may be suggested that the formation and
evolution of microcavities of a large size must depend on the motion of large
segments of macromolecules or even structural fragments with a long period of
regrouping.
The lifetime of an ortho-positronium atom regarding the pick-offannihilation
allows estimation of the size of the microcavity in which it was localized before
annihilation [140]. The calculation results are also shown in Table 8.

52

In line with the model [140], positronium is considered in a spherical pit


surrounded by a layer of electrons, R thick. For wave functions in spherical
coordinates:
(2R0 )1 / 2 1 sin (r / R0 )
in the pit;
r
%( r ) =
(II.35)
0
outside the pit.
The probability of positronium existence outside the limits of density will be:
2 R
R
1
,
W ( R) = 1
+ sin
(II.36)
R0 2 R0

where R = R0 R.
Suggesting that the rate of ortho-positronium annihilation inside the electron
layer equals 0.5 ns1, the decomposition rate averaged over spins will be:

D = 1/D = 2W(R)

(II.37)

with the constant R = 1.66 , selected empirically for solids.


Let us consider the results of measurements of PTMSP films porous structure
because of their aging.
Long-term relaxation of PTMSP films was investigated with the help of
measuring positron lifetime spectra. As Table 9 and Figure 9 indicate displaying a
series of characteristics of time spectrum decomposition into three components and
the calculated radius of micropores R, and durability of samples aging, lifetime of the
long-living component decreases with growth of PTMSP exposure time at room
temperature. In practice, the intensity of the long-living component does not depend
on the relaxation time.
Table 9
Long-term relaxation of PTMSP from the data of measurement of the longest component
parameters of positron lifetime spectrum (n is lifetime of intermediate component)
Aging time, days
RD 0.25, %
D 0.03, ns
n 0.080, ns
13
5.78
38.40
0.687
17
5.68
37.53
0.607
24
5.72
38.09
0.678
83
5.40
38.08
0.507
210
5.09
37.91
0.453

Figure 9. Dependence of sizes R of the positron-sensitive microcavity on time of exposure tc at 25C


for PTMSP

53

The result observed is connected with slow structural relaxation but not the
aging (if by the aging occurrence of the main chain fission is meant), because the
latter process is usually accompanied by changes in intensity ID (results of observing
long-term aging of polyethylene by the method of positron lifetime variation may be
displayed as an example, although aging in polymers is a very specific process).
Taking into account the relation between D and the radius of micropores in
polymers [140], it must be concluded that in long-term relaxation of PTMSP sizes of
pores decrease (see Figure 9) and, probably, the mobility of macromolecular chains
reduces due to free volume decrease.
As follows from the constancy of ID, the concentration of positronium traps is
independent of the exposure time in the studied time interval.
Let us now discuss the results of investigation of positron annihilation in
polyimide.
As the measurements have shown [48], annihilation of positrons in polyimide
is significantly different from the one usually observed in most polymers. The
annihilation spectrum in polymers is usually characterized by the presence of three or
four components with average lifetimes from 100 ps to 4 ns [54, 164, 187]. However,
the different structure of the spectrum is observed for polyimide. It displays a single,
short-term, component with 0 = 0.388 ns (Figure 10). Time distribution is
approximated well by a single decay line, the tangent of which determines the average
lifetime.

Figure 10. Positron lifetime spectrum of the starting polyimide film (here N is the number of
readings in a channel)

The value of lifetime and the spectrum structure allow a supposition that
annihilation in polyimide proceeds from the positron state without forming a
positronium atom as it is typical of metals and semiconductors with high mobility of
electrons and a regular crystalline structure.
In this meaning, polyimide forms an electron structure unique for polymers,
characterized by high values and high homogeneity degree of the density function for
electrons.

54

Figure 11. Lifetimes and intensities of components (%) in the spectra of the original sample (I) and
deformed samples of polyimide after recovery lasting for 1 (II) and 24 (III) hrs.
Table 10
Sample
Initial
Deformed
Deformed

Annihilation characteristics of polyimide film


Recovery
0, ps
1, ps
2, ps
lasting, hr
3855

1
29430 44017

24
36110 53130

I2, %

595
92

Count rate,
k109, s

0.600.15
0.120.05

In relation to interaction with positrons, the microstructure of the initial


(undistorted) polyimide film possesses no defects. However, time spectra change after
deformation (Figure 11 and Table 10). Two components instead of a single one are
observed in the deformed sample: with shorter and longer lifetimes. After recovery
(resting) during 24 hours at room temperature, an increase of lifetimes of both
components and reduction of intensity of longer-term ones are observed. The
character of changes taking place allows a supposition that the submolecular structure
of polyimide is rebuilt during deformation; intermolecular bonds break, and
microdefect free volumes enough for positron localization are formed. In this case,
the value of the long-term component 2 must reflect changes in the average size, and
intensity I2 concentration of these defects. Analogous changes in the spectra were
also observed in annealing defects in metals and semiconductors. These changes are
usually analyzed with the help of a positron entrapment model. This model is
qualitatively good in reflecting changes in the time spectra observed in polyimide
deformation. Reduction of the lifetime of the short component, bound to annihilation
in the undistorted part of the polymer, depends on the high rate of capture in the
deformed sample. After partial contraction during recovery, the concentration of
defects decreases and lifetime 2 approaches the characteristic one of the original
polymer. Therewith, the intensity of the long-term component, I2, formed due to
positron annihilation on defects, decreases, too. Growth of the lifetime 2 may be
explained by coagulation (consolidation of small defects into larger ones) during
recovery or fast relaxation of small pores and, consequently, by growth of the average
capture radius.
As indicated in estimations, the concentration of microdefects after partial
relaxation decreases more than 7-fold. Therewith, the free volume induced by
deformation decreases by a factor of 4 [48]. The values obtained indicate that two
processes proceed fusion of microdefects and relaxation of the smallest ones,
though, apparently, the intensity of the latter process is higher.

55

Hence the one-component spectrum is typical of the original polyimide film.


In deformed samples, at least two components are observed in time spectra, which are
bound to the positron annihilation from the free state and the one localized in
micropores, formed at stretching. The lifetime increases and the intensity of the defect
component decreases during relaxation.
The results obtained with the help of the model of positron capture describe
clearly the changes of time distributions observed and allow a supposition that the
structure of the free volume during relaxation changes not only as a result of fast
recombination of the smallest pores, but also because of their consolidation with the
formation of long-term large-size microcavities.
Basing on the analysis performed in ref. [48], the following model of positron
annihilation and relaxation mechanism bound to it are suggested: before deformation
all positrons, captured in small traps with the bond energy slightly higher that the heat
energy, annihilate; after deformation, rather long (compared with the positron
diffusion length) areas occur, in which the concentration of small traps (of the size
~10 nm) decreases significantly, loosened up areas with deep centers of positron
capture are formed simultaneously in which the lifetime of positrons is longer;
relaxation happens in the way that pores formed during deformation recombine and,
moreover, increase when consolidate.
Hence, measuring the lifetime of positrons, the data on changes in structure of
the free volume occurring after polymeric film deformation may be obtained.
However, interpretation of the information obtained requires a detailed study of the
nature of components of a complex time spectrum of annihilation typical for a nonequilibrium state of polymer. No solution of this problem with the help of one of the
positron methods was obtained [3, 110, 156]. That is why a complex study of positron
annihilation was performed [49] in deformed polyimide with the help of measuring
the lifetime of positrons and angular correlation of annihilation radiation.
Two series of experiments are described in ref. [49]. In the first series, a
polyimide film was stretched by 20%. Then, the film was set free and relaxed freely.
Lifetime spectra for the freely relaxed film were measured every 1.5 hours.
Parameters of angular distribution were determined every hour during the day.
Table 11
Change of annihilation characteristics of polyimide film depending on duration of relaxation
after deforming by 20%
Lifetime
Angular correlation
Relaxation lasting
I21.5,
after deforming, h
%
FWMH
avg1,
110,
10.07, 0.07, I1.5,
ps
ps
mrad
mrad
%
0.05, mrad
0
365
201
74.3
10.44
10.49
7.14
28.2
1
360
176
73.6
10.77

5
368
208
77.2
10.60

24
362
205
73.0
10.48
10.64
7.14
34.7
240
364
200
74.1
10.43
10.72
6.95
32.3
Separated
368
220
76.3

Note. avg, 1 and I2 are characteristics of positron lifetime spectra; FWMH is the full width on the
middle height of the full spectrum; 1 is FWMH of the first Gaussian; and I are characteristics of
the parabolic component of the angular correlation spectrum.

In the second series of experiments, stress relaxation at deformation 0 = 20%


was studied. The characteristics of angular distributions were determined for films
with fixed ends. Measurements were performed with the help of a device that
performs deformation of samples directly in the measurement chamber. Stress

56

relaxation curves (dependences of stress on time ) and recovery curves


(dependences of deformation on time ) were taken simultaneously.
The values of the positron lifetime obtained from spectra are shown in Table
11 and Figure 12. Similar to the above-described results of two-component analysis,
changes of annihilation characteristics, which then relaxed gradually to those typical
of the initial polyimide sample, were observed in the structure of the time spectrum,
approximated by three components, after deformation.

Figure 12. Positron lifetime spectrum as a function of relaxation time for freely relaxing polyimide
films (for designation see Table 11).

Three components were separated: the lifetime of the first short-term


components (170220 ps) significantly depend on relaxation time; as displayed by
investigations [49], the lifetime of the second one (38810 ps) is independent of or
weakly depends on the sample state. However, significant changes in the intensity of
this component are observed. The characteristics of the third component have not
changed during the experiment.
In the work cited, experiments on measuring the angular correlation were
performed (alongside the measurement of the positron lifetime). Making no detailed
analysis of the results of these measurements, note that in experiments with fixed ends
(under stress relaxation conditions) the free volume significantly increases after
deformation, and its further slow relaxation is displayed well, happened at the
sacrifice of a decrease of micropore concentration.
In most cases, changes of macro- and microparameters of the polyimide film
during stress relaxation and recovery after deformation were indicated by the method
of positron diagnostics. Non-monotonous changes in the characteristics of positron
lifetime spectra and angular distributions of annihilation photons during recovery
were observed. Two ranges of changes in positron-sensitive properties of polyimide,

57

associated with fast and slow relaxation processes, were separated, and differences
in the type of relaxation of the polymer microporous structure depending upon the
condition of deformation and rest were observed. The effects observed are stipulated
by formation of areas of the local defrosting of molecular mobility.
All these experimental facts indicate that the microporous structure of the
polymer is rearranged during stress relaxation; this is expressed by the redistribution
of the sizes of micropores and their merging. Hence the method of positron
annihilation allows not only estimation of the microporous structure of polymers, but
also following its change under mechanical loading.

Chapter III. Temperature coefficient of volumetric


expansion

The thermal expansion of solids is a consequence of anharmonicity of thermal


oscillations of the substance particles. The thermal expansion of polymers has a
number of peculiarities connected with various physical transitions occurring in the
polymer as temperature is increased. To estimate experimentally the temperature
coefficient of volumetric expansion, the temperature dependence of the specific
volume of the polymer is determined. Schematically, this dependence is depicted in
Figure 13.

Figure 13. Schematic representation of the dependence of specific volume V on temperature T


(dilatometric curve) (rate of heating q1 > q2 > q3 > q4).

This dependence as a broken line is typical of many polymers near the glass
transition temperature, Tg. At temperatures below the glass transition temperature this
dependence is flatter than in the range of temperatures above it. Hence if T < Tg, the
temperature coefficient of volumetric expansion (which represents a tangent of
dilatometric dependence) is smaller than when T > Tg. In the first case, the
temperature coefficient of volumetric expansion is designated as G, and in the
second one L. In this connection, the specific volume of the polymeric substance
may be calculated by equations
V = Vg[1 + G(T Tg)],
V = Vg[1 + L(T Tg)],

(T < Tg);
(T > Tg),

(III.1)
(III.2)

where Vg is the specific volume of the polymer at the glass transition temperature; T is
temperature.
The dilatometric dependence shown in Figure 13 is rather simplified. In fact,
we are dealing not with a broken line, but with a curve called dilatometric. The
curvature of dilatometric dependences may be ambiguous. First of all, transition from
the glassy state into the rubbery one is characterized not by an abrupt fracture on the
dilatometric curve but by a smooth transition of one branch of the dilatometric curve
into another. This is clearly seen from Figure 14 which displays an experimental
dilatometric curve for polystyrene, determined near the glass transition temperature
(105C) of this polymer. The glass transition temperature itself is determined by
intersection of tangents of two branches of the dilatometric curve. Secondly, if the

59

dilatometric curve is determined in a wide temperature range, we may ensure that in


the area below the glass transition temperature it is not linear all the way, but indicates
a clear curvature.

Figure 14. Dependence of specific volume V on temperature T for polystyrene.

Figure 15. Dependence of the temperature coefficient of volumetric expansion G on temperature T for
poly(methyl methacrylate).

According to this curve, with decreasing temperature the coefficient of volumetric (or
linear) expansion is not the constant of the polymeric substance. Figure 15 displays
the experimental temperature dependence of the coefficient of linear expansion for
poly(methyl methacrylate), determined in a wide range of temperatures [154]. It is
easy to verify that the value of G decreases with temperature, i.e. the dilatometric
dependence at T < Tg is not linear. In this case, to calculate the volume of the
polymeric substance, it is not enough to use equation (37), but it is necessary to turn
to a more general relation
1 V
G =
,
(III.3)
V0 T
where G is the thermal coefficient of volumetric expansion depending on
temperature; V0 is the specific volume of the polymer near the absolute zero.
Knowing this dependence, the specific volume of a polymeric substance at any
temperature T may be calculated by equation (III.3).
Dilatometric dependences are not only of practical meaning as the ones
allowing searching of the glass transition temperature of polymers. They are also
theoretically valuable. First of all, slope changes not associated with the polymeric
substance transition from the glassy into the rubbery state are observed at some
temperatures below the glass transition temperature. These transitions occur at
temperatures below the glass transition temperature (e.g. inside the glassy state area)

60

and are of a somewhat different nature as compared with the main transition. Hence,
temperatures of these transitions may be determined by the dilatometric curve.
Secondly, according to the concept developed by Boyer and Simha, and formulated
by Flory, the transition from the glassy state into the rubbery one takes place at the
temperature at which part of the free volume in the polymer becomes the same and
equal to fc = 0.025. Due to this concept, the following relation holds [205]:
(L G)Tg = 0.113.

(III.4)

This relation is a rough approximation, because it takes into account no


curvature of the dilatometric dependence. Taking into account this curvature enables
Simha [154] to refine the free volume concept and to determine it with higher
accuracy. However, even in this case, the concept is just a rough approximation,
although it allows a description of the glassy staterubbery state transition.
Table 12 displays experimental values of G for a series of glassy polymers.
The lower the glass transition temperature, the higher is the coefficient of thermal
expansion. This correlates with the SimhaBoyer concept and equation (III.4). Hence,
heat-resistant polymers displaying high glass transition temperatures possess lower
G, and traditional polymers softening at low temperature display higher coefficients
of thermal expansion which, as a consequence, depend on the chemical structure of
the polymer.
Table 12
Calculated G,calc and experimental G,exp values of thermal coefficients of volumetric expansion
and the glass transition temperature Tg for a series of glassy polymers
Tg, K
Polymer
G,exp104, K1
G,calc104, K1
Poly(methyl methacrylate)
378
2.69
2.55
Poly(ethylene methacrylate)
338
2.99
2.84
Poly-n-propyl methacrylate
308
3.19
3.05
Poly-n-butyl methacrylate
293
3.34
3.22
Poly(methyl acrylate)
293
3.03
2.80
Polystyrene
378
2.50; 2.83
2.50
Polycarbonate based on bisphenol A
423

2.27

In ref. [35], the problems discussed above had been studied in detail for
polymer networks based on epoxy resins. Cured bulky samples were obtained using
epoxy resin ED-20, methyltetrahydrophthalic anhydride as a curing agent, and azelaic
acid (to elongate linear fragments between network cross-linked points), and oleic
acid (to obtain suspended chains). It was found that the coefficients of molecular
packing for cured networks based on epoxy resins are higher than those for linear
polymers. This is typical of systems containing no dangled chains (branches), i.e.
when azelaic acid is used as a co-curing agent. At room temperature, the coefficient of
molecular packing for them is almost independent of the network composition, and its
average value is kavg = 0.694, which is somewhat higher than the average value kavg =
0.681 for linear glassy polymers. Therewith, the average coefficient of molecular
packing of cured networks at their glass transition temperature is kg = 0.681, which is
also greater than kg = 0.667, typical of linear polymers. According to the data of these
measurements
(L G)Tg = 0.106.

(III.5)

61

Calculations and measurements have also indicated that for cured epoxy resins
the fraction of the free volume, formed due to thermal expansion, is 0.078. This value
is calculated from the formula
f =

Vg V0
Vg

= G Tg ,

(III.6)

where Vg and V0 are specific volumes of the polymer at the glass transition
temperature Tg and near the absolute zero, respectively.
The same value is determined from the relation
k
f = 0 1,
kg

(III.7)

where kg and k0 are the coefficients of molecular packing at the glass transition
temperature Tg and near the absolute zero, respectively. For linear polymers, the value
of f equals 0.096.
As mentioned above, the thermal expansion of substances is a consequence of
anharmonicity of thermal oscillations of the substance particles. On this basis, it can
be suggested that the coefficient of thermal expansion consists of contributions of
various oscillations of these particles.
Above all, the role of a weak dispersion interaction must be taken into
account. It should be noted that every atom is characterized by self-dispersion
interaction, which depends on both the type of the atom and its surrounding atoms, i.e.
on the atoms chemically bonded with it.
In the calculation scheme [28, 43], to calculate the coefficient of thermal
expansion, it was suggested that contributions of each atom are proportional to the
part of the Van-der-Waals volume Vi of it in the total Van-der-Waals volume
Vi of the repeat unit of the polymer.
i

Moreover, it is essential to take into account the influence of strong


intermolecular interactions, which appear in the presence of various polar groups in
C O,
the repeat unit of polymer. To them are corresponded, first of all, ester
O

nitrile CN groups, and various halogens which substitute hydrogen atoms (CHCl,
CHF, CF3), etc. These groups cause dipoledipole interactions of various types.
The most significant influence is also caused by polar groups, which lead to
occurrence of hydrogen bonds. They are, for example, amide NH C
,
urethane

NH C O

, hydroxylic OH, acidic

O
C OH groups.

O
O
Clearly, the energy of hydrogen bonds, similar to the dipoledipole
interaction, will depend on the chemical structure of polar groups. Seemingly, their
contribution to the coefficient of thermal expansion must be different. However, if
different parameters to characterize the energy of strong intermolecular interaction are
introduced for each type of the dipoledipole interaction and hydrogen bonds, this
will not only make the calculation scheme more complicated, but will also make
impossible calculations of the coefficient of thermal expansion for polymers

62

containing new polar groups. That is why discussion in works [28, 43] was limited by
the first approximation, according to which contribution of any dipoledipole
interaction is defined by the same parameter d, independent of the chemical structure
of the polar group.
However, since the Van-der-Waals volume of each polar group is different, it
will be seen in discussion below that the contribution of each polar group to the
coefficient of thermal expansion is also different. Concerning hydrogen bonds, we
may also confine to a single parameter h, which characterizes the energy of hydrogen
bonds.
The exception is only the class of polyamides which have a specific behavior
and require several parameters h characterizing the energy of hydrogen bonds.
Consequently, a relation to calculate the thermal coefficient of volumetric
expansion was obtained for polymers existing in the glassy state as follows:

i Vi + j
G =

Vi

(III.8)

where i are partial coefficients of thermal volumetric expansion, stipulated by weak


dispersion interaction of the i-th atom with the neighbor atoms; Vi is the Van-derWaals volume of the i-th atom; j are parameters characterizing contribution of each
type of specific intermolecular interaction (dipoledipole, hydrogen bonds) to the
coefficient of thermal expansion.
Let us consider the physical meaning of parameters i which characterize the
weak dispersion interaction.
It is well known that the coefficient of volumetric expansion is described by
the relation

i =

3 R i

i2 r0, i

(III.9)

where R is the universal gas constant; i is the anharmonicity coefficient,

i = 1

3
r 3 r
0,i

; is the potential of the i-th atom interaction with the adjacent ones; i

is the harmonic force constant, i =

2
r 2 r
0,i

; r0,i is the distance between the

considered i-th atom and adjacent atoms.


To estimate the coefficient i and the harmonic force constant i the Lennard
Jones potential can be used

(r) = D[(r0/r)12 2(r0/r)6].

(III.10)

In equation (III.10), the value D characterizes bond energy, and r0 is the


equilibrium distance between atoms, unbonded chemically, but participating in the
intermolecular interaction. Then, it may be written that

63

i =

72 Di
r02,i

; i =

756 Di
r03,i

; i =

7 R
7 R
.
; Di =
16 Di
16 i

(III.11)

Table 13
Values of constants i, j, ai and bj for various atoms and types of intermolecular interaction
Atom or type of intermolecular
ai103,
bj103,
Symbol
Symbol
i
interaction
K1
3K1
Carbon
0.00
aC
0.02

C
Hydrogen
19.98

1.92
aH
H
Oxygen in the backbone
22.95

2.21
aO,m
O,m
Oxygen in the side group
aO,s

O
16.00
1.54
O,s

=O
0.77
8.00
aO,s
O,s
Nitrogen in the backbone
8.62

0.83
aN,m
N,m
Nitrogen in the side group
6.35

0.61
aN,s
N,s
Chlorine
4.01

0.39
aCl
Cl
Fluorine
6.90

0.66
aF
F
Sulfur in the backbone
7.50

0.72
aS,m
S,m
Sulfur in the side group
2.04

0.20
aS,s
S,s
Silicon in the backbone
8.30

0.80
aSi,m
Si,m
Silicon in the side group
0.20

0.00
aSi,s
Si,s
Boron in carboranes
10.00

0.96
aB
B
Dipoledipole interaction*

55.4
5.31
bd
d
Hydrogen bond**

139.6
13.44
bh
h
Type of substitution of benzene rings***
bp
25.6
2.41

parap
16.0
1.54

methabm
m
16.0
1.54

orthobo
o
Coefficient for polydienes
12.96

135.0
b

Aliphatic cycle

120.0
11.52
bcycle
cycle
* Parameters d and bd are introduced for each branching in the main or side chain; they are also
introduced in the presence of a polar group of any type; if aliphatic polymers possess two CH3-groups
or two atoms of F or Cl at carbon atom, then constant bd is neglected.
For fragments - CH (bd = 51) - CH (bd = 32) - CH (bd = 51) - CH - (bd = 32)
|
|
|
|
COOCCSSC||
||
||
||
O
O
O
O
the additional constant are introduced shown in brackets.
** Constant bh is introduced in the presence of a hydrogen bond of any type for all polymers, except
polyamides; for the latter, constants h and bh are shown in Table 18.
*** Constants p, m, o and bp, bm, bo are introduced at substitution of aromatic rings in para-, methaand ortho-positions, respectively; the number of these constants equals to the number of substituted
rings. In the case of

structure, 2p and 2bp are introduced.

The expressions (III.11) allow estimation of the energy of the dispersion


interaction for each atom. These values are shown in Table 13. They indicate that
values Di really correspond to the energies of the intermolecular interaction, but not to
energies of the chemical bond. However, if these values are estimated by other
methods [66] (designate them as Di0), it is found that values Di differ several times
from Di0. This happens because every atom in any low-molecular substance or

64

polymer is in intermolecular interaction not with a single atom, but with several
atoms, with which it coordinates. Then, it should be written:
Di = zDi0,

(III.12)

where z is the coordination number.


In this case, the total average energy of the weak dispersion interaction is
determined from the relation z<D0>, where

< D0 > =

Vi
i

1 D0,i Vi

).

(III.13)

For copolymers, equation (III.8) gives:

1 i Vi + j + 2 i Vi + j + ... +

j
j
i
1
i
2

G =

1 Vi + 2 Vi + ... +

i
1
i
2

+ n i Vi + j

j
i
n

+ n Vi

i
n

(III.14)

where 1, 2, , n are molar parts of the components 1, 2, , n (do not mix these
values up with the values i present in brackets in the expressions);

V + , V + , , V + represent the

i
i
j
i
i
j
i
i
j

j
j
j
i
1 i
2
i
n

selection of constants i, j for the components 1, 2, , n of copolymer; Vi ,

i
1

Vi , , Vi are Van-der-Waals volumes of the repeat units of these

i
n
i
2
components.
In reduced form, the relation (III.14) can be written as follows:

65

k =n

k =1

k i Vi + j

G =

k =n

k Vi
k =1
i
k

(III.15)

If it is advisable to express the thermal coefficient of volumetric expansion G


of the copolymer via analogous coefficients G,1, G,2, , G,n for appropriate
homopolymers, the substitution of (III.8) into (III.14) gives:

1 G,1 Vi + 2 G,2 Vi + ... + n G, n Vi

i
1
i
2
i
n
,
G =

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n

(III.16)

where 1, 2, , n are the molar parts of the components 1, 2, , n of the


copolymer.
In reduced form, the relation (III.16) becomes

G =

k =n

k =1
k =n

k G, k Vi

k Vi
k =1
i
k

(III.17)

It should be noted that the dependence of G on the copolymer composition


is not linear but has a curvature dependent on the energy of intermolecular interaction
of components and their Van-der-Waals volume (Figure 16).

Figure 16. Schematic representation of the dependence of the thermal coefficient of volumetric
expansion G on the composition of copolymer .

The calculated values of the coefficients of thermal expansion of a series of


polymers in the glassy state are shown in Table 12. Generally, it should be taken into

66

account that, despite the seeming simplicity of this physical characteristic, its
experimental determination is rather difficult. That is the reason why greatly different
values of G for the same polymer can be found in the literature. All calculated and
experimental values of G shown in Table 12 characterize a part of the dilatometric
straight line which directly adjoins the glass transition temperature.
With regard to the thermal coefficient of volumetric expansion L in the
rubbery state, as mentioned above, it can be determined with the help of relation
(III.4), although significant errors are possible in this case. The monograph [214] by
Van-Krevelen indicates another relation for estimating L:

L = l/VM = l/M,

(III.18)

where l = 103VM, VM is the molar volume (per repeat unit of a polymer);


VM = N A Vi , where NA is the Avogadro number, Vi is the Van-der-Waals
i

volume of the repeat unit; M is the molecular mass of the repeat unit; is the polymer
density.
Taking into account that according to the data by Van-Krevelen VM = 1.60VW
in the rubbery state, it follows from the relation (III.18) that the coefficient of thermal
volumetric expansion for polymers in the rubbery state is the same and equals

L 6.3104 K1.

Chapter IV. Glass transition temperature of polymers

IV.1. Thermomechanical and other methods of evaluation of the glass transition


temperature of polymers
The thermomechanical method of studying polymers is one of the most
widespread methods of experimental determination of glass transition temperature Tg.
This method has been developed by V.A. Kargin and T.I. Sogolova. The essence of
the method is the following. A polymeric substance is subjected to a permanent or
variable load. Therewith, its deformation at each temperature and selected time of the
load effect is recorded. It is known that if a polymeric sample is subjected to a
constant stress, creep develops in it. Graphically, this looks as depicted in Figure 17.
To perform these experiments under comparable conditions, deformation must be
measured during a strictly constant time of observation, which may be any but
advisable so that reading of deformation would performed on the second, flattened
part of the creep curve. Having performed such an experiment at different
temperature, a temperature dependence of deformation may be plotted which, in the
general case, will be the same as depicted in Figure 18.
Before we proceed to consideration of features of the graphic in Figure 18,
note that an analogous curve may also be obtained in the case, when the sample is
subjected to effect of a low constant force, and temperature rises permanently with
time. Such a view of thermomechanical tests is most widespread at present, with
temperature usually increasing in accordance with a linear law
T = T0 + t,

(IV.1)

where T0 is the temperature of the onset of the experiment; t is time; is the rate of
temperature growth with time.

Figure 17. Schematic representation of creep (dependence of deformation on time t at constant


stress).

Usually, the standard rate of temperature growth is from 1.5 to 4 deg/min. This
method of thermomechanical investigation is more suitable for the apparatus,
although it gives less information than the first one. This concerns especially the cases
when fine structural transformations occurring in the sample at heating are studied by
the thermomechanical method. We will return to this question below, and let us now
consider the thermomechanical curve of the ideal polymer displaying no structural

68

transformations at heating. Such a curve is shown in Figure 18, and it is obtained in


the case when an experiment starts from rather low temperatures.

Figure 18. Schematic representation of the thermomechanical curve (dependence of deformation on


temperature T) on which I is the glassy state of polymer, II is the rubbery state of polymer
and III is the viscous flow state of polymer.

In this temperature range, deformation of the sample is small and increases


insignificantly with temperature. When the temperature definite for each polymer is
reached, deformation begins growing rapidly. If the load is low, increase of
deformation proceeds not infinitely but ends quite fast. A plateau occurs on the curve,
which is called the rubber-like plateau. The extent of this plateau on the temperatre
axis may be quite long and, as indicated below, it depends on the molecular mass of
the polymer. Further heating induces again a sharp increase of deformation, and the
polymer sample spreads.
The thermomechanical curve depicted in Figure 18 clearly divides the
temperature area of the test into three intervals. In the first of them, the polymer is in
the solid, glassy state (if it is amorphous). In this state, deformation is rather low and
grows only slowly with temperature: the polymer behaves itself almost like lowmolecular substances.
The first sharp increase on the thermomechanical curve is connected with the
polymer transition from the solid glassy state into the rubbery state. This state is
characterized by acquiring rubbery properties by the polymer, i.e. it possesses the
ability to develop high reversible deformations under low stress. This state is typical
only for polymers, and it is not displayed in the case of low-molecular weight solids.
At further heating, a sharp increase of deformation is connected with
occurrence of the viscous flow state, for which viscous flowing of the polymeric
substance is typical. Appropriate temperatures of transitions from the glassy state to
the rubbery state and from the rubbery state into the viscous flow state have received
the names the glass transition temperature and the flow temperature. Before we
discuss the nature of each of the physical states of polymers, note that depending on
the chemical structure of the polymer, i.e. on flexibility or rigidity of its
macromolecules, the glass transition temperature may obtain different values. At the
present time, there are polymers which display the glass transition temperature from
123 to 600C. An example of the first of them is polydimethylsiloxane with the
following formula:
CH3
Si
CH3

O
n

69

An example of the second one is polynaphthoylenebenzimidazole of the


following chemical structure:
N

...

C
N

N
C

...
N
O
Synthesis of polymers with the lowest possible glass transition temperatures is
of great importance for obtaining polymers resistant to cold. Synthesis of polymers
with the highest possible glass transition temperature is of the same importance for
obtaining solid heat-resistant polymers, capable of service at increased temperatures
and stresses. Many polymers, applied at present, possess glass transition temperatures
lying inside this broad interval. Shown in Table 14 are some of them and, considering
these data, the influence of the chemical structure of polymers on their glass transition
temperature could be easily imagined.
It should be noted that transition from the glassy state into the rubbery state
appears not at a definitely stated temperature but in some temperature range which
may be several tens of degrees wide. This is quite clear because the transition
considered is not the phase transition (such as melting, for example) but represents a
physical transition from one phase to another, keeping one of the phases. In the case
of amorphous glassy substances, this phase is liquid with no regard to the aggregate
state of the polymeric substance glassy (solid), rubbery or viscous flow state.
Fundamentally, every polymer, if its molecular mass is quite high, may exist
in one of the three physical states. Rubbers differ from plastics by the only thing that
the glass transition temperature of polymers, on which they are based, is below room
temperature, and for plastics above room temperature. However, if a rubber is
cooled down to temperatures below the glass transition point (for example, to 80C),
then a rubber-like substance becomes solid and behaves itself similar to usual glassy
polymers.
Vice versa, if a solid glassy polymer (for example, poly(methyl-methacrylate))
is heated up to a temperature above its glass transition point (above 100C), it
becomes a rubber-like material and behaves itself as usual rubber at room
temperature. Hence, the subdivision of polymeric substances into rubber-like and
solid glassy ones is rather conditional and has been formed historically.
Table 14
Glass transition temperature Tg of representatives of various series of polymers
Name
Structural formula
Tg, K
1
2
3
1,4-polybutadiene
170
CH2CH=CHCH2
150
Polydimethylsiloxane

CH3
Si

CH3
202

1,4-polyisoprene

CH2

CH C

CH2

CH3

70

1
Polyisobutylene

CH3
CH2

3
199

C
CH3

Polyvinylmethyl ether

CH2

258

CH
O CH3

Polyvinylethyl ether

CH2

248

CH
O C2H5

298

Polyvinylacetate

CH2

CH
O C

CH3

O
Polyvinylchloride

CH2

CH

355

Cl
Poly(methyl methacrylate)

CH3
CH2

378

C
C O CH3
O

Poly(ethyl methacrylate)

CH3
CH2

338

C
C O C2H5
O

Poly(hexyl methacrylate)

CH3
CH2

268

C
C O C6H13
O

Poly(methyl acrylate)

CH2

CH

293

C O CH3
Poly(butyl acrylate)

CH2

O
CH

233

C O C4H9
Polystyrene

CH2

O
CH

378

71

1
Poly(vinyl alcohol)

CH2

3
358

CH
OH

353

Poly(ethylene terephthalate)

Polyacrylonitrile

C O

CH2

(CH2)2

O
418

CH
C N

Polyacrylamide

CH2

426

CH
C NH2
O

Polycarbonate based on
bisphenol A

422

CH3
O

O C

CH3
Polyester based on sebacic
acid and phenolphthalein

O
368

(CH2)8

C O
C

O
C
Polyester based on
isophthalic acid and
phenolphthalein

O
543

C O

O
C
Polyamide based on sebacic
acid and aniline-phthalein

O
451

(CH2)8

C NH

NH
C

O
C
Polyimide based on
pyromellitic dianhydride
and aniline-phthalein

CO

783

CO
N

N
CO

CO
O
C

72

A polymer cannot exist in the gaseous state, because it is much easier to disrupt a
chains of macromolecules than to overcome total forces of intermolecular interaction:
polymer chains are extremely long and possess numerous intermolecular bonds
between neighboring chains; that is why it is impossible to transfer the polymer to the
gaseous state without decomposing it.
It was mentioned above that in each of three physical states the polymer may
exist only if its molecular mass is quite large. Then, at heating up a solid plastic, it
transforms consequently from the solid glassy state to the rubbery state, and then
to the viscous flow state (Figure 18). If the molecular mass of the polymer is low, it is
transferred from the glassy state directly to the viscous flow state, and the
thermomechanical curve has the form schematically represented in Figure 19. Such a
curve is also typical for low-molecular weight substances. Hence, the rubbery state is
typical for polymers only and is not displayed by low-molecular weight substances.

Figure 19. Schematic representation of thermomechanical curve for a polymer with small molecular
mass on which I is the glassy state of polymer and II is the viscous flow state of polymer.

Figure 20. Schematic representation of a series of thermomechanical curves for polymers having
different molecular mass M1 < M2 < M3 < M4 < M5 < M6 .

Now it is necessary to determine the molecular mass at which the polymer


begins displaying the rubbery state. To accomplish this, let us consider a series of
thermomechanical curves of polymer-homologues with various molecular masses.
Figure 20 represents schematically such a series; it is clearly seen that with growth of
the molecular mass the glass transition temperature increases at first and then
practically stops changing. The molecular mass, at which this change stops, is the
segment molecular mass. From this molecular mass, the polymer begins indicating the

73

rubbery state. This is quite clear, because the segment is that part of a macromolecule
the ends of which are able to slide independently. Consequently, if a macromolecule
is longer than the segment length, thermal motion may be performed in its separate
places only, not touching the macromolecule as a whole. This causes typical rubbery
properties, whereas high reversible deformations without significant mechanical
stresses are typical for the polymer. This is the rubbery state.
Clearly, if the molecular mass of the polymer exceeds the segment value, its
further increase cannot cause growth of the glass transition temperature, because
motion of segments as separate parts of macromolecules is not displayed yet. On the
contrary, the temperature of polymer transition to the viscous flow state grows
permanently with molecular mass, because flow is nothing but slide of separate
molecules relative to each other on the whole. It is clear that the longer
macromolecule is, the greater is the heat energy required for moving macromolecules
relative to each other, i.e. to induce flow. That is why the temperature of transition
into the viscous flow state increases permanently with molecular mass.
Therefore, it is absolutely clear that the value of the mechanical segment can
be determined with the help of the thermomechanical method. This value will be
correspond to the molecular mass at which the rubbery state occurs, and exceeding of
which may not increase of the glass transition temperature. Figure 21 represents the
dependence of the glass transition temperature on molecular mass Mp for polystyrene.
This dependence confirms the validity of the above discussion and is described well
by some relations, the most widespread among which is the Flory equation
Tg = Tg, /M,

(IV.2)

where Tg, is the glass transition temperature at molecular mass tending to infinity;
is a parameter; and the equation

1
1
=

,
Tg Tg, M

(IV.3)

where is a parameter.

Figure 21. Dependence of the glass transition temperature Tg on molecular mass Mp for polystyrene.

Let us now discuss experimental and calculation methods of determination of


the mechanical segment value, i.e. molecular mass Ms, starting from which the

74

rubbery state occurs. As mentioned above, one of suitable methods of experimental


determination of the mechanical segment value is the thermomechanical method.
Considering again Figure 20, it may be said that the molecular mass of the
segment determined by the thermomechanical method equals M3.
The value of the segment depends on the chemical structure of the
macromolecule, and the more rigid macromolecule is, the higher the value of its
segment is. The lowest value of the segment is displayed by flexible macromolecules,
which display quite free rotation of separate units relative to each other. Table 15
shows molecular masses of segments for macromolecules of various polymers. It is
clearly seen from this Table that in dependence on the chemical structure of polymers,
their macromolecules display absolutely different sizes of segments. The shortest
segment is typical for polyisobutylene macromolecules, and the longest for
polyarylate macromolecules1. The most rigid-chain polymers possess extremely large
mechanical segments and, in many cases, the value of this segment equals the length
of a macromolecule. In other words, no separate parts could be moved in such
macromolecules, not touching the whole macromolecule. Since when heating such a
polymer macromolecules of the polymer move relative to each other as a whole, the
temperature of transition to the viscous flow state and the glass transition temperature
coincide for it, because the fluidity of polymers by definition is sliding of separate
molecules relative to each other.
One more interesting feature is noteworthy, clearly seen in Table 15. Different
polymers possessing, however, absolutely the same backbone and differing by side
substituents only, display different segment lengths. At first glance, this contradicts
the fact that the flexibility of macromolecules is not a consequence of their deflection,
but the result of rotation of separate units relative to ordinary bonds, i.e. the bonds
which link the backbone atoms (in the present case, they are carbon atoms in
polyisobutylene, polyethylene, polystyrene, etc.). However, this contradiction is just
apparent. Freedom (or restriction) of rotation of separate groups relative to each other
depends not only on the macromolecule backbone, but also on type of its side
substituents. The more bulky and polar the substituent is, all other factors being the
same, rotation of separate units is restricted. Hence, we must consider not only
backbone rigidity, but also kinetic rigidity associated with the restriction of rotation,
occurring due to the presence of bulky side substituents. In this respect, polystyrene is
typical; it has a phenyl ring as a side substituent in the repeat unit. Consequently, this
polymer possesses a larger mechanical segment compared with the other polymers
shown in Table 15.
Now, it is necessary to relate the molecular mass of the segment with the
parameters of the chemical structure of the polymer and with its typical temperatures
the glass transition temperature and the temperature of transition into the viscous
flow state. However, first of all, we must agree on the method of determining these
temperatures. If such a determination is performed thermomechanically, it is very
important to choose a method of obtaining these temperatures from the
thermomechanical curve.
The theory of the thermomechanical method, developed by V.A. Kargin and
G.L. Slonymskii, states that the glass transition temperature and the temperature of
transition to the viscous flow state must be determined in the manner depicted in
Figure 22. A value of deformation 0 is chosen, plotted on the x-coordinate and in
1

Mechanical segment of macromolecules, the point of our discussion, must not be identified with the
Kuhn segment, which characterizes flexibility of isolated chain and is determined from behavior of
macromolecules in solution.

75

relation to the height of the rubbery plateau. A line parallel to the abscissa axis is then
drawn and points of intersection of these lines with the thermomechanical curve give
the required glass transition temperature to the viscous flow states. Hence, according
to this definition, the glass transition temperature is the temperature at which
deformation under the effect of a specific load develops by the value 0. The
temperature of transition to the viscous flow state is the temperature at which
irreversible deformation (flow) changes by the same value 0. The value 0 may be
chosen arbitrarily, but it must not be too high in order not to exceed the rubbery
plateau height. In practice, a specific fraction of the rubbery plateau height is taken for
the value 0.
Table 15
Values of molecular mass of a segment Ms for different polymers determined by the
thermomechanical method
Polymer
Ms
10001200
Polyisobutylene

CH3
CH2

C
CH3
12000

Polyvinylchloride

CH2

C
Cl
3000040000

Polystyrene

CH2

200 000

Poly(ether ketone)
CH3
C

CH3

300 000

Polyarylate F-1

C O
O

O
C

O
C

The KarginSlonymskii theory based on the consideration of deformation of a


polymeric body model with the calculation of temperature change leads to the
following relation:

76

lg M = lg M s +

B Tf Tg
,
C + Tf Tg

(IV.4)

where M is the molecular mass of the polymer; Ms is the molecular mass of the
segment; Tf is the temperature of transition to the viscous flow state; Tg is the glass
transition temperature; B and C are parameters of the polymer.

Figure 22. Evaluation of temperatures Tg and Tf according to the value 0 chosen on the
thermomechanical curve in accordance with KarginSlonymskiis theory .

Equation (IV.4) links the molecular mass of the polymer with the molecular
mass of the segment and the difference of the temperature of transition into the
viscous flow state and the glass transition temperature (Tf Tg). Not dwelling on the
practical meaning of this equation yet, note that its parameters B and C have a clear
physical meaning. The parameter B is determined from the relation:
B = lg

t 0
,
0 0

(IV.5)

where is the constant stress applied to the sample during the thermomechanical test;
t0 is loading time; 0 is the deformation at which Tg and Tf values are determined (its
meaning is clearly seen from Figure 22); 0 is the pre-exponential multiplicand in the
FulcherTammann equation which links the viscosity of the system with temperature.
This equation has the form:
A

= 0 e T T0 ,

(IV.6)

where is the viscosity of the system at temperature T expressed in Kelvin degrees;


0 is the pre-exponential multiplicand; A and T0 are parameters of the material.
Hence, all the values composing the parameter B are determined. The constant
C in equation (IV.4) is found from the relation
C = 0.434

A
.
B

(IV.7)

77

It should be noted that the FulcherTammann equation describes well the


temperature region of the glassy state near the glass transition temperature. We are
now interested in this very region.
Let us now consider briefly the practical meaning of the KarginSlonymskii
equation. If parameters Ms, B and C are known, this equation allows determination of
the molecular mass of the polymer without its dissolving. This requires determination
of the thermomechanical curve for this polymer, determination of the temperature of
transition to the viscous flow state Tf and the glass transition temperature Tg and their
substitution into the relation (IV.4). Therewith, parameters B and C need not be
determined from equations (IV.5) and (IV.7). This only requires determination of
three thermomechanical curves of the polymer with the known molecular masses M
and solution of the system of three equations composed on the basis of the equation
(IV.4) in which the unknown quantities are Ms, B and C. The molecular mass of any
amount of polymer-homologues of the given series may then be determined.
Table 16
Determination of polymerization degree Ns
Ms [174]

M0

Ns [174]

Tg, K

Vi , cm /mol

Ns

Ns*

3460
15625
38073
5625
10000
24287
30246

28
56
104
54
68
86
100

124
279
366
104
147
282
302

195
199
378
171
200
298
378

20.60
41.30
66.00
36.48
48.90
47.73
58.05

128
165
366
136
175
259
351

112
144
320
119
153
227
307

Polymer
Polyethylene
Polyisobutylene
Polystyrene
Polybutadiene
Polyisoprene
Poly(vinyl acetate)
Poly(methyl
methacrylate)

However, it should be noted that basing only on the chemical structure of the
polymer, the value Ms of it could not be calculated by equation (IV.4).
The value of mechanical segment Ms of a macromolecule may be calculated
on the basis of the chemical structure of the polymer. The following expression for
calculation of Ms (or corresponding polymerization degree Ns = Ms/M0, where M0 is
the molecular mass of the repeat unit) has been obtained in the ref. [96]:
1/ 3

N s = const Tg N A Vi

(IV.8)

The value of Ns may be calculated on the basis of the chemical structure of the
repeat unit with the help of atomic approaches, because the glass transition
temperature Tg and the Van-der-Waals volume of the unit Vi (NA is the
i

Avogadro number) are estimated with their help. The constant in the formula (IV.8)
may be expressed via parameters of the polymeric system and also from the
experimental data for standard systems. The estimation gives const = 0.21, if
calibration is performed by polystyrene according to data obtained in ref. [177], and
const = 0.24 for polystyrene calibration according to refs. [174, 214].
Table 16 shows values of Ns obtained from the data of the ref. [174] and the
formula (IV.8), with various constants (Ns with const = 0.24, and Ns* with const =

78

0.21). If const = 0.21 the difference in the values obtained from the ref. [174] does not
exceed 10%.
So far, we have discussed such physical characteristics of polymers as the
glass transition temperature, the temperature of transition to the viscous flow state, the
value of the macromolecule segment, which were determined experimentally with the
help of the thermomechanical method of polymer investigation.
Definite difficulties are met when determining temperature ranges of the solid
(glassy), rubbery and viscous flow states of polymers by this method. This especially
concerns new polymers.
Let us consider generally the possible deformation behavior of polymers in
thermomechanical tests. Recall that under these conditions the sample is loaded at
increasing temperature. In most cases, the stress acts permanently during the
experiment and temperature grows linearly.
Fundamentally, the thermomechanical method of investigation allows
immediate determination of temperature ranges of all three physical states of the
polymer. However, the existence of one or another physical state and appropriate
temperature range may be determined reliably only if it is known that the polymer
studied behaves itself as a classic one, i.e. gives the classic thermomechanical curve
depicted in Figure 18. As it is observed in the considerations below, even if the form
of the thermomechanical curve coincides with the classic one, in estimation of the
properties of a new polymer it is not yet possible to determine unambiguously the
temperature ranges of physical states and even of the states themselves.
Before we consider this point, let us discuss some procedural questions. A
question which appears most often is about the method of determination of transition
points from the thermomechanical curve. As mentioned above, the following method
is suitable: a definite strain 0 is chosen, plotted from the temperature axis and from
the rubbery plateau. The glass transition temperature and the temperature of transition
to the viscous flow state will correspond to temperatures, at which one and the same
value 0 of rubbery and plastic strain occur, respectively.
This method is most correct but suitable only when the thermodynamic curve
is of the classic form with abrupt bends of the curves in transition temperature ranges.
Then, the change of 0 will not cause large shifts in determination of Tg and Tf. If
deformation develops more smoothly, then the adjusted transition points Tg and Tf
will be quite undefined. They will be sufficiently dependent on the value of 0 (Figure
22).
That is why another method is used in practice: values of Tg and Tf are
determined by cross-points of tangents to two correspondent branches of the
thermomechanical curve (Figure 23). In this case, values of Tg and Tf are less
dependent on the shape of the thermodynamic curve, and this method is warranted for
comparative estimation of polymers.
Comparing thermomechanical curves of a series of polymers, the glass
transition point may be defined as the temperature at which deformation is developed
by the value of a specific percentage of the rubbery plateau height. Then, for each
polymer this typical deformation will display different values, because heights of the
rubbery plateau are also different.
Selection of the determination method of Tg and Tf depends on the shape of the
thermomechanical curve of polymers, and any of these methods may be chosen under
comparative estimation of their properties. Of importance is just the comparison of
several polymers to be performed always by the same method.

79

Figure 23. Evaluation of temperatures Tg and Tf by the points of intersection of tangents to two
branches of thermomechanical curve.

Let us now describe deformation of polymers during thermomechanical tests.


Very often, especially in the initial estimation, a polymer is tested as a powder. The
powder may only be compressed, and as it is placed in a special cell, compression is
close to the hydrostatic one. First of all, it should be taken into account that a powderlike polymer is not yet a polymeric material, but its basis only. As a result of the
thermomechanical test, the substance disposed beyond the punch is characterized,
which transmits the load to the sample. That is why we obtain the thermodynamic
curve of not a polymeric material, but of a polymer powder. Since the powder may be
not densely packed in the cell (not due to the fault of a scientist, but because of its
properties), and in connection with probable existence of small amounts of solvent
and by-products of the reaction, the shape of the thermomechanical curve may be
significantly distorted. That is why estimation of physical states of the polymer on
powder samples is of the most rough type, especially in the case of new, unknown
polymers.
Even if the thermomechanical curve is of the classic shape (see Figure 18) and
consists of three parts, the statement that the polymer displays all three physical states
transiting from one into another at heating should be neglected. It should be taken into
account that growth of deformation in a powder sample may be caused by secondary
reasons. When the thermomechanical curve is determined, it is best to take a look on
the last branch of it. If it exists in the temperature range in which thermal and
thermooxidative degradation does not penetrate deep enough, flowing of polymers
may be discussed. To be certain that development of a large deformation (up to 100%
at compression) is caused by flow and not by deep degradation of the polymer,
thermogravimetric analysis (TGA) must be performed (thermogravimetric curve must
be obtained) simultaneously. This is most important in the case of heat-resistant
polymers for which high strains start to develop in the temperature range from 600 to
800C. So, this deformation, caused by deep thermal degradation of the polymer, may
be erroneously assumed to be the flow. It should also be taken into account that
besides degradation, network formation may also occur during thermomechanical
tests. These two processes always coexist at polymer heating, but one of them
proceeds with much higher rate and defines direction of the whole process.
Structuring may be displayed in formation of cross-links between chains of the

80

polymer, cycle formation, etc. Consequently, the started flowing of polymer will be
stopped, and a plateau appears on the thermomechanical curve, analogous by its shape
to the plateau of the rubbery state for linear polymers. That is why the existence of the
plateau (more correct, retarded deformation) is not the point to state that the polymer
in the present temperature range is in the rubbery state.
Ensuring that large deformation in the last region of the thermomechanical
curve is cause by polymer flow, it should be attempted to transfer it into a bulky
sample by any of existing methods (for example, by hot pressing). After that the
sample is again thermomechanically tested. If the shape of the repeated
thermomechanical curve is close to the initial one (obtained for the powder sample),
chemical changes, possibly occurring during processing, are insufficient. Now we
should only ensure that the plateau on the thermomechanical curve reflects the
rubbery state, and is not caused by secondary reasons. One of the typical features of
the rubbery deformation is its complete reversibility, as well as the ability to be
expressed at very low loads (remind that the rubbery modulus is usually three-four
decimal degrees lower than the modulus of elasticity for a glassy polymer). When a
plateau is found on the thermomechanical curve, it must be checked if deformation is
reversible in this temperature range. This is performed by regular loading of the
sample, with this load exceeding the initial one, sometimes several times. If at
application of additional load deformation increases sharply and recovers fast after its
removal, a reversibility may be considered. This is reflected in Figure 24. If values of
additional load and additional reversible deformation caused by it are known, the
rubbery modulus may be easily calculated dividing the additional load by the
deformation increment (Eel = /). Ensuring that the elasticity modulus is of the
smallest value (below ~1 MPa), it may be stated that the polymer in the plateau area is
really in the rubbery state. For rigid-chain polymers, the elasticity modulus will be
significantly higher (~10 MPa), and it is most important here to check reversibility of
deformation in the plateau area. Therewith, it is often observed that a part of
additional deformation is reversible (rubbery), and the rest of it is irreversible
(plastic). In this case, the elasticity modulus is determined by division of by the
increment of the reversible part of deformation, and it is hardly possible to speak
about the true rubbery state.

Figure 24. Checking of reversibility of deformation in the zone of the supposed rubbery-like plateau
(see text)

81

Figure 25. A lever mechanism with a variable arm for ensuring constant tension in the course of creep
process (scheme).

All the above-said is true when thermomechanical tests are also performed
under the effect of elongating tension. Secondary effects at elongation may not distort
the picture so significantly, as in the case of compression. As thermomechanical
investigations are performed at low elongation and the classic thermomechanical
curve is obtained, it must only be assured that the plateau reflects the true rubbery.
Checking is performed in an analogous way: an additional load is applied, and
reversibility of deformation is checked after its removal. If deformation is reversible,
it is rubbery. In thermomechanical tests on elongation, the tension should be constant.
Contrary to tests on compression, some difficulties are faced here. The cross-section
of the sample decreases at deformation and, consequently, stress in it grows under the
effect of constant force. To compensate the stress increment caused by developed
deformation, the load is transmitted to the sample via a lever mechanism with a
variable arm (Figure 25). The arm decreases as deformation is developed, and the
load decreases consequently so that the stress in the sample remains constant.
After making sure that the plateau on the thermomechanical curve reflects the
rubbery state, it can be said with confidence transition to the rubbery state has taken
place. The transition point is determined by the above-described methods.
Until now, we have discussed the thermomechanical curve of semi-crystalline
polymers, which is of the classic form and displays three physical states and two
transitional temperature areas between them. It should be taken into account that the
thermomechanical curve of crystalline polymers may be of the analogous shape (see
Figure 18). Besides thermomechanical investigations, X-ray analysis must be
performed in order to confirm semi-crystallinity of the sample. Neglecting so far
transitions in crystalline polymers, let us discuss separate cases of behavior of semicrystalline polymers under thermomechanical conditions of tests.
It may be immediately noticed that classic thermomechanical curves are not
always observed. Deviations of the thermomechanical curve shape from the classic
one are caused by different reasons. For example, the thermomechanical curve of a
semi-crystalline polymer is often of the shape shown in Figure 19. It is absolutely
obvious that this polymer displays no viscous flow state. It is also obvious that

82

deformation of the sample (under compression) does not reach 100% and,
consequently, it is burnt off at high temperatures, and the remaining part is distributed
below the punch and prevents it from reaching the bottom.
We should also carefully treat the temperature range, in which deformation
grows sharply. If deformation near the onset of the horizontal plateau is completely
reversible, it is of the rubbery type. However, it may not be stated without additional
experiments that the polymer transits to the rubbery state, and the curve uplift points
out the glass transition temperature. It is entirely probable that at this temperature
chains of the polymer are intensively cross-linked, which therewith deforms, but
subsequently loses meltability and solubility. That is why besides thermomechanical
tests, thermogravimetric analysis is obligatory.
Sometimes a polymer gives the classic thermomechanical curve, the last
branch of which is in the area of very high temperatures (700800C). Clearly,
development of large deformations in this area is caused not by viscous flow, but by
intensive thermal degradation. In this case, it must not be said about the temperature
of transition into the viscous flow state and the transition into the viscous flow state. It
is specially obvious in the cases, when the thermomechanical curve is of the shape
depicted in Figure 26. Distortion of the rubbery plateau (if it exists for the polymer
tested) is caused by thermal degradation proceeding, excretion of gas products, etc.

Figure 26. Thermomechanical curve of polymer undergoing thermal degradation during the course of
measurement (scheme).

Figure 27. Appearance of negative deformation on the thermomechanical curve (see text).

Under compression, the thermomechanical curve may also be of the shape


depicted in Figure 27. In this case, decrease and even appearance of negative
deformation is caused by increase of the sample height, as a result of which the punch
transmitting the load to the sample is slightly lifted. This event may be caused by

83

elimination of residual solvent at heating up powder-like or bulky samples, excretion


of gas products of degradation, etc. In such a case, determination of the true glass
transition temperature and the temperature of transition into the viscous flow state is
quite difficult.
One of the reasons for the appearance of negative deformation may be the
following. Preparing for thermomechanical tests, a polymer powder is pressed at
room temperature, which is usually well below the glass transition area.
Consequently, the polymer is subjected to the induced elastic deformation.
Temperature grows during thermomechanical tests and when it reaches the glass
transition area, the induced elastic deformation rapidly becomes reversible. The
sample expands and, if the test is performed under compression at low loading,
negative deformation appears on the thermomechanical curve. The higher the
pressure applied to the polymer powder at tableting, the greater the negative
deformation is and the greater is the distortion of the shape of thermomechanical
curve. In this case, the glass transition temperature corresponds to the lowering and
not uplifting branch of the thermomechanical curve.
In thermomechanical tests of polymers with elongation, deformation also
decreases sometimes as temperature grows (Figure 28). The sample contraction may
occur as a result of cross-linking, ring formation, etc., proceeding during the test.

Figure 28. Decreasing of deformation on the thermomechanical curve determining under conditions of
tension (see text).

In most cases, the thermomechanical curve consists of two (see Figure 19) and
not three parts. Usually, this is associated with the insufficient molecular mass of the
polymer and with the fact that the polymer transits from the glassy state to the viscous
flow one without reaching the rubbery state. Rising of the appropriate branch of the
thermomechanical curve defines the temperature of transition into the viscous flow
state, and it may be said that the glass transition (softening) and temperatures of
transition into the viscous flow state coincide. When the second branch of the
thermomechanical curve lies in the area of high temperatures, it should be confirmed
that no degradation of the polymer happens in this area, and sharp increase of
deformation is really associated with the polymer flowing and not with its
degradation. In the latter case, softening and flowing are not yet reached, and the
polymer exists in the glassy state in the whole possible temperature range.
When transitions from the glassy state to the rubbery state and from the
rubbery state to the viscous flow state are expressed clearly on the thermomechanical
curve (in temperature ranges of transitions the deformation sharply increases),
transition temperatures Tg and Tf can be easily measured. However, the
thermomechanical curve is often of the shape depicted in Figure 29. This
thermomechanical curve cannot help in determination of the glass transition

84

temperature: the transition is seriously blurred. That is why other methods should be
used. Transitions from the glassy state to the rubbery state and from the rubbery state
to the viscous flow state, and the melting point of polymers as well can be determined
by dilatometric, calorimetric, and optical methods; it is appropriate to use also
dynamic mechanical methods of investigation, etc. (see below).

Figure 29. Thermomechanical curve with blurred transition (scheme).

Let us now turn to crystalline polymers. In some cases, the shape of


thermomechanical curves of crystalline polymers differs from that of the same curves
for semi-crystalline polymers and, in other cases, they nearly coincide.
It is impossible to answer the question as to whether a polymer is semicrystalline or crystalline with the help of only the thermomechanical method of
investigation. An X-ray pattern should be obtained first, and then data of the X-ray
structural analysis must be compared with the results of the thermomechanical study.
Crystalline polymers may be obtained directly during synthesis and further processing.
For crystalline polymers, the thermomechanical curve may be of the shape
depicted in Figure 30. If a polymer is quite extensively crystallized, the rubbery state
is nearly completely suppressed, and the polymer displays no significant deformations
in a wide temperature range. When transited to the melting point, deformation grows
sharply and reaches the maximal value.

Figure 30. Thermomechanical curve of well-crystallized polymer (scheme).

The thermomechanical curve in Figure 30 corresponds to the case in which the


rubbery deformation is completely suppressed by crystallization. Under real
conditions, semi-crystalline areas exist in the crystalline polymer, which behave
themselves at heating as a semi-crystalline polymer. During thermomechanical
investigation of semi-crystalline polymers in the glass transition range, deformation
will increase with further formation of a plateau (Figure 31). However, the rubbery
plateau will not be of the same height as for pure semi-crystalline polymers of the

85

same structure. It will be significantly lower depending on the crystallinity degree. In


any case, the glass transition temperature and the temperature of transition to the
viscous flow state can be determined for the semi-crystalline polymer.

Figure 31. Thermomechanical curve of semi-crystalline polymer (scheme).

Particular emphasis should be placed on polymers, which easily crystallize


during the thermomechanical test. It is then best to perform the experiment with a
regular load2.

Figure 32. Schematic representation of the thermomechanical curve for a semi-crystalline polymer
determined at periodic variation of load (see text).

Rapid crystallization of a semi-crystalline polymer occurs at temperatures


above Tg. That is why the semi-crystalline polymer capable of rapid crystallization in
the glassy state behaves itself similar to the normal semi-crystalline polymer only
slightly crystallized. In the thermomechanical test, it gives a low deformation up to
the glass transition temperature. Passing through this temperature, the polymer
develops a large rubber deformation (Figure 32, curve 2). However, crystallization at
T > Tg makes the polymer rigid, and if the load is applied regularly, the deformation
decreases abruptly (see Figure 32, curve 2). Once the polymer has been deeply
crystallized, its deformation is not yet different from the deformation of the
preliminarily crystallized polymer (see Figure 32, curve 1). The temperature of
transition to the viscous flow states will also coincide (see Figure 32).
Let us now discuss briefly other methods of experimental estimation of the
glass transition temperature of polymers. One of them is the dilatometric method in
which the dependence of the specific volume of the polymer on temperature is
2

In all the cases, when more fine investigations are performed by the thermomechanical method, the
load is best to be applied regularly, giving rest to the sample, i.e. a possibility to recover deformation
after load removal. Then all effects of the glass transition, crystallization, etc. will be displayed more
clearly.

86

measured. Schematically, this dependence is depicted in Figure 13. This dependence


as a broken line is typical of many polymers near the glass transition temperature. At
temperatures below the glass transition temperature, this dependence is more
flattened, than at temperatures above it. This is the main reason observed at the glass
transition temperature (see Figure 13). Obviously, dilatometric dependences may be
obtained at different rates of heating (or cooling) of the sample. Consequently, if the
heating rate is changed in a quite wide range, the dilatometric dependences change in
the manner depicted in Figure 13. The higher the rate of the sample heating, the
higher is temperature at which the bending is observed and, vice versa, the lower the
heating rate, the lower is temperature of the bending and the specific volume of the
sample in the bending point decreases. Hence, the glass transition temperature
depends on the rate of heating or cooling of the sample, which is associated with the
relaxation mechanism of the glass transition (see below).
It should be noted that changing the heating or cooling rate causes has a
smaller effect on the type of dilatometric curve and on the glass transition temperature
than the effect of mechanical tension on the type of the thermomechanical curve and
the glass transition temperature. To cause a significant change of the glass transition
temperature in the first case, the heating or cooling rate must be changed by many
decimal degrees, which is impossible to do in an experiment. At the same time, ten- or
hundred-fold change of the heating rate changes the glass transition temperature by
several degrees only. Meanwhile, the change of the mechanical tension by several
percents only may cause a sharp increase or abrupt decrease of the glass transition
temperature. A special role in this case is played by the rate of mechanical influence.
Analysis of these experimental facts enabled G.M. Bartenev [53] to apply the
viewpoint that mechanical and structural glass transitions must differ. By the
structural glass transition one means transition from the glassy state to the rubbery
state (and the reverse process) in the absence of mechanical tension. This transition
and its temperature are associated only with the heating and cooling rate. Mechanical
glass transition is associated with the rapid mechanical influence on the polymer and
its behavior as a solid under the rapid action of the mechanical force. As temperature
changes, the mechanical glass transition may occur at rather lower temperatures than
the structural glass transition.

Figure 33. Temperature dependence of heat capacity Cp (scheme).

In addition to the dilatometric method of determination of the glass transition


temperature, other methods are also widely used, for example, calorimetric one. In
this case, heat capacity Cp of the polymeric substance is measured, the dependence of
Cp on T is of the typical shape (Figure 33). In the area of the glassy state, heat

87

capacity grows slowly with temperature, but when Tg is reached, a sharp increase of
the heat capacity is observed. When the transitional zone is passed, the heat capacity
again increases only slowly with temperature.
Apparently, dynamic mechanical analysis is used most widely. In this method,
the temperature dependences of the storage modulus E and the loss modulus E of
the complex elasticity modulus E* = E + iE change, as well as the loss-factor tgE =
E/E (Figure 34). The temperature dependence of tgE displays several maxima, with
the most intensive (and high-temperature) one associated with transition from the
glassy state to the rubbery state.

Figure 34. Temperature dependence of the mechanical loss-factor tgE (scheme).

The temperature dependence of the dielectric loss-factor tg = , where


and are the true and the imaginary parts of the complex dielectric constant * = +
i, looks the same (Figure 35).

Figure 35. Temperature dependence of the dielectric loss factor tg (scheme).

Finally, the glass transition temperature may be determined from a sharp


increase of the diffusion coefficient. Neglecting discussion of these methods of
experimental determination of Tg, let us guide the reader to special literature [5, 51,
124].

88

IV.2. Mechanism of glass transition


The mechanism of glass transition has been studied for many years, but there
is no general viewpoint on the glass transition mechanism. That is why we shall
consider main common viewpoints on this mechanism. Note that these viewpoints do
not contradict to each other.
First of all, let us consider the concept of the relaxation mechanism of glass
transition, put forward by A.P. Alexandrov, Yu.S. Lazurkin, P.P. Kobeko, and G.I.
Gurevich. Let us preliminarily note once again that transition to the glassy state and
the opposite transition to the rubbery state are not phase transitions of the first and the
second kind (remember that according to the Erenfest definition, the phase transition
of the first kind is the phase transition at which thermodynamic functions display a
sharp jump at the transition temperature, and the derivative of the temperature
dependence of the thermodynamic function is ruptured). The phase transition of the
second kind is the transition at which temperature dependences of thermodynamic
functions at the transition temperature change smoothly, and a sharp jump at the
transition temperature is displayed only by first derivatives of these functions.
Externally, in the case of polymers, some temperature dependences of
thermodynamic functions look as if they were typical of transition of the second kind.
For example, the temperature dependence of the specific volume (dilatometric curve)
possesses a bend at the glass transition temperature by which this temperature is
identified (see Figure 13). However, this resemblance is only superficial, because the
transition considered is not associated with appearance of a new phase, as in the case
of melting or crystallization from the melt, but is the transition from one physical state
to another keeping the unique liquid phase. Hence, by the phase state solid glass
represents a liquid, like elastomers and flowing polymers. They differ only by the
aggregate state, representing the unique liquid phase.

Figure 36. Location of kinetic unit in the minimum of potential energy (scheme).

Let us now turn to consideration of the relaxation mechanism of glass


transition. Let us begin description of this mechanism from low-molecular substances.
Each molecule of such substance occupies a definite location in space and oscillates
due to thermal fluctuations, i.e. as a result of exceeding the average temperature by
the temperature of the present point. To transit to another location in space, a definite
amount of energy should be consumed. Let the kinetic unit (atom or molecule) exist
in the location, when the potential energy of it is minimal. This case, displayed in
Figure 36, corresponds to the stable equilibrium. To transit the kinetic unit to a new
state of equilibrium, it is necessary to overcome some energetic barrier, which is the
real transition energy and is called the potential barrier. Hence, the potential barrier is
the energy, which must be spent by the kinetic unit in order to transit to a new
location. Let us consider the composition of this barrier.

89

Every thermodynamic system is characterized by the value of the free energy


or isobaricisothermal potential F:
F = U TS,

(IV.9)

where U is the internal energy of the system; T is absolute temperature; S is entropy.


The value of the free energy U is composed of the energy of translation of
particles, the energy of their rotation, and the oscillation energy of particles in relation
to their equilibrium. Concerning entropy, according to the Boltzman definition, it
equals
S = kBlnW,

(IV.10)

where kB is the Boltzman constant; W is the thermodynamic probability of existence


of the system.
The notion of entropy as a statistic characteristic is best demonstrated on the
example of various conformations of macromolecules. Let us consider one of the
border cases, a completely stretched chain (Figure 37,a). Clearly, such chain may
possess only a single conformation, and the distance between its ends h will be equal
to the full length of the macromolecule. Probability of existence of this conformation
is extremely low, because it just the single one. But if we take a globular chain
structure (see Figure 37,b), it will obtain a multiplicity of conformation at the same
distance between its ends, i.e. its elements may locate differently in space keeping this
distance constant. The shorter this distance h (down to some limit), the more
conformations of the macromolecule may be composed and greater is the probability
of existence of the macromolecule in this (globular) form. Since according to the
Boltzman definition entropy is proportional to the logarithm of thermodynamic
probability, it will increase with decrease of the distance between macromolecule
ends. According to the second law of thermodynamics, a spontaneous process always
proceeds with entropy increase. That is why thermal motion in the case of flexible
macromolecules causes their coiling.

Figure 37. Schematic representation of straightened (a) and coiled (b) polymer chain (h is the distance
between the chain ends).

Turning back to the system composed of small molecules, note that transiting
of kinetic unit from one state into another both the internal energy and entropy will
change. Change of the free energy will be determined from the relation
F = U TS,

(IV.11)

90

where U is the change of the free energy; S is the change of entropy.


The probability of the given kinetic unit being in a specific position in space
depends on its energy and absolute temperature. This probability is determined by the
relation
W = 0eE/RT,

(IV.12)

where 0 is the self frequency of oscillation of atoms or molecules; E is the energy of


the particle; R is the universal gas constant; T is the absolute temperature.
Equation (IV.12) is used in cases when calculation is performed per mole of
the substance. If the calculation is performed per molecule, kB (the Boltzman
constant) should be introduced instead of R. These values are connected by the
relation kB = R/NA, where NA = 6.0231023 mol1 (the Avogadro number).
It is common knowledge that the fundamental frequency of oscillation of
atoms in solids is ~1013 Hz, and the oscillation period is determined from the relation
0 = 1/0. It is of interest that the time of settled life of an atom in this location is
10101011 s. Considering transition of a polymer from the glassy state to the rubbery
state, let us introduce such characteristic as the time of this transition. This time tr is
determined from the relation

tr = tr,0eU tr RT ,

(IV.13)

where tr,0 is a pre-exponential multiplicand which represents the value reversible to


the fundamental frequency of oscillations of atoms; Utr is the energy, which must be
spent to perform the transition.
The smaller tr is, the faster the transition is, i.e. the higher is temperature and
the lower is transition energy Utr, the easier the transition conduction is.
Let us consider a rubbery polymeric body, cooled down at a definite rate of
temperature change. If the time of transition of structural elements (kinetic units) from
one state to another is short at a definite temperature, such transition is fast and
change of the structure will proceed immediately after the temperature change.
It may be said that if the transition time tr is much shorter than the
observation time obs, then the transition will proceed very fast and the equilibrium
state of structure of the polymeric body is reached very rapidly immediately after the
temperature change. The heat energy, characterized by the RT product, decreases with
temperature, and the relaxation time (see equation (IV.13)) grows significantly.
However, if it remains shorter than the time of system observation, the transition
proceeds quite rapidly. Further decrease of temperature will lead to a sharp increase of
tr and to fulfillment of the condition obs tr, and then obs < tr. This corresponds to
the case when the observation time becomes much shorter than the transition time.
Clearly, this transition cannot be observed, although it proceeds. The transition
proceeds so slowly that observation of it becomes difficult. That is why the polymeric
substance becomes solid, i.e. transits to the glassy state. It should be noted once again
that the structure of the polymeric substance at this transition is in fact not
equilibrium. It is somehow frozen in the state preliminary to the transition. If we
could manage to observe the transition for a very long time, it would be noted that this
transition proceeds, although very slowly. Hence, according to these notions, the
transition from the rubbery state into the glassy state is the relaxation process and is
associated with the fact that at temperatures definite for every polymer the relaxation

91

time (time of transition) becomes extremely long, and the polymeric substance begins
to behave as a glassy solid.
This transition theory was subsequently mathematically justified by M.V.
Volkenstein and O.B. Ptitsyn, who had performed mathematical analysis of the
relaxation transition from the rubbery state to the glassy state and back, and had
adjusted conditions under which such transitions may be performed. According to this
theory
ln

tr,0U tr
RTg2

U tr
1
= lg ,
RTg
q

(IV.14)

where q is the rate of system cooling.


Equation (IV.14) shows that the glass transition temperature depends on the
rate of polymer cooling.
In its initial form, the VolkensteinPtitsyn theory does not consider the
cooperative character of thermal motion of kinetic units. As shown subsequently, such
calculation is absolutely necessary, because due to large length and flexibility of
macromolecules the local motion of any kinetic unit touches upon its neighbors. For
example, motion in the frames of a single repeat unit of a macromolecule causes
motion of the neighbor units and this is the reason for segmental mobility. Calculation
of the cooperative ability of motion of kinetic units of various sizes, performed by
Ptitsyn and Sharonov, caused refinement of the VolkensteinPtitsyn theory and
allowed reasonable values of the activation energy of the relaxation process to be
determined.

Figure 38. Schematic representation of thermomechanical (a) and dilatometric (b) curves in the zone
of transition from the glassy state to the rubbery state (see text) .

Analysis of a polymeric substance with the help of a non-linear model [87, 89]
shows [90] that transition from the glassy state to the rubbery state must proceed not
in a point but in some interval of temperatures which depends on the rate of heating or
cooling q. Schematically, this is shown in Figure 38, which depicts thermomechanical
and dilatometric curves in the transition area. Here Tg,1 is the temperature, at which
transition from the rubbery state to the glassy state begins at system cooling; Tg,2 is
the temperature at which this transition ends.
The results obtained in work [90] indicate that at low rates of cooling the glass
transition temperature range degenerates and the glass transition temperature only
remains. In the general case of analysis of polymer behavior in the area of T < Tg,1, a
significant role is played by the cooling rate. If the cooling rate is high, temperatures

92

Tg,1 and Tg,2 differ significantly, i.e. transition to the glassy state proceeds in a wide
temperature range.
In practice, the glass transition temperature range is really broadened as the
rate of cooling (or heating) increases, not in a jump-like manner, but quite smoothly.
Therewith, the temperature dependence of deformation in the transition zone becomes
more abrupt as the rate of cooling (or heating) decreases. This is clear from Figure 39,
which indicates thermomechanical curves for a series of cured compounds based on
epoxy oligomers. In accordance with the theoretical results obtained, the glass
transition temperature range windes as the heating rate increases (Figure 40).

Figure 39. Thermomechanical curves of cured epoxy compound measured under conditions of
compression. Rate of heating 0.62 (1), 1.07 (2), 2.39 (3), 4.61 (4), 6.39 (5), 9.44 (6), and
15.47 (7) deg/min. Composition of specimen: ED-20 1 mole, MTHPA 0.6 mole,
azelaic acid 0.7 mole.

Figure 40. Dependence of temperature differences T = Tg,1 Tg,2 on heating rate q: (a) correlates with
the same composition as in Figure 39; (b) corresponds to the composition of specimen: ED20 1 mole, MTHPA 1.3 mole, oleic acid 0.7 mole.

The nonlinear model of polymeric substances [87,89] was also used for
estimating the activation energy of -transition in polymers.
Estimation of many properties of polymers requires the values of the
activation energies associated with various transitions in polymeric systems. If a
polymer is considered as a linear system, then it will be a selection of activation
energies, usually associated with the motion of a definite group of atoms in a chain.
But if it is considered as a non-linear system, described by a model suggested in ref.
[87] (Figure A-6-8), we must know the activation energy of elementary acts only,
associated, for example, with a jump of atoms from one equilibrium state to another in
a rotary isomer, and which is usually associated with the activation energy of the transition. All other relaxation times stipulated by other transitions are expressed via
the desired activation energy and temperature coefficients determined by non-linear

93

model parameters [269]. Hence, a temperature dependence of activation energy in the


area of -transition may be obtained, and the WilliamsLandelFerry formula may be
deduced, etc.
To use actively this model for calculations of particular polymeric systems, the
relation of the parameters of the model with the chemical structure of the repeat unit
must be known.
Let us demonstrate how expressions for determining activation energy which
participate in the temperature dependences of relaxation times (viscosity) may be
deduced from correlations of the ref. [269]. It is shown below that the activation
energy is linked by simple dependences with the glass transition temperature and,
therefore, it may be calculated on the basis of the chemical structure of the repeat unit.
Let us consider expressions for relaxation times 3 and 5 of a non-linear
model, shown in ref. [269], in the range of temperature of the onset Tg,2 and
termination Tg,1 of glass transition. With this aim in view, let us perform series
expansion of 2/5 and 2/3 in relation to Tg,1 and Tg,2, respectively, and confine to
terms of the smallness order one by T Tg,1 and T Tg,2. Then,

2 d 2


(T Tg,1 ) ,
5 dT 5 T =T
g ,1
because

2
5

= 0 and
T =Tg ,1

2
3

= 0,
T =Tg , 2

where 2 is the relaxation time of one of elements of the model, which may be
expressed by the ArrheniusEiring formula.
The expressions 2/5 and 2/3 may be expressed via variable 0, respectively,
as follows:

2 d 2
d 0


(T Tg ,1 ),

5 d 0 5 T =T dT T =T
g
,
1
g ,1
2 d 2
d 0


(T Tg, 2 ) .

3 dT 3 T =T dT T =T
g,2
g,2
d 2
d 2


and
are of
As according to ref. [269], values
d 0 5 T =Tg ,1
d 0 3 T =Tg , 2
the same order of magnitude, future consideration will take into account only the
d
( 0 ) at Tg,1. As
expressions 2/5, and Tg will be taken for Tg,1. Let determine
dT
according to ref. [269],

94

16 K
0 = 1 1 , K1 = f1e Y1 , K 3 = f 3e Y3 ,
21 K 3
Y1,3 = 1 2 1 T Tg1,3 ,

then
d 0 16 2 K1 1
1
1
1
= 0

.
dT 21 K 3 Tg3 1 Y3 Tg1 1 Y1
The relation

K1
K3

= 57/32, 0 = 2.8 [269]. K1 value corresponds to the


T =Tg ,1

elasticity of a chemical bond, and K3 to the elasticity modulus of an intermolecular


bond, that is why Tg,1 = Td (to the temperature of onset of intense thermal
degradation), Tg,3 = Tm (to the melting point). Then,

d 0
dT

T =Tg ,1

T
7.6
1
=
m

Tm
Tg,1 Td
1 T
m

1
.
Tg,1
1
Td

d 2

0.3 may be found. Taking into account the
From ref. [269]
d 0 5 T =Tg ,1
expression

d 0
and the value
dT

d 2


, the dependence 2/5 on temperature
d 0 5 T =Tg ,1

in the area of Tg (suggesting Tg = Tg,1) may be presented as follows:

T
2
2.3
1
1

m
T Tg .

Tm
Td
5
Tg
Tg
1
1 T
Td
m

(IV.15)

The expression obtained, according to ref. [269], describes the transition to the
glassy state. It may be displayed that the Bachinski formula is also reduced to an
analogous form. Actually, to describe the viscosity of liquids , Bachinski [768] has
suggested an empirical formula:

C
.
v

(IV.16)

95

Here v is the specific volume of the liquid; is the molecular limited volume of the
liquid (according to Bachinski, M = i , where i are constants of atoms forming
i

the molecule; M is the molecular mass), C is the constant. The expression (IV.16)
describes well the liquid viscosity in a wide range of temperatures and pressures, the
constant C may be found from the CamerlingOwnnes theory of similarity of
molecular systems. Bachinski has shown [768] that for many liquids the following
correlation is fulfilled:
CM 1/ 6
= 0.531104,
1 / 2 1/ 3
Tg

(IV.17)

Thus, the accuracy of determination of C depending on the chemical structure is


6.2%.
Let us transform the expression (IV.16). For this purpose, let us consider the
temperature dependence of the specific volume, which may be presented in the
following form:
v = [1 + v(T Tg)],

(IV.18)

where v is the coefficient of volumetric expansion which, according to ref. [269] in


the case of the potential 6 12, is equal to 1/42Tg. With regard to equation (IV.16),
correlation (IV.18) has the following form:

0 T Tg
=
,

42Tg

(IV.19)

where 0 = C/. Expression (IV.19) will be analogous to equation (IV.15) obtained


from the non-linear model, therewith, 0 = 2 = K12, and = K15. But contrary to
the Bachinski formula deduced for transition from liquid to glass, the desired
expression describes the transition from the side of low temperatures (transition from
the glassy state to the rubbery state). That is why

Tg
Tm
1
1

2 = 2.3
420 .
Tm
Tg Td
Tg
1
1 T
Td
m

(IV.20)

From equation (IV.20) the expression for the activation energy may be
deduced. Let us estimate the following value for the real polymeric systems:

Tg
T
1
1
m
A = 2.3

Tm
T
Tg
Tg
d
1
1 T
Td
m

42 .

(IV.21)

96

Shown in the Table are initial data (glass transition temperatures Tg, melting
points Tm, temperatures of the onset of intense thermal degradation Td), and results of
calculation of values A for a series of polymers. Values of Tg, Tm and Td are usually
taken from experimental data but, in some cases, when, for example, a polymer is not
crystallized or when its melting point lies above the temperature of onset of intense
thermal degradation and thus may not be measured experimentally, they are
calculated by the expressions shown below.
Calculation results indicate (Table 17) that the value A falls within the range
from 24 to 82. In some cases, it becomes negative. It is typical of polymers, for which
temperature of the onset of intense thermal degradation lies near the glass transition
temperature, or the melting point and the glass transition temperature are close, i.e. for
heat-resistant polymers. As obvious, coefficients A calculated for the most of
polymers are positive that, according to equation (IV.20), correspond to positive
values of viscosity. Similar to a non-linear model of polymeric body, in which
viscosity is associated with relaxation mechanisms, according to the Mandelshtamm
Leontovich theory, the viscosity in our expressions will be the second one, because
the non-linear model suggested [269] describes behavior of a polymeric substance by
a selection of kinetic equations. However, as the Table indicates, for some polymers
A < 0. The latter means that the second viscosity of these polymers is negative. The
result obtained has a physical meaning and, as follows from ref. [768], is associated
with nonequilibrium of the medium in the range of the glass transition temperature.
This result must be analyzed in more detail.
E *
h
expression (IV.20) may
With regard to 0 = C/ and 2 = 2 exp
kTg
Nl q

be presented in the form:


Nl 2 q AC
E *
,
= ln

kTg

h

(IV.22)

where N is the number units in the macromolecule; l is the unit size; q = exp(E/kT);
E* is the difference in the energies of rotary isomers.

97

Table 17
Polymer
1

CH2

Initial data and calculated results of value A for a series of polymers


Tm, K
Td, K
Tg, K
(exper./calc.)
2
3
4
263
449/451
673

Tg/Tm

Tg/Tm

Tm/Td

5
0.583

6
0.391

7
0.667

8
39.1

378

513/509

618

0.743

0.612

0.824

46.6

249

405/413

664

0.603

0.375

0.622

46.5

302

508/500

678

0.605

0.445

0.737

35.2

453

633/625

588

0.724

0.770

1.06

-21.5

205

332/314

663

0.654

0.309

0.474

71.4

CH
CH3

CH2

CH

CH2

CH
C2H5

CH2

CH
CH2

CH CH3
CH3

CH2

(CH2)4

CH

O C

C
O

97

(CH2)4

98

(CH2)3

O C

(CH2)4

O
CH2

2
214

3
311/315

4
659

5
0.679

6
0.325

7
0.478

8
77.6

254

359/387

704

0.655

0.361

0.550

64.1

260

417/398

704

0.653

0.369

0.565

62.2

150

234/234

771

0.640

0.194

0.303

82.2

CH
O CH2
CH2

CH3

CH
O CH3

CH3
Si O
CH3
CH2CH2
CH2

CH

CH3

O CH2

CH

213

410/410

713

0.519

0.575

0.299

37.8

246

438/397

651

0.620

0.379

0.609

50.9

CH3
CH2CH2O
CH2

206

339/348

794

0.592

0.259

0.438

60.4

198

348/356

742

0.556

0.267

0.480

50.5

194

309/311

721

0.622

0.269

0.431

67.4

CH O
CH3

(CH2)4O

98

99

2
318

3
433/465

4
629

5
0.682

6
0.506

7
0.739

8
47.6

223

323/334

655

0.667

0.340

0.510

71.2

351

530/539

621

0.651

0.565

0.868

23.7

323

499/494

622

0.654

0.519

0.794

35.1

323; 348

499/496

621

0.651

0.520

0.799

34.0

319

467/481

653

0.662

0.488

0.737

42.5

298

/458

640

0.651

0.716

0.466

44.8

CH3
C

CH2

C O CH3
O
(CH2)2

O C

(CH2)4 C O

O
C

(C H 2 ) 4

O
NH

(C H 2 ) 6

NH

(CH2)8

C NH

(CH2)6

NH

O
NH

(CH2)5

C
O

NH

(CH2)10

C
O

CH2

CH
O C

99

CH3

100

2
199

3
/300

4
713

5
0.663

6
0.420

7
0.279

8
78.6

620

/801

667

0.774

0.930

1.200

180.9

CH3
CH2

C
CH3

C
O

C O
O

O
C
O
C
O

100

101

Considering a polymer in the glassy state as a frozen liquid and using data of
the CamerlingOwnnes theory of molecular similarity, as Bachinski has done, the
expression for C may be presented in the following form in analog to formula (IV.17):
CM 1/ 6
= 10 b .
Tg1/ 2 1/ 3
Here b is the exponent value which, in the case of polymeric systems, is somewhat
dependent on chemical structure and will be found during comparison with particular

characteristics of polymers. Because ~ N Vi , where Vi


is the
i
r.u.
i
r.u.
1/ 3

Van-der-Walls

volume

of

the

repeating

unit,

l ~ Vi
i

and

C = 10 b Tg1/ 2 1/ 3 M 1/ 6 ,

ATg1/ 2 M 1/ 6
E *
.
= 2.3 27 b + ln
1/ 3

kTg
M
r.u.

(IV.23)

If for real polymer systems logarithm is taken from Tg, M, and Mr.u., average
by polymeric systems, it is observed that fluctuations stipulated by the difference of
parameters Tg, M, and Mr.u. for particular polymeric system from average ones, give a
low error (< 10%) for determination of the activation energy E* of this system. In
turn, the logarithm from average values of parameters by polymeric systems is taken
into account via constant b. That is why, expression (IV.23) may be reduced to the
form:
E* = 2.3(27 b)kTg.
In accordance with the experimental data from refs. [214], the value of b may be
estimated, which is equal to 15.7. Respectively,
E* 26kTg.

(IV.24)

*
*
Shown below are some calculated Ecalc
values
and experimental Eexper
of the activation energy for a series of polymers according to the data from refs.
[214].

Polymer

PEO

PCTFE

PE

PVA

Tg, K
*
Ecalc
,
kJ/mol
*
Eexper
,
kJ/mol

206
44.8

323
70.3

213
46.0

303
65.8

Butyl
rubber
200
45.6

37.7

72.7

50.3

60.8

49.9

102

*
*
Deviations between Ecalc
and Eexper
observed do not exceed ~15%.
Considering jumps of atoms or groups of atoms along the chain in rotary
isomers as diffusion of these atoms in polymer, the ratio E*/d2 may be determined
by data from ref. [214] (p. 298, Figure XVIII.7), where d is the size of the groups of
atoms diffusing. But E* is deduced from expression (IV.23), from which the size d
of the jumping group of atoms may be determined immediately. For example, in the
case of PS, E*/d2 = 600, Tg = 378 K, E* = 82.2 kJ/mol, and d = 5.7 that
corresponds to the group C6H5. However, jumps of a group of hydrogen atoms are
possible in PS. If sizes of these atoms are known, the activation energy may be
deduced from the expression E*/d2. It is equal EH* = 19.7 kJ/mol. From expression
(IV.24) the temperature of transition of this group of atoms Tg,1 may be estimated. If
calculated, Cg,1 = 91 K, which by the order of magnitude corresponds to the
temperature of -transition in PS (accurate coincidence is out of the question, because
the constant in expression (IV.24) is found for the -transition). Estimations
performed allows a supposition that if the repeating unit of polymer possesses side
(asymmetric) groups of atoms different by size, a -transition appears in such a
polymer.
Discussed above is one of the concepts of the polymeric substance transition
from the glassy state to the rubbery state and back. This concept is based on
assumption of the relaxation mechanism of transition, the time of this transition (the
relaxation time) depending on temperature and being determined by the energetic
barrier height to be overcome in order to provide with the transition. Since different
macromolecules possess different flexibility or rigidity, they are characterize by
different heights of the energetic barrier and, consequently, by different glass
transition temperatures. The more flexible the macromolecule is, the lower the height
of the energetic barrier is which characterizes the system composed of such
macromolecules.
The concept of the polymeric substance transition from the glassy state to the
rubbery state and back, based on the relaxation mechanism of transitions, is not
unique. Another concept that explains these transitions is the concept by S.N.
Zhurkov, based on acceptance of a significant role of intermolecular bonds in
polymers. Let us discuss this concept.
Every polymeric macromolecule possesses polar groups, which display
various energies of interaction with each other. They form bonds which are not
chemical, but are of the physical nature and are called intermolecular bonds. Such
bonds may be rather weak as, for example, in polyisobutylene, polypropylene,
polyethylene, and in other analogous polymers, but they may also be very strong as
for example, in polyamides, polyesters, etc. In the former case, it is spoken about
weak Van-der-Waals bonds, the energy of which is very low. In the latter case, strong
NH C
hydrogen bonds which appear due to existence of amide
groups (polyO
amides), and strong dipoledipole interaction which appear due to existence of ester
O
groups are considered. Intermolecular bonds may be stipulated by an
C O
existence of not only these polar groups, but also of other groups capable of strong

         
N groups (for example,
  
 -groups (for example, in poly(vinyl alcohol), and
many others. Polar groups in neighboring macromolecules under definite conditions
may interact with each other forming cross-linked bridges between them, which are

103

quite strong. Recall that we are dealing not with chemical bonds, but with physical
ones. Schematically, it is represented in Figure 41.

Figure 41. Schematic representation of a network of physical (intermolecular) bonds.

It is necessary to pay attention to one more very important circumstance.


Contrary to chemical bonds, intermolecular bonds are rather weak. Energy necessary
for dissociation them is by a decimal degree lower than that for chemical bonds.
Recall that the energy of chemical bonds dissociation is of several tens of kcal/mol,
and that of intermolecular bonds is several kcal/mol. The strongest bonds are
displayed by amide, ester, hydroxyl and some other groups. However, the energy of
dissociation of intermolecular bonds formed by these groups does not exceed ten
kcal/mol (in fact, this energy is lower).
Hence, intermolecular bonds are not stable, but decompose permanently under
the effect of thermal fluctuations and are formed in new places. They somewhat
migrate along the polymeric chain and in doing so link not strictly specified chains of
macromolecules, but the chains, in which they appear at the current moment.
However, at each definite temperature the amount of intermolecular bonds existing
simultaneously is much the same for polymers of a definite chemical structure.
It may be said that under equivalent consideration, intermolecular bonds act in
definite places of macromolecules forming a peculiar network and making impossible
their motion in relation to each other. Taking into account that, in fact,
macromolecular bonds are formed and dissociate permanently under the effect of
thermal fluctuations, an equilibrium may be shifted to one or another side, i.e. to the
side of formation or dissociation of intermolecular bonds. As with usual chemical
reaction, the equilibrium shifts to the side of decomposition of intermolecular bonds
as temperature increases, and to the side of their formation as temperature decreases.
When polymeric substance existing in the rubbery state is cooled, the amount of
forming intermolecular bonds grows permanently. When a definite temperature is
reached, the amount of bridge bonds formed reaches the value enough for fixing
separate macromolecules relative to each other and allowing not translation mobility
of them. Therewith, vibration mobility remains.
Formation of a quite frequent and strong network of intermolecular bonds with
polymer cooling down to a definite temperature promotes a definite loss of mobility
of macromolecules at this temperature, and the polymeric substance as a whole
becomes solid, i.e. transits to the glassy state. Temperature, at which this phenomenon
is observed, is the glass transition temperature. When polymeric substance in the
glassy state is heated, the equilibrium shifts to the side of intermolecular bond
decomposition. When the process of their dissociation is carried far enough,
macromolecules obtain translation mobility, and the polymeric substance softens, i.e.
transits to the rubbery state. According to the concept considered, this is the nature of
transition of polymeric substance from the rubbery state to the glassy state and back.

104

To perform this transition, a definite number of intermolecular bonds must be


formed in each polymeric substance with no dependence on their chemical structure.
Designate this number as x. Concentration of active groups in the polymer, capable of
formation of intermolecular bonds, are designated as N. In this case, taking into
account that intermolecular bonds dissociate and are formed permanently in other
places, and that an equilibrium between the number of formed and dissociated bonds
exists, to describe this process, a common equation of chemical kinetics describing
equilibrium process may be used:
x 2(N 2x).

(IV.25)

Left part of the equation includes the number of intermolecular bonds formed
at a definite temperature; right part of it in brackets represents the number of active
groups remained, capable of forming these bonds. Digit two appears in brackets,
because formation of each bond consumes two active groups. This is also the reason
of appearance of two before brackets. Since the reaction considered is equilibrium, the
law of mass action may be applied to it, according to which the relation of
concentration of the substance formed to concentration of the initial substance is a
constant value, equal to the equilibrium constant. Therewith, stoichiometric
coefficients of the reaction equation are included in indices of degrees of
corresponding concentrations. Hence, in the present case

(N 2 x )2
x

= k eq ,

(IV.26)

where keq is the equilibrium constant.


According to Arrhenius, temperature dependence of the equilibrium constant
is described by the following relation
k eq = A * e

E
RT

(IV.27)

where A* is a steric factor; E is the bond energy (in this case, the intermolecular
bonds energy); R is the universal gas constant; T is absolute temperature.
Substituting (IV.27) into (IV.26), we get
E

x = (N 2 x )2 A * e RT .

(IV.28)

Equation (IV.28) associates the number of intermolecular bonds formed with


the number of active groups capable of formation of these bonds and with
temperature.
To clear up the mechanism of glass transition, it is necessary to interlock a
part of polar groups one way or another in order to exclude them from the interaction.
Interlocking may be performed by different methods, for example, by an insignificant
change of the chemical structure of the macromolecule or by introducing small
molecules of low-molecular substances acting as plasticizers. They may be introduced
by sorption of gaseous small molecules of a sorbate by polymer. Assume that n of
such small molecules were introduced into a polymeric substance, so that the same
number of polar groups were interlocked. In this case, (IV.28) will be written as:

105

x = (N 2 x n )2 A * e RT1 ,

(IV.29)

where n is the number of solvent molecules.


To fulfill the equilibrium condition and the number of intermolecular bonds
formed to become x again, temperature must be changes, i.e. it must be decreased to
decrease the value in brackets in the right part of equation (IV.29). That is why, T is
substituted by T1 in this equation.
Solving jointly the system of equations (IV.28) and (IV.29), we obtain the
following relation
n E T1 T

.
2 ln1
=
N 2 x R T1T

(IV.30)

In equation (IV.30) T correlates with the glass transition temperature of the


ZLWK WKH JODVV WUDQVLWLRQ WHPSHUDWXUH RI
original non-plasticized polymer, and T1
plasticized polymer, because by the condition x is the number of intermolecular bonds
necessary for formation of a firm spatial network. When this network is formed, the
polymer transits to the glassy state.
Equation (IV.30) may be used for description of transition of a polymer from
the solid state to the rubbery state, when a definite amount of plasticizer is introduced
into it. However, this equation may be simplified significantly. First, assume that the
number of intermolecular bonds is significantly smaller than the total number of polar
groups capable of intermolecular interaction (as experiments have indicated, such
happens indeed in practice: to obtain a firm network of intermolecular bonds and to
cure the polymer, it is necessary to fix no more than 10% of polar groups). In other
words, the condition is fulfilled:
N >> 2x.

(IV.31)

Let us introduce the second assumption, namely, operation with a small


number of plasticizer molecules (clearly, such assumption might be not introduced).
Then it will be written that
n/N << 1.

(IV.32)

Taking the logarithm of the equation (IV.30) and taking into account that
under the condition (IV.32) ln(1 n/N) = n/N, we get
2

n E T1 T
=
.
N R T1T

(IV.33)

If a low amount of plasticizer molecules is introduced into the polymer, the


glass transition temperatures of the initial and plasticized polymers will not be
T T1 T
=
. Under
significantly different. So it may be assumed that T T1 and
T1T
T2
these conditions, the relation (IV.33) will obtain the form

106

n E T
=
.
N R T2

(IV.34)

Therefrom,
T =
or

2nRT 2
EN

(IV.35)

T = n,

(IV.36)

2 RT 2
.
EN

(IV.37)

where

Equation (IV.36) reflects the Zhurkov rule of molar parts. According to this
rule, decrease of the glass transition temperature of plasticized polymer compared
with the glass transition temperature of the initial polymer is proportional to the molar
part of introduced plasticizer. This rule means that independent of chemical structure,
depression of the glass transition temperature is the same, if equal molar parts of
plasticizer are introduced. This is quite clear, because according to the present
concept, solidification of polymer (transition from the rubbery state to the glassy
state) proceeds at formation of equal amount of intermolecular bonds. If a part of
polar groups is eliminated from the process in one way or another, then decrease of
the glass transition temperature will depend only on the amount of these groups, but
not on chemical structure of interlocking molecules.
Table 17
Values of energies of intermolecular interaction for various polar groups
E, kcal/mol
Polymer
Group
2.25
Poly(methyl methacrylate)

C
Polyacrylonitrile
Polyisobutylene
Polyethylene

O CH3
CN
CH3
CH2

2.22
0.67
0.32

Experiments and calculations performed by S.N. Zhurkov et al, lead to values


of energies of the intermolecular interaction for various groups, shown in Table 17.
This Table indicates clearly that energies of the intermolecular interaction are
low, by a decimal degree lower than energies of chemical bonds. The highest energy
of intermolecular interaction is displayed by hydrogen bonds and by bonds appeared
at the sacrifice of interaction of carbonyl groups with nitrile groups, as well. The
lowest energy of intermolecular interaction (Van-der-Waals forces) is typical of

&+ JURXSV
2

The higher the energy of intermolecular interaction is, the greater amount of
thermal energy is required for overcoming these bonds and the higher the glass
transition temperature of polymers possessing strong intermolecular bonds is. The
lower the energy of intermolecular interaction is, the lower the amount of thermal

107

energy is required to failure these bonds and the lower the temperature of polymer
transition from the glassy state to the rubbery state is. Polymers such as polyethylene,
polyisobutylene, polypropylene, etc. display low glass transition temperatures. In turn,
polymers such as polyesters, polyamides, polyvinyl alcohol, polyacrylonitrile, etc.,
which possess strong intermolecular interaction, display higher glass transition
temperatures.
Clearly, this consideration neglects skeleton rigidity of the macromolecule, i.e.
possibility or impossibility for separate groups in the main chain to rotate around the
ordinary bonds. Besides, this factor may be decisive and cause a very significant
influence on the glass transition temperature of the polymer. One more important
circumstance is noteworthy. In the case of glass transition of low-molecular liquids,
molecules of which contain polar groups, all of them may interact with each other,
because no steric hindrances for this exist.
In the case of the glass transition of polymers, macromolecules are denuded of
this possibility due to their great length. Only in an ideal polymer crystal, free from
any distortions and dislocations, polar groups in the repeating units are capable of
complete interaction with each other. At the glass transition of a polymeric liquid (i.e.
at transition from the viscous flow state or the rubbery state to the glassy state),
structure of the polymer is far from the ideal crystal. That is why, due to steric
hindrances, only a part of polar groups may enter the intermolecular interaction.
According to tests on measurements of heat capacity of polymeric substances at
temperature change, it was found that only a small part of polar groups enters the
intermolecular interaction during polymer cooling down to the glass transition
temperature. The amount of these groups is ~10% of the total number of groups
capable of intermolecular interaction. And even this small amount of interacting
groups is enough for the polymeric substance to transit from the rubbery state to the
glassy state due to formation of a spatial network of intermolecular bonds.
Formation of these bonds at polymer cooling and their dissociation at its
heating may be detected by direct methods of physical study. The first among them is
the IR-spectroscopy method, with the help of which it was clearly displayed that at
polymer cooling down to the glass transition temperature or polymer heating above
this temperature an intensive formation or dissociation of intermolecular bonds
proceeds.
So detailed consideration of this concept is associated with the circumstance
that in the subsequent discussion describing calculation methods of estimation of the
glass transition temperature of polymers Tg various variants of appearance of the
intermolecular interaction will be considered, which affect on calculated values of Tg.
Actually, intermolecular interaction between polar groups located in neighboring
chains causes formation of a physical network of intermolecular bonds. However,
such interaction may also be performed between polar groups, located in the same
repeating unit of the macromolecule. Then these groups are excluded from the
interaction between the neighboring chains, and the glass transition temperature
decreases. The same effect will also be observed at an anomalous addition of
monomers during polymerization or copolymerization (head-to-head or tail-to-tail
attachment). In this case, intermolecular interaction is performed between polar
groups located in neighboring repeating units of the same chain; clearly, therewith
they are excluded from the interaction between neighboring chains, which causes a
decrease of the glass transition temperature. All these points will analyzed in detail
below.

108

Besides the Zhurkov rule of plasticization, discussed above in detail, there are
many other rules. For example, because viscosity of low-molecular fluidsplasticizers
in by many decimal degrees smaller than viscosity of polymer, it is assumed that
equal volumes of the fluid may equally decrease the glass transition temperature of
the polymer:
T =

v
,
V +v

(IV.38)

where V is the polymer volume; v is the plasticizer volume; is the coefficient equal
for all fluids.
Equation (IV.38) reflects the KarginMalinskii rule of volumetric parts. It
operates better for weak polar polymers, whereas the Zhurkov rule is more valuable
for strong polar polymers. The cases are often, when both rules are unable to describe
behavior of real polymeric systems. Let us dwell in more detail on this point below.

IV.3. Calculation of the glass transition temperature of linear polymers


This section discusses the calculation scheme for estimation of the glass
transition temperature Tg developed in works [6, 128]. According to this scheme, the
coefficients of molecular packing for polymers of various chemical structures are
approximately equal at the glass transition temperature of each of polymers, this value
kg being estimated by the value kg 0.667 for linear polymers. The coefficient of
molecular packing k0 near the absolute zero is also approximately equal for all
polymers and is 0.731.
These measurements indicate that the transition from the glassy state to the
rubbery state proceeds, when a definite part of empty volume is reached, equal fE =
1 kg = 0.333, and expression for the glass transition temperature may be presented in
the following form:
Tg =

k 0 k g 1 0.0906
=
.
G
G

(IV.39)

It should be noted that the material discussed in the present Chapter is tightly
connected with the material discussed in Chapter III. Using expressions for G
deduced in Chapter III, it must be written down that
R
1
1
=
,
Tg 0.455 ZD0
1

where

Z D

1
= i
ZD0

i 0i

Vi

Vi
i

(IV.40)

; Z is the coordination number of the i-th atom; D0i is

the bond energy of the i-th atom with neighboring ones; Vi is the Van-der-Waals
volume increment of the i-th atom.

109

Numerical values of 1/ZiD0i, typical of every atom and every type of


intermolecular interaction, were determined with the help of statistical treatment of
experimental data by the method of least squares. As mentioned above, the values of
bond energies obtained as a result of processing of the experimental data correspond
to energies of the intermolecular interaction. Presence of polar groups in polymers,
which causes strong intermolecular interaction, is taken into account by introduction
of special constants.
For practical calculations, it is advisable to rewrite the formula (IV.40) in the
following form:

Tg =

Vi
i

ai Vi + b j
i

(IV.41)

where ai is a selection of constants, which characterizes energy of a weak dispersion


interaction as the averaged contribution of each atom into this interaction; bj is a
selection of constants, which characterizes energy of a strong specific intermolecular
interaction (dipoledipole, hydrogen bonds, etc.).
Values ai and bj are associated with parameters i by the following relations:
ai =
bj =

i
i
=
;
k 0 / kg 1 0.0906
j
k0 / kg 1

j
0.0906

(IV.42)
.

(IV.43)

Using the relation (IV.41), the glass transition temperature of a tremendous


number of polymers may be calculated. This is associated with the circumstance that
the approach described is atomistic, i.e. each atom is characterized by its constant ai
(their values are shown in Table 13). As for specific intermolecular interactions
(dipoledipole, hydrogen bonds), they are characterized by their constants bj,
independent of chemical structure of the polar group. For example, dipoledipole
interactions of various types are characterized by the same constant bd = 55 103
3 K1. The situation with hydrogen bonds in polyamides is somewhat more
complicated. This is connected with specificity of their influence on Tg in the frames
of the present class of polymers (Table 18)*.
As an example, Table 19 indicates experimental and calculated values of glass
transition temperatures for a series of model polymers, starting with polymers of a
simple chemical structure to quite complicated systems. Selections of parameters of
the relation (IV.41) are also present there.
The approach developed [6, 128] and the relation (IV.41) obtained on its basis
allow taking into account the influence of the type of unit attachment to the polymers
chain on the glass transition temperature of it. If attachment is usual (head-to-tail), as
*
Clearly, such approach displays advantages before popular calculation method, based on group
contributions [214]. If a polymer possesses a group, which contribution into the glass transition
temperature is unknown, the calculation becomes impossible. In the same approach discussed,
contribution of any group is simply composed of the contributions of atoms composing it plus the
contribution of specific interactions, if these groups are polar.

110

schematically represented in Figure 42a, the intermolecular interaction may be


performed between neighboring polymer chains with formation of a physical network
of intermolecular bonds (see Figure 41). But if a part of chains displays an anomalous
attachment (for example, head-to-head), as schematically represented in Figure 42b,
the intermolecular interaction will proceed between two polar groups, located in
neighboring units in the same chain. Obviously, then these polar groups are excluded
from formation of the physical network between neighboring chains (see Figure 42b).
Table 18
Values of constants j and bj IRU DPLGH 1+&2 JURXSV
Polyamide type
bj103, 3K1
j103, 3K1
Aromatic polyamides
11.5
119
Aliphatic polyamides of all types*
20.4
212
Aliphatic aromatic polyamides containing CH2-groups**
14.0
145
Aliphatic aromatic polyamides based on iso-, terephthalic acid
18.1
188
and metha-, para-phenylene diamine, containing an even number
CH2-groups**
Aliphatic aromatic polyamides based on iso-, terephthalic acid
16.3
169
and metha-, para-phenylene diamine containing an odd number
of CH2-groups**
* When the number (n RI &+2 JURXSV SHU 1+&2 JURXS n > 5, additional constants bCH =

23.0 are introduced in amount m(n 5), where m LV WKH DPRXQW RI 1+&2 JURXSV SDUWLFLSDWLQJ LQ
the repeating unit.
** When the number (n RI &+2 JURXSV LQ WKe repeating unit n > 6, additional constants bCH =

23.0 are introduced in amount (n 6).

Figure 42. Nature of intermolecular interaction in case of molecules head-to-tail (a) and head-tohead (b) joined .

Since strong intermolecular interactions are taken into calculation in the relation
(IV.41) by parameters bj, clearly, it should be presented that
Vi
Tg =

ai Vi + (1 ) b j
i

(IV.44)

where is the part of anomalously joined chains.


According to the relation (IV.44), the glass transition temperature of polymer
with anomalous attachment of chains will be decreased. It will be displayed on a
particular example below.

111

Van-der-Waals volumes of the repeating units

Vi , values
i

Polymer

ai Vi , and
i

Table 19

b j and the glass transition temperatures Tg for a series of polymers


j

Vi ,

ai Vi 10 ,
3

CH2

CH C

b j 10 ,

Tg

Calc.

Exp.

2
81.3

K
3
319

K
4
81.5

5
203

6
200

79.0
71.8

279
478

81.5
0

219
150

225
150

51.2

239

55.4

279

258;
293

68.3

319

214

199

68.3

319

45.7

250

248

CH2

CH3

&+ &+ &&O&+


2

CH3
Si

CH3
CH2

CH
CH3
CH3

CH2

C
CH3

CH2

CH

111

C2H5

112

CH

2
60.7

3
295

4
55.4

5
253

6
258

77.8

375

55.4

244

248

O CH3

CH2

CH
O C2H5

CH2

CH

111.9

534

36.0

225

221

CH2

O C4H9
CH

48.9

199

55.4

341

355

63.7
96.4

238
422

0
166.2

268
377

256
378

113.5

502.0

166.2

338

338

147.6

661.0

146.7

287

293

&+ &&O
Cl

CH3
CH2

C
C O CH3
O

CH3
CH2

C
C O C 2H5
O

CH3
CH2

C
C O C4H9
O

112

CH2

113

CH3
CH2

2
181.8

3
820.0

4
127

5
262

6
268

41.5

309.0

195.0

365

358

166

578.4

136.4

375

353

54.0

183

55.4

422

418

231.8

1121

428

335

325

116

560

214

335

335

64.3

363

210

421

426

C
C O C6H13
O

CH2 CH
OH
C

C O

(CH2)2

CH2 CH

KN
C

(CH2)4

O
C

C NH

(CH2)6

NH

O
(CH2)5

NH

O
CH2

CH
C NH2
113

114

CH2

CH

CH3
CH2

CH2

CH

2
109.8

3
320

4
27.7

5
376

6
378

127

399

83.1

401

435

60.2

356

195.0

374

379;
439

238.8

771

162.0

431

422

451.1

1480

300.5

382

363

C OH
O
CH3
O

O C

CH3
C
O

(CH2)8

C O
O

O
C
O
C

114

115

C O

2
465.4

3
1163

4
351.3

5
573

6
593

437.3

1060

242

535

543

395.5

1066

453.3

646

653

519.9

1331

298.4

503

500

O
C

O
C

C O

O
C

NH

C HN
C

O
C

O
C
O
C
N
O

CH3
O

C
CH3

C
N
C
O

115

116

2
368.1

3
1133

4
468

5
553

6
560

262.3

692

76.8

416

420

389.7

1003

326.1

576

583

O
C O
O

O
C

O
C

116

Table 19
Experimental and calculated values of the glass transition temperature, Tg, for the broaden list of polymers, presented in the alphabetic order
( a i Vi +
i

V
,
i
Tg, K
Tg, K
No.
Polymer
References
+ b j )103
i
(exp.)
(calc)
3
j

, 3K1
1.
Cyanoethyl cellulose
132(0 gr) 483 (1 gr);
453
389 (1
326
CH2 OH
188(1 gr) 244 (0 gr)
gr); 541
CH O
(0 gr)

O CH
CH
CH CH
OH O CH2
2.

Ethyl cellulose

CH2 O CH2

CH2

C N

CH3

CH O
O CH
CH
CH CH
CH3
3.

CH2

Methyl cellulose

O CH2

204(2 gr) 654 (2 gr);


168(1 gr) 449 (1 gr)

316

312 (2
gr); 374
(1 gr)

275, 327

151(1 gr) 370 (1 gr);


170(2 gr) 496 (2 gr)

423

343 (2
gr); 408
(1 gr)

326

318

328

472

CH3

CH2 O CH3

CH O
O CH
CH
CH CH
CH3
4.

O CH3

Nylon 8,6 [Poly(iminoadipoyliminooctamethylene)]

(CH2)4

C NH
O

(CH2)8

811

117

NH C

266

2
Nylon 10 [Poly(imino-1-oxodecamethylene)]

NH C

3
184

4
580

5
315

6
317

7
480

368

1161

333, 319

317

472, 473

402

1280

322

314

477

300

929

313

323

472

201

640

315, 316,
365, 333 343

314

480, 668,
681, 716

218

699

313 - 316,
314

312

284, 475,
480, 679

539

1750

323

308

477

235

756

314

311

480

(CH2)9

O
6.

Nylon 10,10 [Poly(iminosebacoyliminodecamethylene)]

NH C

(CH2)8

C NH

O
7.

(CH2)10

Nylon 10,12 [Poly(iminodecamethyleneiminododecanedioyl)]

NH

(CH2)10

NH C

(CH2)10

O
8.

Nylon 10,6 [Poly(iminoadipoyliminodecamethylene)]

NH C

(CH2)4

O
9.

C NH

(CH2)10

Nylon 11 [Poly(imino-1-oxoundecamethylene)]

NH C

(CH2)10

O
10.

Nylon 12 [Poly(imino-1-oxododecamethylene)]

NH C

(CH2)11

O
11.

Nylon 12,18 [Poly(iminododecamethyleneiminooctadecanedioyl)]

NH

(CH2)12

NH C

(CH2)16

O
12.

Nylon 13 [Poly(imino-1-oxotridecamethylene)]

NH C
O

(CH2)12

C
O

118

1
5.

1
13.

2
Nylon 14,18 [Poly(iminotetradecamethyleneiminooctadecanedioyl)]

NH

(CH2)14

NH C

(CH2)16

(CH2)18

NH C

(CH2)16

7
477

641

2102

323

305

477, 478

64.7

170

384

380

469

198

577

316

343

470

215

634

318

339

471, 472

300

929

323

323

284, 317,
320, 325

334

1044

319

320

473, 477,
478

249

752

331

331

464, 473

Nylon 3 [Poly(imino-1-oxotrimethylene)]

NH C

6
307

O
15.

5
321

Nylon 18,18 [Poly(iminooctadecanedioyliminooctadecamethylene)]

NH

4
1866

O
14.

3
573

(CH2)2

O
16.

Nylon 4,6 [Poly(iminotetramethyleneiminoadipoyl)]

NH

(CH2)4

NH C

(CH2)4

O
17.

Nylon 5,6 [Poly(iminopentamethyleneiminoadipoyl)]

NH

(CH2)5

NH C

(CH2)4

O
18.

C
O

Nylon 6,10 [Poly(iminohexamethyleneiminocebacoyl)]

NH

(CH2)6

NH C

(CH2)8

O
19.

C
O

Nylon 6,12 [Poly(iminohexamethyleneiminododecanedioyl)]

NH

(CH2)6

NH C

(CH2)10

O
20.

Nylon 6,7 [Poly(iminohexamethyleneiminopimeloyl)]

(CH2)6

NH C
O

(CH2)5

C
O

119

NH

2
Nylon 6,8 [Poly(iminohexamethyleneiminosuberoyl)]

NH

(CH2)6

NH C

(CH2)6

O
22.

(CH2)6

NH C

(CH2)7

6
328

7
473

283

868

331

326

476

133

405

325

328

325, 473,
479, 480

249

752

318, 333

331

365, 464,
472, 473

266

811

328

328

473

150

464

323

323

480, 680

334

1044

333

320

472

368

1161

323

317

477

C
O

Nylon 7 [Poly(imino-1-oxoheptamethylene)]

NH C

5
330

O
23.

4
811

Nylon 6,9 [Poly(iminohexamethyleneiminoazelaoyl)]

NH

3
266

(CH2)6

O
24.

Nylon 7,6 [Poly(iminoadipoyliminoheptamethylene)]

NH C

(CH2)4

Nylon 7,7 [Poly(iminopimeloyliminoheptamethylene)]

NH C

(CH2)5

C NH

(CH2)7

O
26.

(CH2)7

O
25.

C NH

Nylon 8 [Poly(imino-1-oxooctamethylene)]

NH C

(CH2)7

O
27.

Nylon 8,10 [Poly(iminooctamethyleneiminodecanedioyl)]

NH

(CH2)8

NH C

(CH2)8

O
28.

C
O

Nylon 8,12 [Poly(iminooctamethyleneiminododecanedioyl)]

NH

(CH2)8

NH C
O

(CH2)10

C
O

120

1
21.

1
29.

2
Nylon 8,22 [Poly(iminooctamethyleneiminodocosanedioyl)]

NH

(CH2)8

NH C

(CH2)20

O
30.

4
1750

5
321

6
308

7
477

167

522

319

320

473, 479,
480

283

868

318

326

471

146

394

354
Vicat
softening
temperature

371

589

209

641

353

326

733

264

892

330

296

733

C
O

Nylon 9 [Poly(imino-1-oxononamethylene)]

NH C

3
539

(CH2)8

O
31.

Nylon 9,6 [Poly(iminoadipoyliminononamethylene)]

NH C

(CH2)4

O
32.

C NH

(CH2)9

Poly(1,1,1-trifluoro-isopropyl methacrylate)

CH3
CH2

C
C O CH CF3
O

33.

CH3

Poly(1,1,5-trihydrooctafluoroallyl-2-cyanoacrylate)

C N
CH2

C
CO O CH2 (CF2)3

34.

CF2H

Poly(1,1,7-trihydrododecafluoroheptyl-2-cyanoacrylate)

C N
CH2

C
CF2H

121

CO O CH2 (CF2)5

2
Poly(1,1-dichloro-2-fluoroethylene)

CCl2

3
68.9

4
205

5
~320 Low
molecular
weight

6
336

7
646

86.6

318

263273

272

445

132

314

373

420

733

76.1

409

191

186

491, 492

102

366

253

279

539

85.3

288

263

296

539-542

CH
F

36.

Poly(1,1-difluoro-2-trifluoromethoxyethylene)

CF2

CH
O
CF3

37.

Poly(1,1-dihydrotrifluoroethyl-2-cyanoacrylate)

C N
CH2

C
CO O CH2 CF3

38.

Poly(1,1-dimethylsilazane)

CH3
Si

NH

CH3
39.

Poly(1,1-dimethyltetramethylene)

CH3
C

(CH2)3

CH3
40.

Poly(1,1-dimethyltrimethylene)

CH3
C

(CH2)2

CH3

122

1
35.

1
41.

2
Poly(1,2-butadiene)

CH2

3
64.2

4
240

5
266, 269

6
268

7
734

98.3

426

262

231

498

306

494

653

620

296

418

931

453

449

297

165

579

258 Brittle
point

285

582

CH
CH
CH2

42.

Poly(1,2-dimethyl-1-butenylene)

CH3
C

C CH2

CH2

CH3
43.

Poly(1,3,4-oxadiazolediyl-1,4-phenylene-3,3-phthalidylidene-1,4-phenylene)

N N
C

C
O

C
O
CO

45.

Poly(1,3,4-oxadiazolediyl-1,4-phenyleneoxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4phenylene)

N N
C

CH3

O
46.

CH3

Poly(1,3-dimethylbutyl acrylate)

CH2

CH3

C O CH CH2

CH CH3

CH3

123

CH

2
Poly(1,3-dioxa-2-propyl-4,6-cyclohexylenemethylene)

CH2

CH2

HC
O

48.

CH

CF2

CH2

CH2

CF2

O CF2

CH2

O C

O
(CH2)4

O CF2

CH2

O C

(CF2)3

O
50.

51.

CF2

7
273, 324

324

1232

287

263

328

387

1363

303

284

328

230

816

301

282

328

64.2

375

166, 171

171

735, 736

98.3

489

197

201

507, 507

CH3

O CF2

Poly(1,4-butadiene)

CH2

C
O

Poly(1,4,7-trioxy-3,3,5,5-tetrafluoroheptamethyleneisophthaloyl)

O CH2

6
289

CH3

Poly(1,4,7-trioxy-3,3,5,5-tetrafluoroheptamethylenecarbonyl-1,3-phenylene-hexafluorotrimethylene1,3-phenylenecarbonyl)

O CH2

5
322

O
49.

4
498

CH

Poly(1,4,7-trioxy-3,3,5,5-tetrafluoroheptamethylene-5-pentyloxyisophthaloyl)

O CH2

3
144

O C

CH2CH=CHCH2
52.

Poly(1-ethyl-1-butenylene)

CH CH2

CH2

CH3

CH2

124

1
47.

1
53.

2
Poly(17-fluoroalkyl methacrylate)

3
321

4
1239

5
310

6
259

7
686

86.0

368

241

234

662

132

667

192

198

504

235

1205

220

195

273

119

454

~250

262

525

162

623

251

260

589

CH3
CH2

C
C O CH2

(CF2)7

CF3

O
54.

Poly(1-bromo-1-butenylene)

CH2

C CH CH2
Br

55.

Poly(1-butyl-1-butenylene)

CH2

C CH CH2
(CH2)3

56.

CH2

CH CH2

(CH2)9
57.

CH3

Poly(1-decyl-1-butenylene)

CH3

Poly(1-ethyl-1-methyltetramethylene)

CH3
C

(CH2)3

CH2
58.

CH3

Poly(1H,1H,3H-hexafluorobutyl acrylate)

CH2

CH
C O CH2

CHF

CF3
125

CF2

2
Poly(1H,1H,5H-octafluoropentyl acrylate)

CH2

3
189

4
724

5
238

6
261

7
589

206

696

309

296

603

261

946

286

276

603

316

1197

258

264

603

277

1145

256
Crystalline

242

589

CH
C O

CH2

(CF2)3

CF2H

O
60.

Poly(1H,1H,5H-octafluoropentyl methacrylate)

CH3
CH2

C
C O CH2

(CF2)3

CF2H

O
61.

Poly(1H,1H,7H-dodecafluoroheptyl methacrylate)

CH3
CH2

C
C O CH2

(CF2)5

CF2H

O
62.

Poly(1H,1H,9H-hexadecafluorononyl methacrylate)

CH3
CH2

C
C O CH2

(CF2)7

CF2H

O
63.

Poly(1H,1H,-pentadecafluorooctyl acrylate)

CH2

CH
C O CH2
O

(CF2)6

CF3

126

1
59.

1
64.

2
Poly(1H,1H-heptafluorobutyl acrylate)

CH2

3
167

4
645

5
243

6
259

7
315, 589,
590

184

523

~330

352
(syndio)

602

139

519

247

268

589

119

538

226

221

509

184

939

190

196

286, 505

115

502

221

229

504, 505,
507

CH
C O CH2

(CF2)2

CF3

O
65.

Poly(1H,1H-heptafluorobutyl methacrylate) syndiotactic

CH3
CH2

C
C O CH2

(CF2)2

CF3

O
66.

Poly(1H,1H-pentafluoropropyl acrylate)

CH2

CH
C O CH2

CF2

CF3

O
67.

Poly(1-heptene)

CH2

CH
(CH2)4
CH3

68.

Poly(1-heptyl-1-butenylene)

CH CH2

(CH2)6
69.

CH2

CH3

Poly(1-isopropyl-1-butenylene)

C CH CH2
CH
CH3

127

H3C

CH2

2
Poly(1-methyloctamethylene)

CH

3
154

4
664

5
215

6
232

7
538

81.3

457

159, 183

178

502, 508

140

509

283 low
molecular
wieght

275

499

115

578

196

199

504

132

528

293

250

286, 396,
505, 506

146

537

282

272

624

(CH2)7

CH3
71.

Poly(1-pentenylene)

CH=CH(CH2)3
72.

Poly(1-phenyl-1-butenylene)

73.

CH CH2

CH2

Poly(1-propyl-1-butenylene)

CH CH2

CH2

CH2
CH2
74.

CH3

Poly(1-tert-butyl-1-butenylene)

H3C

CH CH2

CH3

CH2

CH3
75.

Poly(2,2-dimethylbutoxyethylene)

CH2

CH
O

CH3

CH2 C CH2
CH3

CH3

128

1
70.

1
76.

2
Poly(2,3,3,3-tetrafluoropropylene)

CH2

3
71.9

4
240

5
315

6
300

7
291

135

371

378

364

291

160

366

435
softening
point

437

287

137

338

406

405

548, 557

CF
CF3

77.

Poly(2,3,4,5,6-pentafluorostyrene)

CH2

CH

F
F

78.

Poly(2,4,6-trimethylstyrene)

CH2

CH

H3C

CH3

CH3
79.

Poly(2,4-dichlorostyrene)

CH2

CH
Cl

129

Cl

2
Poly(2,4-diisopropylstyrene)

CH2

3
211

4
548

5
~435

6
385

7
287

143

340

385

420

300, 548,
555, 563,
574

137

338

379

405

280, 548,
558

120

307

374

391

287

CH
CH3
CH
CH3

H3C
81.

CH CH3

Poly(2,4-dimethylstyrene)

CH2

CH
CH3

CH3
82.

Poly(2,5-dichlorostyrene)

CH2

CH
Cl

Cl
83.

Poly(2,5-fluorostyrene)

CH2

CH
F

130

1
80.

1
84.

2
Poly(2,5-diisopropylstyrene)

CH2

3
211

4
548

5
441

6
385

7
287

143

340

416

420

548

137

338

440

405

287, 494,
559

188

431

440

437

701

CH
CH3
CH
CH3

H3C
CH
H3C
85.

Poly(2,5-dimethylstyrene)

CH2

CH
CH3

H3C
86.

Poly(2,6-dichlorostyrene)

CH2

CH

Cl

87.

Cl

Poly(2,6-xylenyl methacrylate)

CH3
CH2

CH3

C
C O
O
131

CH3

2
Poly(2-bromoethyl methacrylate)

3
135

4
380

5
325

6
355

7
560

206

624.6

339

330

551

204

711

340

287

545

135

349

450

388

556

94.5

372

256

254

500

CH3
CH2

C
C O (CH2)2

Br

O
89.

Poly(2-butoxycarbonylstyrene)

CH2

CH
C O (CH2)3

CH3

O
90.

Poly(2-butoxymethylstyrene)

CH2

CH
CH2 O (CH2)3 CH3

91.

Poly(2-carboxystyrene)

CH2

CH
C OH
O

92.

Poly(2-chloro-1,4,4-trifluoro-1-butenylene)

F
C

C CH2
Cl

CF2

132

1
88.

1
93.

2
Poly(2-chlorobenzoyloxyethylene)

CH2

CH

3
152

4
425

5
335

6
355

7
278, 628,
653

128

374

>365

342

361

124

302

392

410

548

159

562

313

283

514

226

518

418

436

706

Cl

O
O C
94.

Poly(2-chloroethyl methacrylate)

CH3
C

CH2

C O (CH2)2

Cl

O
95.

Poly(2-chlorostyrene)

CH2

CH
Cl

96.

Poly(2-cyclohexylethylethylene)

CH2

CH
CH2

97.

CH2

CH2
CH
CH2

Poly(2-decanehydronaphthyl methacrylate)
CH3
CH2

CH2
CH2
CH2

CH2 CH2
C O CH CH CH2
O

133

CH2 CH CH2
CH2 CH2

2
Poly(2-ethoxycarbonylphenyl acrylate)

CH2

3
200

4
617

5
303

6
325

7
560

172

444.6

391

387

551

170

531

347 low
viscosity

320

545

126

406

345

310

273

234

659

340, 328338

355

723

CH
C

O
O
99.

C 2H 5

Poly(2-ethoxycarbonylstyrene)
CH2 CH

C O CH2

CH3

O
100.

Poly(2-ethoxymethylstyrene)

CH2

CH
CH2

101.

O CH2

CH3

Poly(2-ethyl-1,3-dioxa-4,6-cyclohexylenemethylene)

CH2
CH2

CH CH
O

O
CH
CH2

102.

CH3

Poly(2-ethyl-2-methylpropylene terephthalate)

CH3
CH2

C CH2

H3C CH2

O C
O

C O
O

134

1
98.

1
103.

2
Poly(2-ethylhexyl methacrylate)

3
216

4
797

5
263 Brittle
point

6
271

7
581

143

404

376

354

555

131

337

384

389

562

240

801.6

318

299

551

121

346

358

350

690

CH3
CH2

104.

C
C O CH2

CH

(CH2)3

CH2

CH3

Poly(2-ethylstyrene)

CH2

CH
CH2

105.

CH3

CH3

Poly(2-fluoro-5-methylstyrene)
CH2 CH
F
H3C

106.

Poly(2-hexyloxycarbonylstyrene)

CH2

CH
C O

(CH2)5

CH3

O
107.

Poly(2-hydroxyethyl methacrylate)

CH3
CH2

C O CH2

CH2 OH

135

2
Poly(2-hydroxymethylstyrene)

CH2

4
327

5
433

6
410

7
273

138

433 (80%
isotactic), 369
(atactic), 357
(58% syndio)

328, 359,
311 (80%

560, 604606, 663

isotactic),

319(80%
isotactic),
374

393 (58%

(atactic),

CH
CH2

109.

3
134

OH

Poly(2-hydroxypropyl methacrylate)

CH3
CH2

110.

C O CH2

CH

OH

syndiotactic)

CH3
206

548.6

400

376

551

223

637.6

341

350

551

CH
C O CH2

CH CH3

CH3

Poly(2-isopentyloxycarbonylstyrene)

CH2

387
(58%
syndiotact.)

Poly(2-isobutoxycarbonylstyrene)

CH2

111.

CH

O C O CH2

CH2 CH
CH3

CH3

136

1
108.

1
112.

2
Poly(2-isopentyloxymethylstyrene)

CH2

3
221

4
725

5
351 (low
viscosity)

6
305

7
545

204

636

~330

321

553

189

469.6

419

402

551

187

555

361 (low
viscosity)

337

545

CH
CH2 O

(CH2)2

CH CH3
CH3

113.

Poly(4-isopentyloxystyrene)

CH2

CH

CH3
O
114.

(CH2)2

CH CH3

Poly(2-isopropoxycarbonylstyrene)

CH2

CH
CH3
C O CH
CH3
O

115.

Poly(2-isopropoxymethylstyrene)

CH2

CH
CH2

137

CH3
O CH
CH3

2
Poly(2-isopropyl-1,3-dioxa-4,6-cyclohexylenemethylene)

CH2

4
424

5
329

6
340

7
273

182

534

319

341

560

155

421

403 (low
viscosity)

368

556

123

496

223

248

587

153

451

362 (low
viscosity)

339

545

CH2
CH CH
O

CH3
117.

3
144

O
CH
CH

CH3

Poly(2-methoxycarbonylphenyl acrylate)

O CH3
C O
CH2

CH

O C O
118.

Poly(2-methoxycarbonylstyrene)

CH2

CH
C O CH3
O

119.

Poly(2-methoxyethyl acrylate)

CH2

CH
C O CH2

CH2

O CH3

O
120.

Poly(2-methoxymethylstyrene)

CH2

CH
CH2

O CH3

138

1
116.

1
121.

2
Poly(2-methoxystyrene)

CH2

3
136

4
363

5
~348

6
375

7
568

109

318

355

343

273

148

556

241 Brittle
point

266

582, 706

165

645

235 Brittle
point

256

582

126

314

293, 371375, 453,


443, 455

401

670, 671,
689, 697,
709, 726

CH
O CH3

122.

Poly(2-methyl-1,3-dioxa-4,6-cyclohexylenemethylene)
CH2
CH2 CH CH
O

O
CH
CH3

123.

Poly(2-methylbutyl acrylate)

CH2

124.

C O CH2

CH CH2

CH3

CH3

Poly(2-methylpentyl acrylate)

CH2

125.

CH

CH
C O CH2

CH CH2

CH3

Poly(-methylstyrene)

CH3
CH2

CH2

CH3

139

2
Poly(2-methylstyrene)

CH2

3
126

4
315

5
409

6
400

7
282, 548,
555, 563,
568

183

487

358

376

585

199

833

228 Brittle
point

239

582

255

977

286

261

575

223

712

365

313

551

CH
CH3

127.

Poly(2-naphthyl acrylate)

CH2

CH
C O
O

128.

Poly(2-octyl acrylate)

CH2

CH

CH3

C O

CH

(CH2)5

CH3

O
129.

Poly(2-octyloxystyrene)

CH2

CH
O

130.

(CH2)7

CH3

Poly(2-pentyloxycarbonylstyrene)

CH2

CH
C O
O

(CH2)4

CH3

140

1
126.

1
131.

2
Poly(2-pentyloxymethylstyrene)

CH2

(CH2)4

6
277

7
545

246

730

336 low
viscosity

337

545

213

462

397 (low
viscosity)

461

556

189

533.6

381

354

551

187

621

370 low
viscosity

301

545

CH
CH2 O

133.

5
320

CH3

Poly(2-phenethyloxymethylstyrene)

CH2

4
798

CH
CH2 O

132.

3
221

CH2

CH2

Poly(2-phenoxycarbonylstyrene)
CH2 CH
C O
O

134.

Poly(2-propoxycarbonylstyrene)

CH2

CH
C O CH2

CH2 CH3

O
135.

Poly(2-propoxymethylstyrene)

CH2

CH
CH2 O CH2

CH2 CH3
141

2
Poly(2-tert-butylaminoethyl methacrylate)
CH3
CH2

4
647

5
306

6
300

7
560, 597

206

563

345

367

585

66.7

217

300

307

291, 620,
647, 648

206

609

293

338

650

182

553

318

329

599, 706

CH3

C O (CH2)2
137.

3
194

NH C CH3

O
Poly(2-tert-butylphenyl acrylate)
CH2 CH

CH3

C O
O
H3C C CH3

138.

139.

CH3
Poly(3,3,3-trifluoropropylene)
CH2 CH

CF3
Poly(3,3-dimethyl-3-phenylpropionyloxyethylene)

CH2

CH
O
C O CH3
CH2

C
CH3

140.

Poly(3,3-dimethylbutyl methacrylate)

CH3
CH2

C
C O CH2
O

CH3
CH2

CH3

CH3

142

1
136.

1
141.

2
Poly(3,3-phthalidylidene-1,4-phenyleneoxy-5-tert-butylisophthaloyloxy-1,4-phenylene)

O C

H3C

CO

4
870

5
552

6
525

7
313

296

550

513, 538

538

313, 323

390

720

591, 543

542

6, 313

137

338

401

405

548, 558

C O
O

3
457

CH3

CH3
142.

Poly(3,3-phthalidylidene-1,4-phenyleneoxycarbonyloxy-1,4-phenylene)

O C O

O
CO
143.

Poly(3,3-phthalidylidene-1,4-phenyleneoxyisophthaloyloxy-1,4-phenylene)

O C
O

C O
O

CO
144.

Poly(3,4-dichlorostyrene)

CH2

CH

Cl
143

Cl

2
Poly(3,4-dimethylbenzoylethylene)

CH2

3
162

4
411

5
315

6
346

7
267

143

340

384

420

548

233

747

274

312

601

143

340

377

420

563

CH
C O

CH3
CH3
146.

Poly(3,4-dimethylstyrene)

CH2

CH

CH3
CH3
147.

Poly(3,5,5-trimethylhexyl methacrylate)

CH3
CH2

CH3

C O CH2
O
148.

Poly(3,5-dimethylstyrene)

CH2

H3C

CH

CH3

CH2

CH CH2

C CH3

CH3

CH3

144

1
145.

1
149.

2
Poly(3-bromobenzoyloxyethylene)

CH2

CH

3
159

4
436

5
331

6
365

7
278, 628

151

428

338

352

278, 628,
653

124

302

363

410

548

176

652

248

270

511

184

505

320

364

560

Br

O
O C
150.

Poly(3-chlorobenzoyloxyethylene)

CH2

CH

Cl

O
O
151.

Poly(3-chlorostyrene)

CH2

CH

Cl
152.

Poly(3-cyclohexylpropylethylene)

CH2

CH
CH2

153.

CH2

CH2

CH2
CH
CH2

Poly(3-dimethylaminophenyl acrylate)

CH2

CH2
CH2
CH2

CH
C O
O
145

H3C N CH3

2
Poly(3-ethoxycarbonylphenyl acrylate)

CH2

3
200

4
617

5
297

6
324

7
560

143

404

~303

354

555

117

311

398

376

686

129

369

355

349

686

121

451

277, 278,
223

268

685, 718,
724

CH
C O
O
O C O C2H5

155.

Poly(3-ethylstyrene)

CH2

CH

CH2 CH3
156.

Poly(3-fluoroalkyl -fluoroacrylate)

F
CH2

O C O CH2
157.

CF3

Poly(3-fluoroalkyl methacrylate)

CH3
CH2

O C O CH2
158.

CF3

Poly(3-hydroxybutyrate)

CH2

CH
O

O C

(CH2)3

OH

146

1
154.

1
159.

2
Poly(3-hydroxymethylstyrene)

CH2

(CH2)2

7
273

157

611

217

257

594

182

534

311

341

560

140

586

198

239

587

148

556

228 Brittle
point

266

582

CH CH3
O CH3

O
Poly(3-methoxycarbonylphenyl acrylate)

CH2

6
410

CH
C O

161.

5
398

OH

Poly(3-methoxybutyl acrylate)

CH2

4
327

CH

CH2
160.

3
134

CH
C O
O
C O
O CH3

162.

Poly(3-methoxypropyl acrylate)

CH2

CH
C O (CH2)3 O CH3
O

163.

Poly(3-methylbutyl acrylate)

CH2

CH
C O CH2

CH3

147

CH2 CH CH3

2
Poly(3-methylstyrene)

CH2

3
126

4
315

5
370

6
400

7
548, 555,
563, 572

157

561

289

280

596, 597

148

565

267, 257

262

560, 582

145

363.4

389

399

544

208

632

318

329

651

CH

CH3
165.

Poly(3-oxabutyl methacrylate)

CH3
CH2
O
166.

167.

C
C O (CH2)3

Poly(3-pentyl acrylate)
CH2 CH

CH3

CH3

CH2
O C O CH
CH2
Poly(4-acetylstyrene)

CH2

CH3

CH

O C CH3
168.

Poly(4-benzoylbutyryloxyethylene)

CH2

CH
O

O C

(CH2)3 C
O

148

1
164.

1
169.

2
Poly(4-benzoylstyrene)

CH2

3
204

4
528

5
371

6
386

7
544

213

539

383

395

578

175

476

317

329

627

159

436

365

365

278, 628,
652, 653

CH

O C
170.

Poly(4-biphenyl acrylate)

C H2
O
171.

CH
C

Poly(4-bromo-3-methoxybenzoylethylene)

CH2

CH
C O

O CH3
Br
172.

Poly(4-bromobenzoyloxyethylene)

CH2

CH
O

Br
149

Poly(4-bromostyrene)

CH

3
131

4
310

5
391, 414430

6
423

7
548-550, 668

174.

Br
Poly(4-butoxycarbonylphenyl acrylate)

234

909

286

257

560

254

669

401
softening
point

380

611

206

625

349

330

544

204

711

< 283

287

552

CH2

175.

177.

CH
C O

C O

(CH2)3

CH3

Poly(4-butoxycarbonylphenylmethacrylamide)
CH3

CH2

176.

CH2

C
C NH

C O

Poly(4-butoxycarbonylstyrene)
CH2 CH

O C O (CH2)3
Poly(4-butoxymethylstyrene)
CH2

(CH2)3

CH3

CH3

CH

CH2 O (CH2)3 CH3

150

1
173.

1
178.

2
Poly(4-butoxystyrene)

CH2

(CH2)3

7
552, 553

178

586

279

304

552, 553,
555

217

715

334

303

278, 628

183

284

CH3

Poly(4-butylyloxybenzoyloxyethylene)

CH2

6
301

CH

(CH2)3
180.

5
~320

CH3

Poly(4-butylstyrene)

CH2

4
621

CH

O
179.

3
187

CH
O
C

(CH2)2

CH3

O
181.

Poly(4-carboxyphenylmethacrylamide)

CH3
CH2

527
>473
(Softening (decompoint)
position)

611

C
C OH

151

C NH

2
Poly(4-carboxystyrene)

CH2

3
135

4
293

5
386

6
388

7
544

140

354

418

396

548

129

324

395

398

548

140

354

387

396

548

CH

COOH
183.

Poly(4-chloro-2-methylstyrene)

CH2

CH
CH3

Cl
184.

Poly(4-chloro-3-fluorostyrene)

CH2

CH

F
Cl
186.

Poly(4-chloro-3-methylstyrene)

CH2

CH

CH3
Cl

152

1
182.

1
187.

2
Poly(4-chlorobenzoylethylene)

CH2
O
188.

3
142

4
372

5
310, 362

6
333

7
627, 628

151

428

357

352

278, 628,
652, 653

124

302

383, 388401

410

287, 548550, 667

150

624

233-238

240

754

204

731

~263

279

650

CH
Cl

Poly(4-chlorobenzoyloxyethylene)
CH2 CH

O
O

Cl

189.

Poly(4-chlorostyrene)
CH2 CH

190.

Poly(4-cyanobutyl acrylate)

Cl

CH2

CH
C O

C N

(CH2)4

O
191.

Poly(4-cyclohexylbutyryloxyethylene)

CH2

CH
O
C O
CH2

(CH2)2

CH2
CH2
CH2

153

CH2
CH
CH2

193.

2
Poly(4-decylstyrene)
CH2 CH

(CH2)9 CH3
Poly(4-dimethylaminocarbonylstyrene)
CH2 CH

3
280

4
1120

5
208

6
250

7
554

174

437

398

398

544

314

1138

221

276

554

181

568

343

319

278, 628

CH3
O C N
CH3
194.

Poly(4-dodecylstyrene)

CH2

CH

(CH2)11
195.

CH3

Poly(4-ethoxybenzoyloxyethylene)

CH2

CH
O

O CH2

CH3

154

1
192.

1
196.

2
Poly(4-ethoxycarbonylphenyl acrylate)

CH2

197.

3
200

4
617

5
310

6
324

7
766

220

491

441
(Softening
point)

448

611

172

445

367

387

544

153

442

~359

346

553

162

476

325

305

627, 628

CH
C O

C O C2H5

Poly(4-ethoxycarbonylphenylmethacrylamide)

CH3
CH2

C
C NH

C O CH2

198.

Poly(4-ethoxycarbonylstyrene)
CH2 CH

199.

Poly(4-ethoxystyrene)

O C O CH2

CH2

CH3

CH

O CH2
200.

CH3

Poly(4-ethylbenzoylethylene)

CH2
O

CH3

CH
C

CH2

CH3
155

2
Poly(4-ethylstyrene)

CH2

4
404

5
300, 351

6
354

7
552, 554,
555

139

390

368

356

686

151

448

353

337

686

115

286

368

402

548-550,
563-565

CH

CH2
202.

3
143

CH3

Poly(4-fluoroalkyl -fluoroacrylate)

CH2

C
C O CH2

CF2

CF2H

O
203.

Poly(4-fluoroalkyl methacrylate)

CH3
CH2

C
C O CH2

CF2

CF2H

O
204.

Poly(4-fluorostyrene)

CH2

CH

156

1
201.

1
205.

2
Poly(4-hexadecylstyrene)

CH2

O
208.

(C H 2 ) 4

213

721.4

339

295

544

240

801.6

339

299

554

238

888

253

268

552

CH

(C H 2 ) 5 C H 3

Poly(4-hexyloxymethylstyrene)

C H2

7
554

CH3

Poly(4-hexyloxycarbonylstyrene)

CH2

6
325

CH

O C
207.

5
278

CH3

Poly(4-hexanoylstyrene)

C H2

4
1175

CH

(CH2)15
206.

3
382

CH

(C H 2 ) 5

C H3

157

C H2

2
Poly(4-hexylstyrene)

CH2

5
246

6
277

7
554

206

549

363

376

544

189

470

368

402

544

189

520

342

363

278, 628

CH3

Poly(4-isobutoxycarbonylstyrene)

CH2

4
765

CH

(CH2)5
210.

3
212

CH

CH3
O C O CH2
211.

CH CH3

Poly(4-isopropoxycarbonylstyrene)

CH2

CH

CH3
O C O CH CH3
212.

Poly(4-isopropylbenzoyloxyethylene)

CH2

CH
O
C
O

CH3
CH
CH3

158

1
209.

1
213.

2
Poly(4-methoxy-2-methylstyrene)

CH2

3
152

4
386

5
~358, 363

6
394

7
568

200

482

379

415

560

202

410

453
(Softening
point)

493

611

155

421

386

368

544

CH
CH3

O CH3
214.

Poly(4-methoxycarbonylphenyl methacrylate)

CH3
CH2

215.

C
C O

C O CH3

Poly(4-methoxycarbonylphenylmethacrylamide)

CH3
CH2

216.

C
C NH

C O CH3

Poly(4-methoxycarbonylstyrene)

CH2

CH

159

O C O CH3

2
Poly(4-methoxymethylstyrene)
CH2 CH

CH2

3
153

4
451

5
350

6
339

7
552

136

363

386

375

553, 568570, 672

126

315

366, 374,
382

400

548, 555,
569, 572,
573, 708

206

509

400

405

544

O CH3

218.

Poly(4-methoxystyrene)
CH2 CH

219.

Poly(4-methylstyrene)
CH2 CH

220.

Poly(4-morpholinocabonylstyrene)
CH2 CH

O CH3

CH3

CH2
O

C N
CH2

CH2
O
CH2

160

1
217.

1
221.

2
Poly(4-nonadecyl styrene)
CH2 CH

(CH2)18
222.

263

1031

220

255

554

417

1198

305

348

554

247

843

323

293

544

CH3

Poly(4-octanoylstyrene)

CH2

7
273

CH

(CH2)17
224.

6
360

CH3

Poly(4-octadecylstyrene)

CH2

5
305

CH

(CH2)8
223.

4
1206

CH3

Poly(4-nonylstyrene)

CH2

3
434

CH

(CH2)6

CH3

161

O C

226.

2
Poly(4-octylstyrene)
CH2 CH

(CH2)7 CH3
Poly(4-p-anisoylstyrene)
CH2 CH

3
246

4
943

5
228

6
261

7
554

211

521

376

405

544

194

515

~373
(softening
point)

377

287

185

423

434

437

571, 576

O CH3
227.

Poly(4-phenoxystyrene)
CH2 CH

O
228.

Poly(4-phenyl styrene)

CH2

CH

162

1
225.

1
229.

2
Poly(4-phenylacetylstyrene)

CH2

Poly(4-piperidinocarbonylstyrene)
CH2 CH

CH2
CH2

232.

6
393

7
544

214

534

387

401

544

170

453

375

375

544

189

534

365

354

544

CH2
CH2
CH2

Poly(4-propionylstyrene)
C H2 C H

O C C H2 C H3
Poly(4-propoxycarbonylstyrene)
C H2 C H

CH2

CH2

C H3

163

5
351

CH2

CO N
231.

4
561

CH

O C
230.

3
221

234.

2
Poly(4-propoxymethylstyrene)
C H2 C H

C H2 O
Poly(4-propoxystyrene)

C H2

C H2

C H2

Poly(4-p-toluoylstyrene)

CH2

5
295

6
301

7
552

170

531

343

320

553

220

524

372

420

544

178

511

359

348

555

CH3

Poly(4-sec-butylstyrene)

C H2

CH

C H3

CH

CH3

CH

H3C
236.

C H2

4
621

CH

O
235.

C H2

3
187

C H2

C H3

164

1
233.

1
237.

2
Poly(4-tert-butylbenzoylethylene)
C H2 C H

H 3C
238.

C H3

C H3
Poly(4-tert-butylbenzoyloxyethylene)
CH2 CH

3
196

4
515

5
377

6
344

7
628

206

572

374

360

278, 628,
655

238

607

356, 403451

392

758, 760

206

563

344

366

585

C O

H3C

C CH3
CH3

239.

Poly(4-tert-butylcyclohexyl methacrylate)
C H3

C H2

C
C

240.

C H2
CH
C H2

O
Poly(4-tert-butylphenyl acrylate)

C H2

CH
C

C H3
C

C H3

C H3

C H3
O

C
C H3

CH3
165

C H2
CH
C H2

2
Poly(4-tert-butylstyrene)
CH2 CH

CH3

3
178

4
446

5
399404

6
399

7
666

348

1156

237

301

554

196

631

343

311

544

127

236

429

494

714

C H3

C H3

242.

Poly(4-tetradecylstyrene)

CH2

CH

(CH2)13
243.

Poly(4-valerylstyrene)

CH2

O
244.

CH3

CH

C (CH2)3

Poly(4-vinyl phenol)
CH2 CH

OH

CH3

166

1
241.

1
245.

2
Poly(5,7-dihydro-1,3,5,7-tetraoxobenzo[1,2-c:4,5-c']dipyrrole-2,6[1H,3H]-diyl-1,4-phenyleneoxy1,4phenylene)

N
246.

CO

CO

CO

CO

CO

CO

CO

CO

CO

CO

CO

CO

CO

CO

CO

CO

Poly(5-bromo-2-butoxystyrene)
CH2

5
523, 773,
653

6
649

7
298, 300-309

319

519

618

615

303

286

709

408

403

299

303

734

408

413

299

303

844

383

359

299

208

667

320

312

547

(CH2)2 CH (CH2)4
CH3

CH3

(CH2)3

(CH2)3

CH3

(CH2)9

CH
O

(CH2)3

CH3

167

Br

4
476

Poly(5,7-dihydro-1,3,5,7-tetraoxobenzo[1,2-c:4,5-c']dipyrrole-2,6[1H,3H]-diylnonamethylene)

N
250.

CO

Poly(5,7-dihydro-1,3,5,7-tetraoxobenzo[1,2-c:4,5-c']dipyrrole-2,6[1H,3H]-diyl-4,4dimethylheptamethylene)

N
249.

CO

Poly(5,7-dihydro-1,3,5,7-tetraoxobenzo[1,2-c:4,5-c']dipyrrole-2,6[1H,3H]-diyl-3methylheptamethylene)

N
248.

CO

Poly(5,7-dihydro-1,3,5,7-tetraoxobenzo[1,2-c:4,5-c']dipyrrole-2,6[1H,3H]-diyl-1,4-phenylenethio-1,4phenylene)

N
247.

CO

3
309

2
Poly(5-bromo-2-ethoxystyrene)

CH2

3
174

4
487

5
353

6
357

7
547

225

680

310 (low
viscosity)

331

547

191

511

308

374

547

157

408

359

385

547

225

755

322 (low
viscosity)

298

547

CH
O CH2

CH3

Br
252.

Poly(5-bromo-2-isopentyloxystyrene)

CH2

CH

O (CH2)2

CH3

Br
253.

CH CH3

Poly(5-bromo-2-isopropoxystyrene)

CH2

CH
CH3
O CH
CH3

Br
254.

Poly(5-bromo-2-methoxystyrene)

CH2

CH
O CH3

Br
255.

Poly(5-bromo-2-pentyloxystyrene)

CH2

CH
O

Br

(CH2)4

CH3

168

1
251.

1
256.

2
Poly(5-bromo-2-propoxystyrene)
CH2 CH

257.

5
327 (low
viscosity)

6
331

7
547

177

644

250

275

580

186

677

223

274

755

156

492

350

317

686

194

469

360

414

548

CH

CH2 CH2 C N
C O CH
O CH2 CH3
O

Poly(5-cyano-3-thiapentyl acrylate)

CH2

4
577

CH3

Br
Poly(5-cyano-3-oxapentyl acrylate)

CH2

258.

(CH2)2

3
191

CH
C O

(CH2)3

(CH2)2

C N

O
259.

Poly(5-fluoroalkyl methacrylate)

CH3
CH2

C
C O CH2

CF2

CF3

O
260.

Poly(5-tert-butyl-2-methylstyrene)
CH2 CH

H3C

CH3

H3C C
169

CH3

2
Poly(6-cyano-3-thiahexyl acrylate)

CH2

3
203

4
769

5
215

6
264

7
755

203

769

215

264

755

204

823

233

248

587

237

944

214

251

755

206

696

320

296

686

CH
C O

(CH2)3

(CH2)3

C N

O
262.

Poly(6-cyano-4-thiahexyl acrylate)

CH2

CH
C O

(CH2)4

(CH2)2

C N

O
263.

Poly(7,7,8,8-tetrafluoro-3,6-dioxaoctyl acrylate)

CH2

CH
C O (CH2)2

(CH2)2

O CF2

CF2H

O
264.

Poly(8-cyano-7-thiaoctyl acrylate)

CH2

CH
C O

(CH2)7

CH2

C N

O
265.

Poly(8-fluoroalkyl methacrylate)

CH3
CH2

C
C O CH2
O

(CF2)3

CF2H

170

1
261.

1
266.

2
Poly(9-iodononylethylene)

CH2

5
267

6
260

7
649

125

303

513

413

287, 291,
295, 566

143

210

487-618;
628-649

682

287, 494,
751-753, 759

129

335

395

385

728

173

551

330

314

729

190

576

304

330

732

CH2I

Poly(,,-trifluorostyrene)

CF2

4
858

CH
(CH2)8

267.

3
223

CF

268.

Poly(acenaphthylene)

269.

Poly(allyl-2-cyanoacrylate)

CH

CH

C N
CH2

C
CH CH2

CO O CH2
270.

Poly(allyloxyethyl-2-cyanoacrylate)

C N
CH2

C
CO O (CH2)2

271.

O CH2

CH CH2

Poly(allyloxyisopropyl-2-cyanoacrylate)

C N
CH2

C
CO O CH2 CH O CH2

171

CH3

CH CH2

2
Poly(aniline)

3
87.7

4
247

5
373

6
355

7
761

128

390

314, 347

329

627, 628

138

391

344

353

278, 398,
568, 580,
628, 652,
653

127

381

333

333

511, 514,
661

112

498

218, 217

225

588, 600,
620-624, 709

145

472

330 Vicat
softening
point

307

599

NH
273.

Poly(benzoylethylene)

CH2

CH

O C
274.

Poly(benzoyloxyethylene)

CH2

CH
O

O C
275.

Poly(benzylethylene)

CH2

CH
CH2

276.

Poly(butoxyethylene)

CH2

CH
O

277.

(CH2)3

CH3

Poly(butyl chloroacrylate)

Cl
CH2

C
C O
O

(CH2)3

CH3

172

1
272.

1
278.

2
Poly(butyl cyanoacrylate)

3
150

4
457

5
358

6
328

7
613

245

706

321

347

725

193

785

223

246

699

200

643

291

311

699

200

602

304

332

699

102

449

223

227

509-513

C N
CH2

C
C O

(CH2)3

CH3

O
279.

Poly(butylene 2,6-naphthalate)

C O
C

(CH2)4

O
280.

Poly(butylene adipate)

(CH2)4

C O

O
281.

282.

283.

(CH2)4

Poly(butylene isophthalate)

C O

(CH2)4

Poly(butylene terephthalate)

C O

Poly(butylethylene)

CH2

(CH2)4

CH
(CH2)3
173

CH3

2
Poly(butylimino-2,2,3,3,4,4-hexafluoropentamethylenebutyliminoadipoyl)

285.

CH2

(CF2)3

CH2 N

(CH2)3

(CH2)3 O

CH3

CH3
C

(CF2)3

(CH2)3 O

C N

6
254

7
481

383

1508

~293
Brittle
point

252

481

121

476

253

254

600

116

358

~303

324

750

206

500

405

412

282

594

480

475

(CH2)6

CH3
CH
S

(CH2)3

CH3

Poly(butyryliminoethylene)

CH2

5
283288

Poly(butylthioethylene)

CH2

4
1508

O (CH2)3

CH3

287.

(CH2)4

Poly(butyliminohexafluoroglutarylbutyliminohexamethylene)

286.

3
383

CH
NH C

(CH2)2

CH3

O
288.

289.

Poly(carbonyl-1,3-phenylenecarbonyloxy-1,3-phenyleneoxy)

C O

Poly(carbonyl-1,3-phenylenecarbonyloxy-1,4-phenylene-1,4-phenyleneoxy)

C O

174

1
284.

1
290.

2
Poly(carbonyl-1,4-phenylenecarbonyloxy-1,4-phenylene-(methyl)phenylmethylene-1,4-phenyleneoxy)

3
391

4
788

5
483-493

6
496

7
6

282

552

540

511

494

834

598

592

554

1053

563

526

CH3

291.

292.

Poly(carbonyl-1,4-phenylenecarbonyloxy-1,4-phenylene-1,4-phenyleneoxy)

C O

Poly(carbonyl-1,4-phenylenesulfonyl-1,4-phenylenecarbonyloxy-1,4-phenylene-3,3-phthalidylidene1,4-phenylene)

SO2

C O
O

O
C
O
CO

293.

Poly(carbonyl-1,4-phenylenesulfonyl-1,4-phenylenecarbonyloxy-1,4-phenylenediphenylmethylene1,4-phenyleneoxy)

C
O

SO 2

C O

O
175

2
Poly(carbonyl-1,4-phenylenesulfonyl-1,4-phenylenecarbonyloxy-1,4-phenyleneisopropylidene-1,4phenyleneoxy)

3
437

4
837

5
518

6
522

7
6

64.5

210

~325, 373

307

238

773

331

308

283, 286,
325, 360,
397, 399,
413, 415,
476, 523,
608, 629-645
706

272

932

329

292

706

153

464

349

330

650

CH3

SO2

O
295.

C O

CH3

Poly(chlorotrifluoroethylene)

CF

CF2

Cl
296.

Poly(cyclodecyl methacrylate)

CH3
CH2

297.

CH2
C O CH
CH2
O

CH2

CH2

CH2

CH2

CH2
CH2
CH2

Poly(cyclododecyl methacrylate)

CH3
CH2

298.

CH2
C O CH
CH2
O

CH2

CH2

CH2

CH2

CH2

CH2

Poly(cyclohexanoyloxyethylene)

CH2

CH
O

CH2
C CH
CH2
O

CH2
CH2
CH2

CH2
CH2
CH2

176

1
294.

1
299.

2
Poly(cyclohexyl acrylate)

CH2

300.

CH

CH2
C O CH
CH2
O

3
153

4
532

5
292

6
288

7
583

167

413

387 Vicat
softening
point

404

599

170

456

384

373

742

170

552

298

308

650

124

326

393, 406

380

514-520

CH2
CH2
CH2

Poly(cyclohexyl chloroacrylate)

Cl
CH2

CH2
C O CH

CH2

CH2

CH2

O
301.

CH2

Poly(cyclohexyl methacrylate)

CH3
CH2

302.

CH2
C O CH
CH2
O

CH2
CH2
CH2

Poly(cyclohexylacetoxyethylene)

CH2

CH
O
C O
CH2

303.

Poly(cyclohexylethylene)
CH2 CH
H2C

CH2
CH
CH2

CH2
CH2
CH2

CH
CH2
177

H2C
CH2
H2C

2
Poly(cyclohexyloxyethylene)

CH2

CH
CH2

306.

5
354

6
306

7
624

204

614

346

332

706

221

695

326

318

706

107

301

348

355

514

250

1050

203

238

277, 581,
596, 756

CH2
CH2
CH2

Poly(cyclooctyl methacrylate)

CH3

CH2

4
438

CH
O CH2

305.

3
134

CH2
C O CH
CH2
O

CH2

CH2
CH2
CH2

CH2

Poly(cyclooctylmethyl methacrylate)

CH3
CH2

C
C O CH2
O

307.

CH2
CH2

Poly(cyclopentylethylene)

CH2

308.

CH2
CH
CH2

CH
CH2
CH
CH2

CH2
CH2

Poly(decyl methacrylate)

CH3
CH2

CH2
CH2
CH2

(C H 2 ) 9

C H3

178

1
304.

1
309.

2
Poly(decyl-2-cyanoacrylate)
C N

CH2
310.

4
996

5
192

6
254

7
731

214

1034

~183, 211

207

600, 620,
622, 623

194

649

289-297

299

597

106

757

134; 138140

140

740, 741

160

489

292

327

560

80.7

504

173, 186

160

493, 711

CO O
Poly(decyloxyethylene)

CH2

(CH2)9 CH3

CH
O

311.

3
253

(CH2)9

CH3

Poly(diethylaminoethyl methacrylate)

CH3
CH2

C
C O (CH2)2

CH3

CH2

CH3

O
312.

CH2

Poly(diethylsiloxane)

C2H5
Si

C2H5
313.

Poly(dimethylaminoethyl methacrylate)

CH3
CH2

C
C O CH2
O

314.

Poly(dimethylsilylenemethylene)

CH3
CH2

N
CH3

CH3
Si

179

CH3

CH2

2
Poly(dimethylsilylenetrimethylene)

3
115

4
665

5
203

6
173

7
490

263

809

311

325

489

208

967

208

215

461

71.3

330

246

216

351, 352

140

649

199

216

351

147

465

296

316

352, 461

123

572

201

215

351

CH3
Si

(CH2)3

CH3
316.

Poly(di-p-tolylsilylenetrimethylene)

CH3

Si

(CH2)3

CH3
317.

Poly(dithiodecamethylene)

SS(CH2)10
318.

Poly(dithioethylene)

SSCH2CH2
319.

Poly(dithiohexamethylene)

SS(CH2)6
320.

Poly(dithiomethylene-1,4-phenylenemethylene)

S
321.

CH2

Poly(dithiopentamethylene)

SS(CH2)5

CH2

180

1
315.

1
322.

2
Poly(dodecafluorobutoxyethylene)

CF2

CF

3
174

4
713

5
263273

6
244

7
445

267

1094

270 Brittle
point

244

581, 586

284

1229

208
conflicting
data

231

273, 581,
598, 600,
601

262

630

420, 425,
418, 430,
405, 431

416

684, 700,
702, 709,
715

604

1250

483

483

738, 739

478

988

515

484

738, 739

O
CF2 CF2
323.

CF2

CF3

Poly(dodecyl acrylate)

CH2

CH
C O

(CH2)11

CH3

O
324.

Poly(dodecyl methacrylate)

CH3

CH2

C
C O (CH2)11

CH3

O
325.

Poly(ehter ether ketone)

C
O

326.

Poly(ether imide) 1
O
N

CH3
O

C
C

CH3

327.

Poly(ether imide) 10
O

O
O

C
C

O
181

2
Poly(ether imide) 11
O
N

329.

500

465

738, 739

646

1462

487

442

738, 739

572

1246

485

459

738, 739

488

1030

499

474

738, 739

O
N

Poly(ether imide) 15

1028

478

7
738, 739

Poly(ether imide) 14
O

6
467

Poly(ether imide) 13
O

5
488

Poly(ether imide) 12

332.

C
C

331.

330.

4
1203

3
562

182

1
328.

1
333.

2
Poly(ether imide) 16

O
N

334.

335.

O
S

SO2

484

738, 739

656

1505

451

436

738, 739

582

1143

533

509

738, 739

520

1074

488

484

738, 739

O
N

CH3
O

C
CH3

O
O

C
C
O

N
183

530

1165

564

O
Poly(ether imide) 2

7
738, 739

Poly(ether imide) 19
O

6
455

5
482

Poly(ether imide) 18
N

337.

4
1073

336.

Poly(ether imide) 17
O

3
488

2
Poly(ether imide) 20

O
N

339.

SO2

340.

341.

342.

O
SO 2

7
738, 739

394

812

528

485

738, 739

666

1402

402

475

738, 739

572

1184

483

483

738, 739

487

1008

512

483

738, 739

O
Poly(ether imide) 24
O
N

Poly(ether imide) 23
O

6
514

Poly(ether imide) 22
N

5
538

4
969

Poly(ether imide) 21

3
498

O
N

O
C

184

1
338.

1
343.

2
Poly(ether imide) 25
O

344.

CO

CO

347.

CO

CO

CO

401

379

738, 739

572

1246

507

459

738, 739

488

1073

504

455

738, 739

656

1505

475

436

738, 739

O
N

CO

CO

CO

CO

S
O

1330

CO

Poly(ether imide) 29

504

(CH2)6

Poly(ether imide) 28

7
738, 739

O
CO

6
455

Poly(ether imide) 27

N
346.

CO

5
467

Poly(ether imide) 26

N
345.

4
1442

3
656

O
CO

CO

CO

185

CO

2
Poly(ether imide) 3
O
N

349.

CO

CO

CO

CO

CO

CO

1143

540

509

738, 739

498

969

539

514

738, 739

666

1402

503

475

738, 739

487

1008

521

483

738, 739

CO

CO

CO

CO

CO

Poly(ether imide) 33

582

SO2
O

7
738, 739

SO 2

CO

Poly(ether imide) 32

6
451

Poly(ether imide) 31

5
472

SO2

CO

352.

4
1246

Poly(ether imide) 30

351.

350.

3
562

CO

CO

CO

CO

186

1
348.

1
353.

2
Poly(ether imide) 34

N
354.

CO

CO

CO

CO

Poly(ether imide) 35

N
355.

CO

CO

CO

N
356.

562

1286

466

437

738, 739

553

1122

550

493

738, 739

CO

CO

469

862

548

544

738, 739

638

1381

497

462

738, 739

O
N

CO

CO

CO

CO

Poly(ether imide) 38

7
738, 739

O
CO

6
461

Poly(ether imide) 37

357.

CO

5
514

Poly(ether imide) 36

4
855

CO

3
394

CO

CO

CO

CO

187

2
Poly(ether imide) 39

O
N
359.

CO

CO

CO

CO

478

1028

482

465

738, 739

478

1028

505

465

738, 739

646

1462

471

442

738, 739

597

1164

549

513

737

O
CO

CO

CO

Poly(ether imide) 41

7
738, 739

CO

362.

361.

6
467

Poly(ether imide) 40

5
512

360.

4
1203

Poly(ether imide) 4

3
562

CO

CO

CO

CO

Poly(ether imide) 42

CF3
C
CF3

CO

O
CO

CO
O

CO

188

1
358.

1
363.

2
Poly(ether imide) 43

CF3
C

CF3

364.

CO

CO

CO

CO

CF3

CO

CO

366.

Poly(ether imide) 46

CF3

367.

CO

CO

CO

CO

CO

CO

Poly(ether imide) 47

1164

491

513

737

597

1164

529

513

737

478

988

536

484

738, 739

394

812

532

485

738, 739

CO

CO

597

O
N

7
737

Poly(ether imide) 45

6
513

CO

CF3

5
529

CO
365.

4
1164

Poly(ether imide) 44

CF3

3
597

O
N

CO

CO

CO

CO

N
189

2
Poly(ether imide) 48

O
N
369.

N
370.

371.

CO

CO

CO
O

5
487

6
451

7
738, 739

478

1047

499

465

738, 739

394

855

497

461

738, 739

683

1444

503

473

737

598

1189

554

503

737

CO

CO

CO

CO

O
N

Poly(ether imide) 5
O
N

4
1246

CO

Poly(ether imide) 49

3
562

C
C

O
Poly(ether imide) 50
O

O
N

CO
CO
CO
CO

372.

Poly(ether imide) 51
H3C

CH3
CH2

H3C

N
CH3

O
CO
CO

O
CO
CO

190

1
368.

1
373.

2
Poly(ether imide) 52

CH3

H3C
CH2

H3C
374.

Poly(ether imide) 53
H3C

5
538

6
503

7
737

598

1189

545

503

737

581

1107

581

525

737

471

975

523

483

737

CO

CO

CO

CO

CH3
N

H3C

O
CO

CO

CO

CO

CH3

Poly(ether imide) 54

H3C

CH3
N

H3C
376.

4
1189

CH3

CH2

375.

3
598

CO

CO

CO

CO

CH3

Poly(ether imide) 55

O
N

CO
CO

CO

CO
O

191

CF3

2
Poly(ether imide) 56

O
N

CO

CO

CO

CO

3
471

4
975

5
523

6
483

7
737

471

975

508

483

737

692

1401

526

494

737

692

1401

519

494

737

562

1286

403

437

738, 739

CF3

378.

Poly(ether imide) 57

CO

CO

CO

CO

CF3

379.

Poly(ether imide) 58
CH3

CH3

CH3

CH3

O
CO

CO

CO

CO

380.

Poly(ether imide) 59
CH3

CH3

CH3

CH3

O
N

CO
CO
CO
CO

381.

Poly(ether imide) 6
O

O
O

O
N

192

1
377.

1
382.

383.

2
Poly(ether imide) 60
CH3

CH3

CH3

CH3

CO

CO

CO

CO

CO

CO

CO

O
O

O
CO

CO

CO

Poly(ether imide) 64

O
O

Poly(ether imide) 65
N

CO

6
494

7
737

608

1310

502

464

737

608

1310

501

464

737

608

1351

481

450

737

683

1444

500

473

737

439

916

533

479

737

CO

CO

CO

CO

5
519

CO

Poly(ether imide) 63

4
1401

CO

387.

CO

CO

Poly(ether imide) 62

386.

385.

Poly(ether imide) 61
O

384.

O
CO

3
692

CO
CO

CO
O

193

CO

2
Poly(ether imide) 66

O
N

389.

CO

CO

CO

CO

Poly(ether imide) 67

4
916

5
528

6
479

7
737

439

916

503

479

737

439

916

527

479

737

523

1092

513

479

737

553

1122

502

493

738, 739

O
N

3
439

CO
CO
CO
CO

O
390.

Poly(ether imide) 68

O
N

391.

CO

CO

CO

CO

Poly(ether imide) 69
O

CO
CO

CO

CO
O
392.

Poly(ether imide) 7
O
N

O
O

194

1
388.

1
393.

2
Poly(ether imide) 70

O
N

394.

Poly(ether imide) 71
O

CO

CO

CO

CO

3
523

4
1092

5
508

6
479

7
737

523

1092

522

479

737

523

1092

518

479

737

523

1134

499

461

737

523

1134

498

461

737

O
CO

CO
CO
CO

O
395.

Poly(ether imide) 72

O
N

396.

Poly(ether imide) 73
O

CO

CO

CO

CO
O

CO
CO

CO

CO

O
397.

Poly(ether imide) 74
O

CO

CO
CO
CO

195

2
Poly(ether imide) 75

O
N

400.

Poly(ether imide) 76

CO

CO

CO

CO

CH3

4
1134

5
494

6
461

7
737

565

1137

538

497

737

565

1137

509

497

737

565

1137

520

497

737

656

1442

489

455

738, 739

CH3

3
523

CO
CO

CO

CO

O
401.

Poly(ether imide) 77

CH3
C

CH3

CO
CO
CO
CO

O
402.

Poly(ether imide) 78

CH3
C

CH3
403.

Poly(ether imide) 79

CO

CO

CO

CO

CO
O

CO

CO

CO

CO

O
N

196

1
399.

1
404.

2
Poly(ether imide) 8

O
N

C
C

CO

CO

CO

CO

483

738, 739

520

1075

509

484

738, 739

612

1246

503

491

738, 739

577

1378

408

397

738, 739

CH3

CO
CO

CO

CH2
N

CH3
O

C
CH3

O
CO
CO

(CH2)6

197

CO

508

CH3

CO

CO

1251

CH3

Poly(ether imide) 83

604

CH3

408.

CO
CO

Poly(ether imide) 82

7
738, 739

CO
O

407.

C
CH3

CO

Poly(ether imide) 81

6
496

O
O

406.

5
520

CH3

Poly(ether imide) 80

4
946

O
405.

3
469

2
Poly(ether imide) 84
O

O
N

410.

CO

CO

CO

CO

CO

C
C

Poly(ethyl chloroacrylate)

5
500

6
473

7
738, 739

683

1444

504

473

766

638

1381

478

462

738, 739

111

294

366 Vicat
softening
point

377

588, 599,
614, 615

131

427

300

307

560

176

553

325

318

613

Cl
CH2

4
1444

3
683
N

Poly(ether imide) 9
O

412.

CO

CO

O
N

411.

CO

Poly(ether imide) 85
O

C
C O CH2

CH3

413.

Poly(ethyl ethacrylate)

CH2
CH2

CH3

C
C O CH2

CH3

414.

Poly(ethyl ethoxycarbonyl methacrylate)


CH2

CH2

C O CH2

C O CH2
O

CH3

CH3

198

1
409.

1
415.

2
Poly(ethyl-2-cyanoacrylate)

3
116

4
279

5
422

6
416

7
730

85.3

297

268

287

526

211

548

398

385

725

211

548

398

385

699, 727

315

1150

293

274

481

315

1150

~278

274

481

C N
CH2

C
CO O CH2 CH3

416.

Poly(ethyl-2-propylene)
CH2 CH3

CH2
417.

CH3
Poly(ethylene 2,6-naphthalate)

(CH2)2

O C
O

C
O

418.

Poly(ethylene 2,6-naphthalenedicarboxylate)

(CH2)2

O C
C O

O
419.

Poly(ethylimino-2,2,3,3,4,4-hexafluoropentamethyleneethyliminoadipoyl)

420.

CH2

(CF2)3

CH2 N

CH2

CH2

CH3

CH3

(CH2)4

Poly(ethyliminohexafluoroglutarylethyliminohexamethylene)

CH2 O

C N

(CH2)6

O CH2
CH3

199

CH3

(CF2)3

2
Poly(ethylthioethylene)

CH2

3
86.5

4
297

5
266

6
291

7
601

62.3

207

310 (~60%
syndio)

301

501

128

391

347, 336

327

687, 705

91.8

340

268

270

445

312

1268

269

246

612

137

634

208 - 228

216

496, 509,
512

CH
S
CH3

CH2
422.

Poly(formyloxyethylene)

CH2

CH
O

O C
423.

Poly(glycidyl methacrylate)

CH3
CH2

C
C O CH2

CH
O

O
424.

Poly(hexafluoromethoxyethylene)

CF2

CF
O

425.

CF3

Poly(hexyl hexyloxycarbonyl methacrylate)

CH2

CH2

C O

O C O (CH2)5
426.

CH2

Poly(hexylethylene)

CH2

CH
(CH2)5
CH3

(CH2)5

CH3

CH3

200

1
421.

1
427.

2
Poly(hexyloxyethylene)

CH2

CH

3
146

4
676

5
199

6
216

7
600, 620624, 709

212

436

553, 513

486

378, 474,
486

273

713

383

383

485

212

394

< 500

538

378

212

394

< 500

538

378

212

353

618, 580 600

601

378, 379

258

459

578

562

379

O
(CH2)5

428.

NH

429.

CH3

Poly(imino-1,3-phenyleneiminoisophthaloyl)

NH C

Poly(imino-1,3-phenyleneiminosebacoyl)

NH

NH C

(CH2)8

O
430.

NH C

Poly(imino-1,4-phenyleneiminoterephthaloyl)

NH
433.

NH C

Poly(imino-1,4-phenyleneiminoisophthaloyl)

NH
432.

Poly(imino-1,3-phenyleneiminoterephthaloyl)

NH
431.

NH C

Poly(imino-1,5-naphthyleneiminoterephthaloyl)
NH
C

201

NH C

1
434.

435.

2
Poly(imino-1-methyl-3-oxotrimethylene)

NH CH CH2

CH3

Poly(imino-1-oxoethylene-1,4-phenylene-2-oxoethyleneiminooctadecamethylene)
NH C CH2

436.

C NH

CH2

CH2

CH2

CH2

C NH

CH2

CH2

CH2

CH2

O
438.

7
469

478

1354

351

353

478

392

1107

358

354

478

512

1467

338

349

478

543

1278

438

425

285

266

826

338

322

406

300

992

336

302

483

(CH2)18

Poly(imino-2,2,4-trimethylpentamethyleneiminoadipoyl)

CH3

CH3

C CH2

CH CH2

NH C

CH3
440.

C NH

Poly(imino-1-oxotrimethylenefluoren-9-ylidene-3-oxotrimethyleneiminotrimethylenefluoren-9ylidenetrimethylene)
NH C (CH2)2 C (CH2)2 C NH (CH2)3 C (CH2)3

NH CH2

6
407

439.

5
369

(CH2)11

O
O
Poly(imino-1-oxotrimethylene-1,4-phenylene-3-oxotrimethyleneiminooctadecamethylene)

NH C

4
201

(CH2)18

O
O
Poly(imino-1-oxotrimethylene-1,4-phenylene-3-oxotrimethyleneiminododecamethylene)
NH C

437.

CH2

3
81.8

Poly(imino-2,2-dimethylpentamethyleneiminoazelaoyl)

(CH2)4

C
O

CH3
C
CH3

(CH2)3

NH C
O

(CH2)7

C
O

202

NH CH2

1
441.

2
Poly(imino-2,2-dimethylpentamethyleneiminodipoyl)

3
249

4
801

5
350

6
311

7
483

266

878

344

303

483

256

531

430

481

375

255

627

393

407

334

313

749

465

418

444

CH3
NH CH2

(CH2)3

NH C

CH3
442.

(CH2)4

C
O

Poly(imino-2,2-dimethylpentamethyleneiminopimeloyl)

CH3
NH CH2

(CH2)3

NH C

CH3
443.

(CH2)5

C
O

Poly(imino-2,2-dimethylpentamethyleneiminoterephthaloyl)

CH3
NH CH2

(CH2)3

NH C

CH3
444.

Poly(imino-5-methylisophthaloyliminohexamethylene)

NH C

C NH

(CH2)6

O
CH3

445.

Poly(imino-5-tert-butylisophthaloyliminomethylene-1,3-phenylenemethylene)

NH C

C NH

CH2

CH2

H3C C CH3
203

CH3

2
Poly(imino-5-tert-butylisophthaloyliminomethylene-1,4-phenylenemethylene)

NH C

C NH

O
H3C

3
313

4
708

5
477

6
442

7
444

272

696

343

391

373

388

1057

393

367

285, 317

261

906

278

288

464

256

656

377

390

482

CH2

CH2

O
C CH3
CH3

447.

Poly(iminoadipoyliminomethylene-2,5-dimethyl-1,4-phenylenemethylene)

H3C
NH

(CH2)4

NH CH2

CH2

O
H3C

448.

Poly(iminoadipoyliminotrimethylenefluoren-9-ylidenetrimethylene)

NH

(CH2)4

449.

(CH2)3

(CH2)3

Poly(iminoadipoyliminotrimethylenemethyliminotrimethylene)

NH C
O
450.

C NH

(CH2)4

C NH

(CH2)3

(CH2)3

CH3

Poly(iminocarbonyl-1,4-phenylene-2-oxoethyleneiminohexamethylene)

NH C
O

CH2

C NH
O

(CH2)6

204

1
446.

1
451.

2
Poly(iminoethylene-1,4-phenyleneethyleneimino-1,11-dioxoundecamethylene)

NH (CH2)2

(CH2)2

NH C

(CH2)9

O
452.

(CH2)2

NH C

(CH2)12

(CH2)2

NH C

(CH2)14

(CH2)2

NH C

(CH2)16

(CH2)2

NH C

363

478

443

1241

358

357

478

478

1354

348

353

478

341

900

378

379

478

232

721

355

322

483

(CH2)8

O
456.

366

Poly(iminoethylene-1,4-phenyleneethyleneiminosebacoyl)

NH (CH2)2

1127

O
455.

409

Poly(iminoethylene-1,4-phenyleneethyleneimino-1,18-dioxooctadecamethylene)

NH (CH2)2

7
478

O
454.

6
360

Poly(iminoethylene-1,4-phenyleneethyleneimino-1,16-dioxohexadecamethylene)

NH (CH2)2

5
369

O
453.

4
994

Poly(iminoethylene-1,4-phenyleneethyleneimino-1,14-dioxotetradecamethylene)

NH (CH2)2

3
358

Poly(iminoglutarylimino-2,2-dimethylpentamethylene)

C
O

CH3
NH C

C
O

NH

CH2

C
CH3

(CH2)3
205

(CH2)3

2
Poly(iminohexamethyleneimino-1-oxotrimethylenefluoren-9-ylidene-3-oxotrimethylene)

NH

(CH2)6

NH C

(CH2)2

(CH2)2

458.

5
395

6
367

7
285, 317

249

752

323

331

476

382

1000

427-437

382

318

304

608

<500

500

378

256

572

426

447

375

273

652

398

419

444

NH C

(CH2)5

O
459.

4
1057

Poly(iminohexamethyleneimino-4-methylpimeloyl)

NH (CH2)6

3
388

Poly(iminohexamethyleneiminocarbonyl-1,4-phenylene-2,2-butylidene-1,4-phenylenecarbonyl)

CH3
NH

(CH2)6 NH C
O

CH2

CH3
460.

Poly(iminoisophthaloylimino-1,4-phenylenemethylene-1,4-phenylene)

NH C

C NH

O
461.

CH2

Poly(iminoisophthaloylimino-2,2-dimethylpentamethylene)

CH3
NH C

C NH

O
462.

CH2

C (CH2)3
CH3

Poly(iminoisophthaloylimino-3,4-dimethylhexamethylene)

NH C
O

C NH
O

(CH2)2

CH3
CH CH
CH3

(CH2)2

206

1
457.

1
463.

2
Poly(iminoisophthaloylimino-4,4'-biphenylylene)

NH C

Poly(iminoisophthaloyliminohexamethylene)

C NH

C NH

604

390

396

285, 317,
357, 484

246

599

438

411

285, 317,
444

273

717

388

381

485

395

828

448

477

285, 317

329

715

473

460

444

CH2

CH2

Poly(iminoisophthaloyliminooctamethylene)

C NH

(CH2)8

Poly(iminoisophthaloyliminotrimethylenefluoren-9-ylidenetrimethylene)

NH C

C NH

(CH2)3

(CH2)3

Poly(iminomethylene-1,3-cyclohexylenemethyleneimino-5-tert-butylisophthaloyl)

NH CH2

239

NH C

468.

7
378, 379

(CH2)6

Poly(iminoisophthaloyliminomethylene-1,3-phenylenemethylene)

467.

6
544

NH C
466.

5
558

NH C

465.

4
529

C NH

O
464.

3
288

CH
CH2

CH2
CH2

CH CH2
CH2

NH C

H3C C

CH3
207

CH3

2
Poly(iminomethylene-1,3-phenylenemethyleneimino-1-oxotrimethylenefluoren-9-ylidene-3oxotrimethylene)

NH CH2

CH2

NH C

(CH2)2

(CH2)2

O
470.

CH2

NH C

(CH2)4

CH2

CH2

NH C

(CH2)10

CH2

NH C

(CH2)16

CH2

285

321

757

432

424

285

341

897

378

380

463, 478

443

1234

348

359

478

392

1104

363

355

478

Poly(iminomethylene-1,4-phenylenemethyleneiminopentadecanedioyl)

NH CH2

347

O
474.

346

Poly(iminomethylene-1,4-phenylenemethyleneiminooctadecanedioyl)

NH CH2

689

O
473.

239

Poly(iminomethylene-1,4-phenylenemethyleneiminododecanedioyl)

NH CH2

7
285, 317

NH C
O

472.

6
432

Poly(iminomethylene-1,3-phenylenemethyleneiminocarbonyl-2,2'-biphenylenecarbonyl)

NH CH2

5
423

O
471.

4
914

Poly(iminomethylene-1,3-phenylenemethyleneiminoadipoyl)

NH CH2

3
395

NH C
O

(CH2)13

C
O

208

1
469.

1
475.

2
Poly(iminomethylene-1,4-phenylenemethyleneiminotridecanedioyl)

NH CH2

CH2

NH C

(CH2)11

CH2

NH C

(CH2)9

CH2

NH C

(CH2)4

O
H3C

7
478

324

878

380

369

463, 478

306

841

382

364

444

313

749

461

418

444

307

783

383

392

478

Poly(iminomethylene-5-tert-butyl-1,3-phenylenemethyleneiminoadipoyl)

NH CH2

6
361

O
477.

5
373

Poly(iminomethylene-1,4-phenylenemethyleneiminoundecanedioyl)

NH CH2

4
992

O
476.

3
358

C
O

CH3

CH3
478.

Poly(iminomethylene-5-tert-butyl-1,3-phenylenemethyleneiminoisophthaloyl)

CH2

NH CH2

H3C

NH C

C CH3
CH3

479.

Poly(iminooctamethyleneimino-1-oxoethylene-1,4-phenylene-2-oxoethylene)

NH

(CH2)8

NH C

CH2

C
O

209

CH2

2
Poly(iminooxalylimino-2,2-dimethylpentamethylene)

3
181

4
482

5
382

6
376

7
483

239

648

376

369

482

307

751

423

409

488

304

566

< 500

537

378

307

751

418

409

361, 406

307

751

421, 432

409

406

CH3
NH C

C NH CH2

O O
481.

CH2

CH2

CH3

Poly(iminopentamethyleneiminocarbonyl-1,4-phenylene-2-oxoethylene)

NH

(CH2)5

NH C

CH2

O
482.

CH2

C
O

Poly(iminoterephthaloylimino-1,4,4-trimethylheptamethylene)

CH3

CH3
NH C

C NH CH

O
483.

CH3

C NH

CH2

Poly(iminoterephthaloylimino-2,2,4-trimethylheptamethylene)

CH3
NH C
O
485.

(CH2)3

Poly(iminoterephthaloylimino-1,4-phenylenemethylene-1,4-phenylene)

NH C
484.

(CH2)2

C NH CH2 C
O

CH3
CH2

C NH CH2 CH CH2
O

(CH2)3

CH3

Poly(iminoterephthaloylimino-2,4,4-trimethylheptamethylene)

NH C

CH

CH3

CH3
C
CH3

(CH2)3

210

1
480.

1
486.

2
Poly(iminoterephthaloylimino-3-ethylhexamethylene)

NH C
O
487.

C NH

(CH2)2

C NH

(CH2)2

CH2

CH3

CH

5
403

6
383

7
406

290

727

416

399

406

288

488

613

590

378, 379

290

690

425

420

406

341

900

393

379

484, 487

239

561

413

426

406

(CH2)3

Poly(iminoterephthaloylimino-4,4'-biphenylylene)

NH C

4
713

CH
CH3

H3C

489.

(CH2)3

Poly(iminoterephthaloylimino-3-isopropylhexamethylene)

NH C

488.

CH

3
273

C NH
O

Poly(iminoterephthaloylimino-4,4-dimethylheptamethylene)

CH3
NH C
O
490.

(CH2)3

C NH
O

C NH
O

(CH2)6
211

(CH2)3

(CH2)12

Poly(iminoterephthaloyliminohexamethylene)

NH C

C
CH3

Poly(iminoterephthaloyliminododecamethylene)

NH C
491.

C NH

2
Poly(iminoterephthaloyliminomethylene-2,5-dimethyl-1,4-phenylenemethylene)

3
279

4
564

5
498

6
495

7
373

290

769

388

377

406

348

841

446-455

414

318

222

568

357

391

482

456

1373

358

332

285, 317

H3C
NH C

C NH

CH2

CH2

O
H3C

493.

Poly(iminoterephthaloyliminononamethylene)

NH C

C NH

O
494.

(CH2)9

Poly(iminotetramethyleneiminocarbonyl-1,4-phenylene-2,2-butylidene-1,4-phenylenecarbonyl)

CH3
NH

(CH2)4 NH C
O

CH2

CH3
495.

Poly(iminotetramethyleneiminocarbonyl-1,4-phenylene-2-oxoethylene)

NH

496.

(CH2)4 NH C

CH2 C

Poly(iminotrimethylenefluoren-9-ylidenetrimethyleneiminosebacoyl)

NH

(CH2)3

(CH2)3

NH C
O

(CH2)8

C
O

212

1
492.

1
497.

2
Poly(isobutoxyethylene)

CH2

CH
O

498.

CH2

CH
C O CH2
O

499.

4
424

5
254

6
264

7
273, 333,
495, 588,
600, 620,
622-626

131

466

249 Brittle
point, 230

281

582, 706

145

398

363

364

588, 617

CH3
CH
CH3

Poly(isobutyl acrylate)

CH2

3
112

CH3
CH
CH3

Poly(isobutyl chloroacrylate)

Cl
CH2

CH3

C O CH2
O
500.

Poly(isobutyl methacrylate)

CH
CH3
148

CH3
CH2

CH3

C O CH2 CH
O
501.

Poly(isobutylethylene)

CH2

CH

440
326, 337
336
(random), (random), random),
530 (iso)
281
279 (iso)
(isotactic)

506, 581,
583, 596,
607, 706,
710

CH3
102

375

302, 297 333

272

281, 524,
527-537

CH2
CH CH3
213

CH3

2
Poly(isohexylethylene)

CH2

CH2 CH2

5
239

6
246

7
511

135

364

372

371

278, 370,
658

119

465

259

256

511

94.9

344

270, 261

276

588, 600,
620, 624,
709

114

388

267-270

294

506, 582,
583

CH3
CH
CH3

Poly(isonicotinoyloxyethylene)

CH2

4
557

CH
CH2

503.

3
137

CH
O
C
O

504.

Poly(isopentylethylene)

CH2

CH
CH2

CH3
CH
CH3

CH2
505.

Poly(isopropoxyethylene)

CH2

CH
O CH CH3
CH3

506.

Poly(isopropyl acrylate)

CH2
O

CH
C O

CH3
CH
CH3

214

1
502.

1
507.

3
128

Poly(isopropyl chloroacrylate)

4
318

Cl
CH2

CH3

C
C O
O

508.

Poly(isopropyl methacrylate)

131

CH3
C

CH3
C O CH
CH3
O

Poly(isopropylethylene)

CH2

6
402

7
588, 599

CH
CH3
CH2

509.

5
363 Vicat
softening
point

306 (iso),
428 (iso),
300
(isotactic),
314
417
(syndio)
(syndio)
358
(syndiotactic)

706

85.3

287

323

297

514, 521,
528, 534

349

1195

303-308

292

481

349

1195

~293

292

481

60.7

240

242, 246,
228, 245,
248, 251

253

273, 333,
577, 600,
691, 694,
696, 704,
709, 719

CH
CH CH3
CH3

510.

Poly(isopropylimino-2,2,3,3,4,4-hexafluoropentamethyleneisopropyliminoadipoyl)

N CH2

(CF2)3

CH CH3

CH2

H3C CH

CH3
511.

(CH2)4

C
O

CH3

Poly(isopropyliminohexafluoroglutarylisopropyliminohexamethylene)

(CF2)3

H3C CH O

(CH2)6

CH3

Poly(methoxyethylene)

CH2

C N

O CH CH3

CH3
512.

CH
CH3

215

2
Poly(methyl chloroacrylate)

3
94.2

4
216

5
413 Vicat
softening
point

6
437

7
599

84.5

198

404

426

579, 618

112

289

357

387

579, 616

Cl
CH2

C
C O CH3
O

514.

Poly(methyl fluoroacrylate)

F
CH2

C
C O CH3
O

515.

Poly(methyl fluoromethacrylate)

CF3
CH2

C
C O CH3
O

516.

Poly(methyl methacrylate)

96.4

CH3
CH2

C
C O CH3

301 (iso), 311, 323 320 (iso),


224
(isotactic),
431
(syndio)
(syndio)
378
(syndiotactic)

766

O
517.

Poly(methyl phenylacrylate)

155

CH2
O

C
C O CH3

364

391, 397

426

619

216

1
513.

1
517.

2
Poly(methylimino-2,2,3,3,4,4-hexafluoropentamethylenemethyliminoadipoyl)

CH2 (CF2)3

CH2 N

CH3
518.

CH3

(CH2)4

Poly(methyliminohexafluoroglutarylmethyliminohexamethylene)

(CF2)3

CH3 O
519.

C N

4
989

5
298-303
(Brittle
point)

6
284

7
481

281

989

298-303

284

481

138

496

301

278

707

133

437

319

304

757

186

407

500, 357,
423, 481

457

273, 362,
495, 595,
669, 698

270

1154

198-320

234

580, 757

C
O

(CH2)6

O CH3

Poly(methylphenylsilylenemethylene)

Si

3
281

CH2

CH3
520.

Poly(N-butylacrylamide)

CH2

CH
C NH

(CH2)3

CH3

O
521.

Poly(N-carbazolylethylene)
CH2 CH

522.

Poly(N-dodecylacrylamide)

CH2

CH
O

(CH2)11

CH3

217

C NH

2
Poly(neopentyl methacrylate)

3
165

4
465

5
299-312

6
355

7
664, 706

119

399

332
softening
point,
crystalline
sample

298

511

182

630

213 Brittle
point

225

580, 706

135

364

360

371

278, 658

216

986

215 Britte
point, 184

219

580, 706

CH3
CH2

524.

CH3

C O CH2

CH3

CH3

Poly(neopentylethylene)

CH2

CH
CH2

H3C

CH3

CH3
525.

Poly(heptyl acrylate)

CH2

CH

O C O C7H15
526.

Poly(nicotinoyloxyethylene)

CH2

CH
O
C
O

527.

Poly(nonyl acrylate)

CH2

CH
C O
O

(CH2)8

CH3

218

1
523.

1
528.

2
Poly(nonylethylene)

CH2

3
188

4
900

5
236

6
209

7
512

148

630

216

235

706

133

373

390

357

580

133

307

401

433

580

150

331

433

453

579

123

436

270

282

291

CH
(CH2)8
CH3

529.

Poly(n-pentyl acrylate)

CH2

CH
CH3

O C O (CH2)4
530.

Poly(N-sec-butylacrylamide)

CH2

531.

C NH

CH

CH3

CH2

Poly(N-tert-butylacrylamide)

CH2

532.

CH

CH

CH3

C NH

CH3

CH3

CH3

Poly(N-tert-butylmethacrylamide)

CH3
CH2

533.

CH3
C NH C CH3
CH3
O

Poly(octafluoro-4-methyl-1-butenylene)

CF

CF

CF
CF3

219

CF2

535.

2
Poly(octamethyene p,p'-dibenzoate)

C O

Poly(octyl acrylate)

CH2

3
344

4
1055

5
315

6
326

7
722

199

896

208 Brittle
point

222

580, 592,
706

180

853

194

211

600, 620-623

112

394

308

284

333-336

77.8

401

204

194

333

77.8

290

277

268

333

(CH2)8 O

CH
C O

(CH2)7

CH3

O
536.

Poly(octyloxyethylene)

CH2

CH
O
(CH2)7

537.

CH3

Poly(oxy-tert-butylethylene)

O CH2
CH3

CH
C

CH3

CH3
538.

Poly(oxy-1,1-dimethylethylene)

CH3
O C

CH2

CH3
539.

Poly(oxy-1,2-dimethylethylene)

CH3
O CH
CH3

CH

220

1
534.

1
540.

2
Poly(oxy-1,3-phenylenecarbonyloxycarbonyl-1,3-phenyleneoxytetramethylene)

C O C
O

541.

(CF2)3

O C

(CF2)3

543.

544.

O C

521

1598

345

326

314

206

500

411

412

380, 409,
426, 432-434

458

962

476

473

456

851

543-553

536

SO 2

Poly(oxy-1,4-phenylene-3,3-phthalidylidene-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

7
350

Poly(oxy-1,4-phenylene(methyl)phenylmethylene-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4phenylene)
CH3

6
297

Poly(oxy-1,3-phenyleneoxyisophthaloyl)

5
< 293

O
542.

4
956

(CH2)4

Poly(oxy-1,3-phenylenehexafluorotrimethylene-1,3-phenyleneoxycarbonyl-1,3phenylenehexafluorotrimethylene-1,3-phenylenecarbonyl)

3
284

SO2

CO

221

2
Poly(oxy-1,4-phenylene-9,9-anthronylidene-1,4-phenyleneoxyterephthaloyl)

O C

3
455

4
753

5
627

6
604

7
6

389

894

428

435

297

366

787

478

465

400

726

571

551

376

340

833

428

408

375

CO
546.

Poly(oxy-1,4-phenylenecarbonyl-1,4-phenyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
O

O
547.

Poly(oxy-1,4-phenylenecarbonyl-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

O
548.

CH3

CO

SO 2

Poly(oxy-1,4-phenylenecarbonylimino-1,4-phenylenesulfonyl-1,4-phenyleneiminocarbonyl-1,4phenylene)

C NH

SO2

NH C

O
549.

Poly(oxy-1,4-phenylenecarbonylimino-2,2-dimethylpentamethyleneiminocarbonyl-1,4-phenylene)

CH3
O

C NH CH2

(CH2)3

CH3

NH C
O

222

1
545.

1
550.

2
Poly(oxy-1,4-phenylenediphenylmethylene-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

551.

O C
C

552.

4
1071

5
503

6
482

7
6

499

1279

424

390

395

420

917

448

458

297

296

610

463, 554

485

308, 378,
379

SO 2

Poly(oxy-1,4-phenylenefluoren-9-ylidene-1,4-phenyleneoxysebacoyl)

3
516

(CH2)8

C
O

Poly(oxy-1,4-phenylenehexafluoro-2,2-propylidene-1,4-phenyleneoxy-1,4-phenylenecarbonyl-1,4phenylene]
CF
3

CF3
553.

Poly(oxy-1,4-phenyleneiminoisophthaloylimino-1,4-phenylene)

NH C

C NH
O

223

2
Poly(oxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4-phenyleneazo-1,4-phenylene)
CH3

O
555.

3
389

4
960

5
448

6
405

7
297

389

894

433

435

482

1090

438

442

399

854

458, 463473

467

555

1394

393

398

297

482

1090

438

442

297

N N

CH3
Poly(oxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4-phenylenecarbonyl-1,4-phenylene)

CH3
O

CO

CH3
556.

Poly(oxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4-phenylenecarbonyl-1,4phenylenecarbonyl-1,4-phenylene)

CH3

CH3
557.

Poly(oxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

CH3
O

SO2

CH3
558.

Poly(oxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4phenylenesulfonylmethyliminotetramethylenemethyliminosulfonyl-1,4-phenylene)
CH3
CH3
CH3
O

SO 2

(CH2)4

SO 2

CH3

559.

Poly(oxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4-phenyleneterephthaloyl-1,4-phenylene)

CH3
O

C
CH3

224

1
554.

1
560.

2
Poly(oxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4-phenylenethionyl-1,4-phenylene)

3
394

4
940

5
438

6
419

7
6

365

806

453

453

357

808

453

442

290

633

423

458

188

373

487

504

297, 322,
381383

366

787

478

465

297

CH3
O

CH3
561.

Poly(oxy-1,4-phenylenemethylene-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

O
562.

SO 2

SO 2

O C

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenylene)

O
565.

Poly(oxy-1,4-phenyleneoxy-1,4-phenyleneoxyterephthaloyl)

564.

CH2

Poly(oxy-1,4-phenyleneoxy-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

O
563.

SO2

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenylenecarbonyl-1,4-phenylene)

SO 2

C
225

2
Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenylenecyclohexylidene-1,4-phenylene)

SO 2

3
438

4
948

5
478

6
462

7
297

516

1081

503

482

297

399

854

449

467

292, 297,
361, 369

394

940

438

419

297

C
H2C CH2
H2C CH2
CH2

567.

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenylenediphenylmethylene-1,4-phenylene)

568.

SO 2

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
O

SO 2

C
CH3

569.

Poly(oxy-1,4-phenylenesulfinyl-1,4-phenyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
O

S
O

C
CH3

226

1
566.

1
570.

2
Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenylenemethylene-1,4-phenylene)

O
571.

SO 2

3
365

4
806

5
453

6
453

7
319

458

962

473

476

297

377

748

518

504

367

850

448

432

297

465

951

508

489

297

CH2

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenylenemethylphenylmethylene-1,4phenylene)

CH3

572.

SO2

SO 2

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenylenethio-1,4-phenylene)

O
574.

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

O
573.

SO 2

SO 2

Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxy-2,6-dimethyl-1,4-phenyleneisopropylidene-3,5dimethyl-1,4-phenylene)
CH
HC
3

CH3
O

SO 2

H3C

CH3

227

CH3

2
Poly(oxy-1,4-phenylenesulfonyl-1,4-phenyleneoxyterephthaloyl)

O
576.

SO 2

SO 2

Poly(oxy-1-methyltrimethylene)

O CH CH2

5
522

6
540

7
380

338

637

523

531

322

368

665

533

553

322

367

850

448

432

77.8

346

223

225

364

122

473

271

258

338

SO 2

Poly(oxy-1,4-phenylenethio-1,4-phenyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

4
574

SO2

SO 2

O
579.

Poly(oxy-1,4-phenylenesulfonyl-4,4'-biphenylylenesulfonyl-1,4-phenylene)

O
578.

Poly(oxy-1,4-phenylenesulfonyl-2,7-naphthylenesulfonyl-1,4-phenylene)

577.

O C

3
310

SO 2

CH2

CH3
580.

Poly(oxy-2,2,2-trichloroethylethylene)

O CH2

CH
CH2
CCl3

228

1
575.

1
581.

2
Poly(oxy-2,2,3,3,4,4,5,5-octafluorohexamethyleneoxycarbonyliminohexamethylene-iminocarbonyl)

O CH2

(CF2)4

CH2

O C NH

(CH2)6

(CF2)4

CH2

O C

(CH2)2

(CH2)2

(CH2)2

O
583.

(CF2)3

CH2

O C

(CH2)2

(CH2)2

(CH2)2

585.

(CF2)3

CH2

(CF2)3

CH2

O C

(CH2)4

(CF2)3

CH2

221

421

312

1412

~233

221

421

326

1025

314

318

357

241

1000

216

241

309-422

460

1655

301

278

314

C
O

Poly(oxy-2,2,3,3,4,4-hexafluoropentamethyleneoxycarbonyl-1,3-phenylenedecafluoropentamethylene1,3-phenylenecarbonyl)

O CH2

~235
Brittle
point

O
586.

1534

Poly(oxy-2,2,3,3,4,4-hexafluoropentamethyleneoxyadipoyl)

O CH2

339

Poly(oxy-2,2,3,3,4,4-hexafluoropentamethyleneoxy-4,4'-octafluorobiphenylylene)

O CH2

7
745, 746

O
584.

6
272

Poly(oxy-2,2,3,3,4,4-hexafluoropentamethyleneoxy3,6-dithiaoctanedioyl)

O CH2

5
271

Poly(oxy-2,2,3,3,4,4,5,5-octafluorohexamethyleneoxy-3,6-dithiaoctanedioyl)

O CH2

4
1199

NH C

O
582.

3
326

O C

C
O

229

(CF2)5

2
Poly(oxy-2,2,3,3,4,4-hexafluoropentamethyleneoxycarbonyl-1,3-phenylenehexafluorotrimethylene1,3-phenylenecarbonyl)

O CH2

(CF2)3

CH2

O C

(CF2)3

589.

(CF2)3

(CF2)4

CH2

O CH2

CH3

CH2
CH
CH2
O

C
O CH
H3C

CH3

CH2

CH3
C

O C

CH3
H3C

1035

318

313

314

272

1058

247

257

358

271

642

442

422

279

460

1281

318

359

404

CH3
O

324

CH C

Poly(oxy-2,6-dimethyl-1,4-phenyleneisopropylidene-3,5-dimethyl-1,4-phenyleneoxysebacoyl)

H3C

7
314

CH2

CH O C
C

6
288

CH2

Poly(oxy-2,2,4,4-tetramethyl-1,3-cyclobutyleneoxycarbonyl-trans-1,4-cyclohexylenecarbonyl)

H3C

591.

O C

Poly(oxy-2,2,3,3,4,4-hexafluoropentamethyleneoxymethylene-1,4-phenylenemethylene)

O CH2
590.

CH2

5
290

Poly(oxy-2,2,3,3,4,4-hexafluoropentamethyleneoxycarbonyl-3,3'-biphenylylenecarbonyl)

O CH2

4
1406

O
588.

3
405

O
CH3

(CH2)8

C
O

230

1
587.

1
592.

2
Poly(oxy-2,6-diphenyl-1,4-phenylene)

3
236

4
483

5
493, 498

6
489

7
447, 450,
546, 673

663

1842

365

360

404

190

642

322

296

329

390

1114

380

350

329

593.

Poly(oxy-2,6-diphenyl-1,4-phenylenemethylene-3,5-diphenyl-1,4-phenyleneoxysebacoyl)
O

CH2

O C

(CH2)8

594.

Poly(oxy-2-acetoxytrimethyleneoxy-1,4-phenylene)

O CH2

CH CH2

C
O

O
O C CH3
595.

Poly(oxy-2-acetoxytrimethyleneoxy-1,4-phenylene-1-ethyl-1,3-cyclohexylene-1,4-phenylene)

CH2
O CH2

CH CH2
O

CH

H2C CH2 CH2


CH3 CH2

231

O C CH3

2
Poly(oxy-2-acetoxytrimethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

3
317

4
906

5
333

6
350

7
329

294

799

403

368

329

372

944

373

394

276

375

1036

338

362

276

CH3
O CH2

CH CH2

C
CH3

O
O C CH3
597.

Poly(oxy-2-acetoxytrimethyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

O CH2

CH CH2

SO 2

O
O C
598.

CH3

Poly(oxy-2-acetoxytrimethyleneoxy-2,6-dichloro-1,4-phenyleneisopropylidene-3,5-dichloro-1,4Cl
Cl
phenylene)
CH3

O CH2

CH CH2

C
CH3

O
C CH3

Cl

Cl

O
599.

Poly(oxy-2-benzoyloxytrimethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
O CH2

CH CH2
O

O C

C
CH3

232

1
596.

1
600.

2
Poly(oxy-2-butenylene oxycarbonyliminohexamethyleneiminocarbonyl) cis-trans

O CH2

CH CH CH2

O C NH

(CH2)6

HC

CH CH2

O C

(CH2)8

O
602.

O CH2

257

1163

232

221

408

127

472

~293

269

363

152

492

333

309

359

330

940

348

351

359

352

964

413

365

359

CH3
Cl

Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenylene)

O CH2

7
408, 427

C CH2
CH2

603.

6
269

Poly(oxy-2-ethyl-2-chloromethyltrimethylene)

CH2

5
234, 228

Poly(oxy-2-bytenyleneoxysebacoyl) cis, trans

O CH2

4
918

NH C

O
601.

3
247

CH CH2 O
OH

604.

Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenylene-1,3,3-trimethyltrimethylene-1,4-phenylene)

CH3
O CH2
605.

CH CH2 O

CH CH2

OH

CH3

CH3

Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenylene-1-ethyl-1,4-cyclohexylene-1,4-phenylene)

O CH2

CH CH2 O
OH

CH2

CH2

CH2

CH2

C
H2C

CH

233

CH3

2
Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenylene-1-methyl-1,4-cyclohexyleneisopropylidene-1,4phenylene)

H3C

O CH2

CH CH2 O

CH2

607.

5
408

6
385

7
359

296

836

368

354

359

279

756

373

369

359

337

862

388

391

329

245

708

353

346

359

CH C
CH2

OH

4
1003

CH3

CH2

3
386

CH2

CH3

Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenyleneisobutylidene-1,4-phenylene)

CH3
O CH2

CH CH2 O

CH2

OH
608.

C
CH3

Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
O CH2
609.

CH CH2 O

OH

CH3

Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenylenemethyl(phenyl)methylene-1,4-phenylene)

CH3
O CH2

CH CH2 O

OH

610.

Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenylenemethylene-1,4-phenylene)

O CH2

CH CH2 O
OH

CH2

234

1
606.

1
611.

2
Poly(oxy-2-hydroxytrimethyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

O CH2

CH CH2

3
256

4
648

5
428

6
395

7
329

334

793

388

421

359

306

939

358

326

359

342

1276

290

268

424

SO2

OH
612.

Poly(oxy-2-hydroxytrimethyleneoxy-2,6-dichloro-1,4-phenyleneisopropylidene-3,5-dichloro-1,4phenylene)

Cl

Cl
CH3
O CH2

CH CH2 O

OH

CH3
Cl

613.

Cl

Poly(oxy-2-hydroxytrimethyleneoxy-2-chloro-1,4-phenyleneisopropylidene-3-chloro-1,4-phenylene)

Cl

Cl
CH3

O CH2
614.

CH CH2 O

OH

CH3

Poly(oxy-2-pentyloxyisophthaloyloxy-2,2,3,3,4,4-hexafluoropentamethylene)

(CH2)4

CH3

O
O

C O CH2

CF2

CF2

CF2

CH2
235

2
Poly(oxy-2-propionyloxytrimethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

3
334

4
997

5
331

6
335

7
329

410

1155

330

355

404

646

1587

371

407

404

311

1239

~248

251

421

CH3
O CH2

CH CH2

O
O C
616.

CH3
CH2

CH3

Poly(oxy-3,3',5,5'-tetramethyl-4,4'-biphenylylenesebacoyl)

H3C

CH3

O C

(CH2)8

O
H3C
617.

CH3

Poly(oxy-3,3',5,5'-tetraphenyl-4,4'-biphenylylenesebacoyl)

O C

(CH2)8

618.

C
O

C
O

Poly(oxy-3-heptafluoropropylglutaryloxy-2,2,3,3,4,4,5,5-octafluorohexamethylene)

O C
O

CH2 CH CH2

CF2

CF2

CF3

O CH2

(CF2)4

CH2

236

1
615.

1
619.

2
Poly(oxy-3-heptafluoropropylglutaryloxy-2,2,3,3,4,4-hexafluoropentamethylene)

O C

CH2

CH
CF2
CF2

620.

CH2

O CH2

(CF2)3

(CH2)2

4
1224

5
243 - 248

6
254

7
421

281

749

353

375

373

348

725

503

480

297

342

1276

282

268

424

274

794

359

345

409

CH2

O
CF3

Poly(oxy-3-oxotrimethyleneiminomethylene-2,5-dimethyl-1,4-phenylenemethyleneimino-1oxotrimethylene)
H3C

3
311

C NH CH2

CH2

NH C

(CH2)2

O
H3C

621.

Poly(oxy-4,4'-biphenylyleneoxy-1,4-phenylenesulfonyl-1,4-phenylene)

O
622.

SO 2

Poly(oxy-4-pentyloxyisophthaloyloxy-2,2,3,3,4,4-hexafluoropentamethylene)

C O CH2

CF2

CF2

CF2

CH2

O
(CH2)4
623.

CH3

Poly(oxy-5-butyl-1,3-phenyleneoxyisophthaloyl)

CH3

O
237

(CH2)3

O C

2
Poly(oxy-5-ethyl-1,3-phenyleneoxyisophthaloyl)

O C

3
240

4
615

5
395

6
390

7
409

308

972

335

317

409

223

526

426

424

409

364

1260

307

289

430

CH2
CH3
625.

Poly(oxy-5-hexyl-1,3-phenyleneoxyisophthaloyl)

O C

(CH2)5
CH3
626.

Poly(oxy-5-methyl-1,3-phenyleneoxyisophthaloyl)

O C

C O

CH3
627.

Poly(oxy-5-nonyl-1,3-phenyleneoxy-2-fluoroisophthaloyl)

(CH2)8
CH3

O C

238

1
624.

1
628.

2
Poly(oxy-5-nonyl-1,3-phenyleneoxy-5-fluoroisophthaloyl)

O C

(CH2)8

3
364

4
1260

5
293

6
289

7
334

359

1238

304

290

409

342

1152

314

297

409

316

972

343

325

374

CH3
629.

Poly(oxy-5-nonyl-1,3-phenyleneoxyisophthaloyl)

O C

(CH2)8
CH3
630.

Poly(oxy-5-octyl-1,3-phenyleneoxyisophthaloyl)

O C

(CH2)7
CH3
631.

Poly(oxy-5-oxopentamethyleneiminomethylene-1,4-phenylenemethyleneimino-1-oxopentamethylene)

(CH2)4

C NH CH2

NH C
O

(CH2)4
239

CH2

2
Poly(oxy-5-pentyloxyisophthaloyloxy-1,4-phenylenemethylene-1,4-phenylene)

O C

C O

(CH2)4

7
380

404

1033

443

391

380

342

1276

243

268

424

376

1011

411

372

380

SO2

O
O

(CH2)4

CH3

Poly(oxy-5-pentyloxyisophthaloyloxy-2,2,3,3,4,4-hexafluoropentamethylene)

O C

C O CH2

(CF2)3

CH2

O
O

(CH2)4

CH3

Poly(oxy-5-pentyloxyisophthaloyloxy-4,4'-biphenylene)

6
360

CH3

C O

635.

5
383

CH2

Poly(oxy-5-pentyloxyisophthaloyloxy-1,4-phenylenesulfonyl-1,4-phenylene)

O C

634.

4
1092

O
O

633.

3
393

C O

O
O

(CH2)4

CH3

240

1
632.

1
636.

2
Poly(oxy-5-propyl-1,3-phenyleneoxyisophthaloyl)

O C

3
257

4
704

5
394

6
365

7
409

428

1597

291

268

409

393

1419

295

277

409

326

865

341

377

407

CH2
CH2
637.

CH3

Poly(oxy-5-tridecyl-1,3-phenyleneoxyisophthaloyl)

O C

(CH2)12
CH3
638.

Poly(oxy-5-undecyl-1,3-phenyleneoxyisophthaloyl)

O C

(CH2)10
CH3
639.

Poly(oxyadipoyloxy-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
O C

C O

CH3

241

(CH2)4

2
Poly(oxyadipoyloxy-2,6-dimethyl-1,4-phenyleneisopropylidene-3,5-dimethyl-1,4-phenylene)

H3C

3
392

4
961

5
366

6
408

7
404

341

832

381

410

404

295

1261

217

234

290, 406

77.8

374

185

208

709

112

533

203

210

330

332

1114

293

298

314

CH3
CH3

O C

(CH2)4

C O

CH3
H3C

641.

CH3

Poly(oxyadipoyloxy-3,3',5,5'-tetramethyl-4,4-biphenylylene)

H3C
O C

(CH2)4

CH3

C O
O
H3C

642.

Poly(oxyadipoyloxydecamethylene)

O C

(CH2)4

C O

(CH2)10

O
643.

CH3

Poly(oxybutylene)

O(CH2)4
644.

Poly(oxybutylethylene)

O CH2

CH
(CH2)3

645.

CH3

Poly(oxycarbonyl-1,3-phenyleneoxy-1,3-phenylenecarbonyloxy-2,2,3,3,4,4hexafluoropentamethylene)

O C
O

C O CH2
O

(CF2)3

CH2

242

1
640.

1
646.

2
Poly(oxycarbonyl-1,4-cyclohexylenecarbonyloxy-1,4-phenyleneisopropylidene-1,4-phenylene) trans

CH2
O C
647.

CH C O
CH2

4
825

5
423

6
422

7
279

148

523

240

283

676

348

1188

287

293

411

131

498

220

263

675

521

1541

313

338

441

326

970

335

336

442

CH3

CH2

CH

3
348

CH2

C
CH3

Poly(oxycarbonyl-1,5-dimethylpentamethylene)

CH3
O C CH

(CH2)3

CH3

O
648.

CH

Poly(oxycarbonyl-2,6-naphthylenecarbonyloxydecamethylene)

O C
C O

(CH2)10

O
649.

Poly(oxycarbonyl-3-methylpentamethylene)

O C

(CH2)2

O
650.

CH

(CH2)2

CH3

Poly(oxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyloxyhexadecamethylene)

O C NH

CH2

NH C O

O
651.

Poly(oxycarbonylimino-4-methyl-1,3-phenyleneiminocarbonyloxynonamethylene)

O C NH

NH C O
CH3

(CH2)9
243

(CH2)16

2
Poly(oxycarbonylimino-4-methyl-1,3-phenyleneiminocarbonyloxyoctamethylene)

O C NH

NH C O

O
653.

CH3

Poly(oxycarbonyliminomethylene-1,4-phenylenemethyleneiminocarbonyloxydecamethylene)

CH2

NH C O

CH2

NH C O

1094

322

329

441

394

1220

318

323

441

463

1475

320

314

441

227

672

323

338

316

270

882

319

306

314

(CH2)12

Poly(oxycarbonyliminomethylene-1,4-phenylenemethyleneiminocarbonyloxy-hexadecamethylene)

CH2

NH C O

(CH2)16

Poly(oxycarbonylneopentylenesulfonylneopentylene)

CH3

CH3
CH2

CH2

SO 2

CH2

CH2

CH3

CH3

Poly(oxycarbonyloxy-1,3-phenylenehexafluorotrimethylene-1,3-phenylene)

O C O
O

360

O C NH CH2

657.

7
442

(CH2)10

Poly(oxycarbonyliminomethylene-1,4-phenylenemethyleneiminocarbonyloxydodecamethylene)

O C

6
340

O C NH CH2

656.

5
337

(CH2)8

655.

4
909

O C NH CH2
654.

3
309

(CF2)3

244

1
652.

1
658.

2
Poly(oxycarbonyloxy-1,4-phenylene-2,2-butylidene-1,4-phenylene)

3
256

4
643

5
407

6
398

7
385, 394

273

732

410

373

384

280

636

394

440

385, 388

278

650

448

428

384, 385,
388

CH3
O C O

CH2
CH3

659.

Poly(oxycarbonyloxy-1,4-phenylene-2,2-pentylidene-1,4-phenylene)

CH3
O C O

C
CH2

CH2
CH3
660.

Poly(oxycarbonyloxy-1,4-phenylenebenzylidene-1,4-phenylene)

O C O

CH

661.

Poly(oxycarbonyloxy-1,4-phenylenecyclohexylidene-1,4-phenylene)

O C O
O

C
CH2

H3C

CH2
CH2

245

H2C

663.

2
Poly(oxycarbonyloxy-1,4-phenyleneethylidene-1,4-phenylene)

O C O

CH

CH3

Poly(oxycarbonyloxy-1,4-phenylenehexafluoro-2,2-propylidene-1,4-phenylene)

3
222

4
530

5
403

6
419

7
388

270

577

449

468

387

256

643

422

398

388

365

859

393

425

388

205

506

420

405

388, 402

207

551

~383

376

402

CF3
O C O

O
664.

CF3

Poly(oxycarbonyloxy-1,4-phenyleneisobutylidene-1,4-phenylene)

CH3
O C O

C
CH2

CH3
665.

Poly(oxycarbonyloxy-1,4-phenyleneisopropylidene-1,3-phenyleneisopropylidene-1,4-phenylene)

O C O
O
666.

CH3

CH3

CH3

CH3

Poly(oxycarbonyloxy-1,4-phenylenemethylene-1,4-phenylene)

O C O

CH2

O
667.

Poly(oxycarbonyloxy-1,4-phenylenethio-1,4-phenylene)

O C O
O

246

1
662.

1
668.

2
Poly(oxycarbonyloxy-2,2,3,3,4,4,5,5-octafluorohexamethylene)

O C O CH2

CF2

CF2

CF2

CF2

3
182

4
816

5
232

6
223

7
389

294

593

453, 493,
504

496

384, 386,
387

305

723

443 - 452

422

385

299

676

427

442

385

289

682

385

424

388

CH2

O
669.

Poly(oxycarbonyloxy-2,6-dichloro-1,4-phenyleneisopropylidene-3,5-dichloro-1,4-phenylene)

Cl

Cl

CH3
O C O

C
CH3

O
Cl
670.

Cl

Poly(oxycarbonyloxy-2-chloro-1,4-phenylenecyclohexylidene-3-chloro-1,4-phenylene)
Cl

Cl

O C O
O

671.

H2C

CH2

H2C

CH2
CH2

Poly(oxycarbonyloxy-2-chloro-6-methyl-1,4-phenyleneisopropylidene-3-chloro-5-methyl-1,4phenylene)
Cl
Cl

CH3
O C O

CH3
H3C

672.

CH3

Poly(oxycarbonyloxy-2-isopropyl-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
CH3

CH
CH3

C
CH3

247

O C O

2
Poly(oxycarbonyloxy-2-methoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

H3C

4
625

5
418

6
424

7
388

311

697

408

446

385, 393

255

577

413

442

388

272

714

368, 373,
418, 363 383

381

384-386, 393

O C O

O
674.

3
265

CH3

Poly(oxycarbonyloxy-2-methyl-1,4-phenylenecyclohexylidene-3-methyl-1,4-phenylene)

H3C

CH3

O C O
O
H2C
H2C
675.

CH2

CH2
CH2

Poly(oxycarbonyloxy-2-methyl-1,4-phenyleneisopropylidene-1,4-phenylene)

H3C
CH3
O C O
O
676.

C
CH3

Poly(oxycarbonyloxy-2-methyl-1,4-phenyleneisopropylidene-3-methyl-1,4-phenylene)

H3C

CH3
CH3

O C O
O

C
CH3

248

1
673.

1
677.

2
Poly(oxycarbonyloxy-3-methyl-1,4-phenylenebenzylidene-2-methyl-1,4-phenylene)

4
685

5
455

6
457

7
385

140

633

230 low
molecular
wieght
sample

221

389

419

1167

321

359

441, 442

214

1070

232

200

330, 333

365

1393

265

262

314

H3C

CH3
O C O

3
313

CH

678.

Poly(oxycarbonyloxyhexamethylene)

O C O

(CH2)6

O
679.

Poly(oxydecamethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

(CH2)10

O C NH

CH2

O
700.

Poly(oxydecylethylene)

O CH2

CH
(CH2)9

701.

NH C

CH3

Poly(oxydimethylsilylene-1,3-phenylenehexafluorotrimethylene-1,3-phenylenedimethylsilylene)

CH3
O

Si

(CF2)3

Si
CH3

249

CH3

CH3

2
Poly(oxydimethylsilylene-1,3-phenylenetetrafluoroethylene-1,3-phenylenedimethylsilylene)

Si

(CF2)2

CH3
Si

CH3

CH3

Si

Si

Poly(oxydimethylsilyleneoxydimethylsilylene-2,4,5,6-tetrafluorophenylenedimethylsilylene)

CH3
O

CH3

228

946

245

241

446

300

1429

212

210

446

446

2084

221

214

428

CH3

CH3

Si

Si
CH3

CH3
F

Poly(oxydimethylsilyleneoxypentyleneoxyisophthaloyloxyneopentylenedimethylsilylene)

CH3

CH3

C O CH2

C CH2

Si

CH3

CH3

CH3
O

371

F
F

706.

286

CH3
F

Si

293

CH3

1021

CH3

705.

292

Si

Poly(oxydimethylsilylene-2,4,5,6-tetrafluorophenylenedimethylsilylene)

7
450

CH3

CH3
704.

6
266

CH3

Poly(oxydimethylsilylene-1,4-phenyleneoxy-1,4-phenylenedimethylsilylene)

5
271

Si

CH3
703.

4
1271

CH3

CH3
O

3
338

Si
CH3

(CH2)5

O C
O

250

1
702.

1
707.

2
Poly(oxydimethylsilyleneoxypentyleneoxyterephthaloyloxyneopentylenedimethylsilylene)

CH3

CH3

C O CH2

C CH2

Si

CH3

CH3

CH3
O

Si

(CH2)5

O C

CH3
708.

Poly(oxydiphenoxymethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)

3
446

4
2046

5
238

6
218

7
429

407

998

365

408

405

390

920

398

424

405

356

868

385

410

405

O
CH3
O C O

C
CH3

O
709.

Poly(oxydiphenoxymethyleneoxy-3,3'-dimethyl-4,4'-biphenylene)

CH3

O C O
O
710.

CH3

Poly(oxydiphenoxymethyleneoxy-4,4'-biphenylene)

O
O C O
O
251

2
Poly(oxydiphenylsilylene-1,3-phenylene)

712.

713.

Si

CH3

CH3

Si

Si

CH3

CH3

Poly(oxydodecamethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

(CH2)12

O C NH

CH2

5
~331

6
341

7
295

354

1378

~273

286

451

453

1291

316

351

441

115

542

220

212

351, 352

181

586

291

309

279

NH C

O
714.

4
765

Si

Poly(oxydiphenylsilyleneoxydimethylsilylene-1,4-phenylenedimethylsilylene)

3
261

Poly(oxyethylenedithioethylene)

O(CH2)2SS(CH2)2
715.

Poly(oxyethyleneoxycarbonyl-1,4-cyclohexylenecarbonyl) trans

O CH2

CH2

CH2
O C CH
CH2
O

CH2
CH C
CH2

252

1
711.

1
716.

2
Poly(oxyethyleneoxycarbonyl-1,4-phenylene-sec-butylidene-1,4-phenylenecarbonyl)

O CH2

CH2

O C
O

717.

CH CH2

CH2

CH2

4
840

5
380

6
368

7
318

211

548

386

385

411, 412

211

548

392

385

412

299

726

390

412

443

282

664

412, 366

425

442, 443

333

852

379

391

443

CH3

Poly(oxyethyleneoxycarbonyl-2,6-naphthylenecarbonyl)

O CH2

3
309

O C
O

C
O

718.

Poly(oxyethyleneoxycarbonyl-2,7-naphthylenecarbonyl)

O CH2

719.

CH2

O C

Poly(oxyethyleneoxycarbonylimino-1,4-phenylene ethylene-1,4-phenyleneiminocarbonyl)

(CH2)2

O C NH

(CH2)2

NH C

O
720.

Poly(oxyethyleneoxycarbonylimino-1,4-phenylene methylene-1,4-phenyleneiminocarbonyl)

(CH2)2

O C NH

CH2

O
721.

NH C
O

Poly(oxyethyleneoxycarbonylimino-1,4-phenylene tetramethylene-1,4-phenyleneiminocarbonyl)

(CH2)2

O C NH

NH C
O

253

(CH2)4

2
Poly(oxyethyleneoxyisophthaloyl)

723.

(CH2)2

O C

Poly(oxyethylenetetrathioethylene)

3
166

4
485

5
324

6
342

7
412

153

733

233

209

351

77.8

355

203

219

330

43.7

186

243

235

353, 354

251

1041

218 - 223

241

421

224

918

218 223

244

421

368

979

357

376

442

279

866

317

322

411

O(CH2)2SSSS(CH2)2
724.

Poly(oxyethylethylene)

O CH2

CH
CH2

725.

CH3

Poly(oxyethylidene)

O CH
CH3
726.

Poly(oxyglutaryloxy-2,2,3,3,4,4,5,5-octafluorohexamethylene)

O C

C O CH2

(CH2)3

Poly(oxyglutaryloxy-2,2,3,3,4,4-hexafluoropentamethylene)

O C

C O CH2

(CH2)3

(CF2)3

Poly(oxyheptamethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

(CH2)7

O C NH

CH2

O
729.

CH2

O
728.

CH2

O
727.

(CF2)4

Poly(oxyhexamethyleneoxycarbonyl-2,6-naphthylenecarbonyl)

(CH2)6

O C
O

NH C

C
O

254

1
722.

1
730.

2
Poly(oxyhexamethyleneoxycarbonyliminomethylene-1,4-phenylenemethyleneiminocarbonyl)

(CH2)6

O C NH

CH2

CH2

O
731.

4
846

5
329

6
345

7
441

146

712

206

205

330

122

298

403

410

414, 440

374

806

433, 513

464

288, 291,
425

333

722

462

461

313, 403,
404, 426

298

673

423

443

380

NH C
O

Poly(oxyhexylethylene)

O CH2

3
292

CH
(CH2)5
CH3

732.

733.

Poly(oxyisophthaloyl)

Poly(oxyisophthaloyloxy-1,4-phenylenebenzylidene-1,4-phenylene)

O C
O

734.

O C

C O

CH

Poly(oxyisophthaloyloxy-1,4-phenyleneisopropylidene-1,4-phenylene)

CH3
O C
O
735.

C O

CH3

Poly(oxyisophthaloyloxy-1,4-phenylenemethylene-1,4-phenylene)

O C

CH2
255

C O

2
Poly(oxyisophthaloyloxy-2,2,3,3,4,4-hexafluoropentamethylene)

O C
O
737.

C O CH2

(CF2)3

C O

CH3

H3C

7
410, 424

399

819

498

487

404

365

772

461

473

403

C O

366

882

438

415

401

332

723

418

459

403

CH3
CH2

O
H3C

739.

6
289

CH3

Poly(oxyisophthaloyloxy-2,6-dimethyl-1,4-phenylenemethylene-3,5-dimethyl-1,4-phenylene)

O C

5
298

H3C
738.

4
858

CH2

Poly(oxyisophthaloyloxy-2,6-dimethyl-1,4-phenyleneisopropylidene-3,5-dimethyl-1,4-phenylene)
H3C
CH3
CH3

O C

3
248

CH3

Poly(oxyisophthaloyloxy-2-methyl-1,4-phenyleneisopropylidene-3-methyl-1,4-phenylene)

H3C

CH3
CH3

O C
O
740.

C O

CH3

Poly(oxyisophthaloyloxy-2-methyl-1,4-phenylenemethylene-3-methyl-1,4-phenylene)

H3C
O C
O

C O
O

CH3
CH2

256

1
736.

1
741.

2
Poly(oxyisophthaloyloxy-4,4'-biphenylylene)

O C

4
594

5
437, 583,
480

6
475

7
309, 423

121

644

<193

188

420, 422,
447, 747

245

980

238

250

358

357

838

379

426

444

271

863

325

314

279

329

1154

279

285

358

C O

O
742.

3
282

Poly(oxymethyl-3,3,3-trifluoropropylsilylene)

CH3
O

Si
CH2

743.

CH2

O CH2

744.

CF3

Poly(oxymethylene-1,3-phenylenemethyleneoxy-2,2,3,3,4,4-hexafluoropentamethylene)

CH2

O CH2

(CF2)3

CH2

Poly(oxymethylene-1,3-phenylenemethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4phenyleneiminocarbonyl)

O CH2

CH2

O C NH

CH2

NH C

O
745.

Poly(oxymethylene-1,4-cyclohexylenemethyleneoxycarbonyl-trans-1,4-cyclohexylenecarbonyl)

O CH2
746.

CH2
CH
CH2

CH2
CH CH2
CH2

CH2
O C CH
CH2
O

CH2
CH C
CH2
O

Poly(oxymethylene-1,4-phenyleneoxy-1,4-phenylenemethyleneoxy-2,2,3,3,4,4hexafluoropentamethylene)

O
CH2

O CH2

(CF2)3

CH2

257

O CH2

2
Poly(oxymethylene-5-tert-butyl-1,3-phenylenemethyleneoxycarbonylimino-1,4-phenylenemethylene1,4-phenyleneiminocarbonyl)

O CH2

CH2

O C NH

CH2

3
425

4
993

5
387

6
428

7
444

233

635

357

367

350

180

874

220

206

367

153

750

220

204

367

142

707

214

201

210

1123

197

187

351, 352,
366
351

154

550

248

280

446

NH C
O

H3C C CH3
CH3
748.

Poly(oxymethyleneoxy-1,4-phenylenecarbonyloxycarbonyl-1,4-phenylene)

O CH2

C O C
O

749.

Poly(oxymethyleneoxy-2,2,3,3,4,4,5,5-octafluorohexamethylene)

OCH2OCH2(CF2)4CH2
750.

Poly(oxymethyleneoxy-2,2,3,3,4,4-hexafluoropentamethylene)

OCH2OCH2(CF2)3CH2
751.

Poly(oxymethyleneoxyethylenedithioethylene)

OCH2O(CH2)2SS(CH2)2
752.

Poly(oxymethyleneoxytetramethylenedithiotetramethylene)

OCH2O(CH2)4SS(CH2)4
753.

Poly(oxymethylpentafluorophenylsilylene)

CH3
O

Si

F
F

258

1
747.

1
754.

2
Poly(oxymethylpentafluorophenylsilyleneoxydimethylsilylene)

CH3
O

Si

3
226

4
1027

5
208

6
220

7
446

232

826

303

281

279

217

682

341

318

279

381

1140

303

334

316

402

1104

345

364

442

CH3
O

Si

F CH3

F
F

755.

Poly(oxyneopentyleneoxycarbonyl-1,4-cyclohexylenecarbonyl) trans

CH3
O CH2

C CH2
CH3

756.

CH2
O C CH
CH2
O

CH2
CH C
CH2
O

Poly(oxyneopentyleneoxyterephthaloyl)

CH3
O CH2

C CH2

O C

CH3
757.

Poly(oxyneopentylenesulfonylneopentyleneoxycarbonyliminohexamethyleneiminocarbonyl)

CH3
O CH2

CH3

C CH2

SO2

CH3
758.

CH2

C CH2

O C NH

CH3

(CH2)6

NH C

Poly(oxynonamethyleneoxycarbonyimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

(CH2)9

O C NH

NH C
O

259

CH2

2
Poly(oxyoctamethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

(CH2)8

O C NH

CH2

(CH2)5

(CH2)4

O
762.

(CH2)5

O C NH

CH2

(CH2)5

O C NH
O

764.

(CH2)5

316

350

210

864

204

243

290

333

852

368

391

442

258

721

325

358

442

262

787

311

333

411

O
NH C
CH3

Poly(oxypentamethyleneoxycarbonyl-2,6-naphthylenecarbonyl)

326

NH C

Poly(oxypentamethyleneoxycarbonylimino-4-methyl-1,3-phenyleneiminocarbonyl)

953

O
763.

301

Poly(oxypentamethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

7
442

Poly(oxypentamethyleneoxyadipoyl)

6
370

C O C
O

761.

5
352

Poly(oxypentamethyleneoxy-1,4-phenylenecarbonyloxycarbonyl-1,4-phenylene)

O (CH2)5

4
1041

NH C

O
760.

3
385

C
O

C
O

260

1
759.

1
765.

2
Poly(oxypentamethyleneoxyterephthaloyl)

O
766.

(CH2)5

Poly(oxypimeloyloxy-2,6-dimethyl-1,4-phenyleneisopropylidene-3,5-dimethyl-1,4-phenylene)

H3C

3
217

4
682

5
283, < 318

6
318

7
306, 310,
311, 312

409

1041

357

393

404

358

913

369

392

404

60.7

266

198, 201

228

330, 333,
688

CH3
CH3

(CH2)5

O C
O

C O

CH3
H3C

767.

Poly(oxypimeloyloxy-3,3',5,5'-tetramethyl-4,4'-biphenylylene)

H3C
(CH2)5

O C
O

Poly(oxypropylene)

CH2

CH3

C O
H3C

768.

CH3

CH O

CH3

CH3

261

2
Poly(oxyterephthaloyloxy-1,4-phenylene-9,9-anthronylidene-1,4-phenylene)

O C
O

3
455

4
753

5
570, 627

6
604

7
395

374

763

473

490

425

437

776

654, 590

563

395

364

704

534

517

439

C O
O

C
C
O

770.

Poly(oxyterephthaloyloxy-1,4-phenylenebenzylidene-1,4-phenylene)

O C
O

771.

CH

Poly(oxyterephthaloyloxy-1,4-phenylenefluoren-9-ylidene-1,4-phenylene)

O C
O

772.

C O

C O
O

Poly(oxyterephthaloyloxy-1,4-phenylenehexafluoroisopropylidene-1,4-phenylene)

CF3
O C
O

C O

CF3

262

1
769.

1
773.

2
Poly(oxyterephthaloyloxy-1,4-phenyleneisopropylidene-1,4-phenylene)

3
333

4
681

5
478

6
489

7
400, 403,
404, 407,
426, 435,
438, 439

399

778

498

513

404

360

865

463, 413

416

436

434

1048

403, 413

414

289, 293,
294, 436

366

839

~428, 461

436

401, 417,
436

CH3
O C
O
774.

C O

CH3

Poly(oxyterephthaloyloxy-2,6-dimethyl-1,4-phenyleneisopropylidene-3,5-dimethyl-1,4-phenylene)

H3C

CH3

CH3
O C
O

C O

CH3
H3C

775.

CH3

Poly(oxyterephthaloyloxy-2-chloro-1,4-phenyleneisopropylidene-3-chloro-1,4-phenylene)

Cl

Cl
CH3

O C
O
776.

C O

CH3

Poly(oxyterephthaloyloxy-2-isopropyl-1,4-phenyleneisopropylidene-3-isopropyl-1,4-phenylene)
CH3
CH3

CH3

CH

CH CH3
CH3

O C
O
777.

C O

CH3

Poly(oxyterephthaloyloxy-2-methyl-1,4-phenyleneisopropylidene-3-methyl-1,4-phenylene)

H3C

CH3
CH3

O C

CH3

263

C O

2
Poly(oxyterephthaloyloxy-2-sec-butyl-1,4-phenylene-isopropylidene-3-sec-butyl-1,4-phenylene)

3
468

4
1225

5
373

6
382

7
436

302

1079

298, 268

280

274, 416,
417

251

839

276

299

417

234

762

318, 264

307

274, 416,
417

256

721

358

355

279, 437

CH3
O C
O

779.

782.

CH3

O
CH2

CH2

CH CH3

CH CH3

CH3

CH3

C O

(CH2)10

Poly(oxyterephthaloyloxyheptamethylene)

O C
781.

Poly(oxyterephthaloyloxydecamethylene)

O C
780.

C O

C O

(CH2)7

Poly(oxyterephthaloyloxyhexamethylene)

O C

(CH2)6

Poly(oxyterephthaloyloxymethylene-1,4-cyclohexylenemethylene) 70% trans

O C
O

C O CH2
O

CH2
CH
CH2

CH2
CH CH2
CH2

264

1
778.

1
783.

2
Poly(oxyterephthaloyloxyneopentylenesulfonylneopentylene)

CH3
O C
O
784.

785.

786.

C O CH2

CH2

CH3

SO 2

CH2

6
376

7
316

285

1000

308, 270

285

274, 416,
417

268

921

318

291

416

217

682

318

318

274, 416

CH2

CH3

(CH2)8

Poly(oxyterephthaloyloxypentamethylene)

O C

5
378

(CH2)9

Poly(oxyterephthaloyloxyoctamethylene)

O C

4
878

CH3

Poly(oxyterephthaloyloxynonamethylene)

O C

3
330

(CH2)5

787.

Poly(oxytetrafluoroethylene)

64.4

330

225

195

356, 357

788.

Poly(oxytetramethylene)

77.8

374

189, 178

208

346, 347,
348, 695

789.

Poly(oxytetramethylenedithiotetramethylene)

183

863

197

212

351, 352

790.

Poly(oxytetramethyleneoxy-1,4-phenylenecarbonyloxycarbonyl-1,4-phenylene)

284

874

348

325

350

OCF2CF2
O(CH2)4

O(CH2)4SS(CH2)4
O

(CH2)4

C O C
O

265

2
Poly(oxytetramethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenyleneoxy-1,4-phenylenesulfonyl1,4-phenyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene)
CH3
O

(CH2)4

(CH2)4

SO2

O C NH

CH2

O
793.

794.

(CH2)4

CH2
O C CH
CH2
O

(CH2)4

7
297

316

788

382

401

442

215

824

263

261

279

245

706

349

347

411

261

1106

216

236

273, 408,
427

200

602

290, 353

332

274, 279,
416, 417

CH3

NH C

CH2
CH C
CH2 O

Poly(oxytetramethyleneoxycarbonyl-2,6-napthylenecarbonyl)

6
396

Poly(oxytetramethyleneoxycarbonyl-1,4-cyclohexylenecarbonyl)

5
413

CH3

Poly(oxytetramethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

4
1735

C
CH3

792.

3
687

O C
O

C
O

795.

Poly(oxytetramethyleneoxysebacoyl)

O
796.

(CH2)4

O C

(CH2)8 C

Poly(oxytetramethyleneoxyterephthaloyl)

(CH2)4

O C

266

1
791.

1
797.

Poly(oxytrimethylene)

3
60.7

4
295

5
195

6
206

7
347, 362,
363

798.

Poly(oxytrimethyleneoxy-1,3-phenylenecarbonyl-oxycarbonyl-1,3-phenylene)

267

878

326

304

350

267

795

368

336

350

176

707

214

249

290

299

726

392

412

442

207

570

345, 213

363

442, 744

198

744

267

266

279

O(CH2)3
O

(CH2)3

C O C
O

799.

Poly(oxytrimethyleneoxy-1,4-phenylenecarbonyloxycarbonyl-1,4-phenylene)

(CH2)3

C O C
O

800.

(CH2)3

O C

(CH2)4

O
701.

Poly(oxytrimethyleneoxyadipoyl)

C
O

Poly(oxytrimethyleneoxycarbonylimino-1,4-phenylenemethylene-1,4-phenyleneiminocarbonyl)

(CH2)3

O C NH

CH2

NH C

O
802.

(CH2)3

O C NH
O

803.

Poly(oxytrimethyleneoxycarbonylimino-4-methyl-1,3-phenyleneiminocarbonyl)

NH C
CH3

Poly(oxytrimethyleneoxycarbonyl-1,4-cyclohexylenecarbonyl) trans

(CH2)3

CH2
CH C
CH2 O

267

CH2
O C CH
CH2
O

2
Poly(oxytrimethyleneoxycarbonyl-2,6-napthylenecarbonyl)

(CH2)3

3
228

4
626

5
346

6
364

7
411

183

523

368, 308

350

279, 321,
416, 417

260

585

453

445

712

207

456

420

454

560

165

604

268 Brittle
point

273

581, 706

O C
O

C
O

805.

Poly(oxytrimethyleneoxyterephthaloyl)

O
806.

(CH2)3

O C

Poly(pentabromobenzyl acrylate)

CH2

Br

Br

CH
C O

Br

CH2

O
Br
807.

Poly(pentachlorophenyl acrylate)

CH2

Cl

Br
Cl

CH
C O

Cl

O
Cl
808.

Poly(pentyl methacrylate)

CH3
CH2

O C O (CH2)4

CH3

Cl

268

1
804.

1
809.

2
Poly(pentylethylene)

CH2

3
137

4
620

5
242

6
221

7
509

150

326

467

460

291

138

380

330

363

560, 585

172

558

270

308

560

94.4

260

359, 383,
353, 365

363

692, 709,
713, 720

129

586

207

220

600, 621

CH
(CH2)4
CH3

810.

Poly(perfluorostyrene)

CF2

CF

F
F

811.

Poly(phenyl acrylate)

CH2

CH
C O
O

812.

Poly(phenyl ethyl acrylate)

CH2

CH
C O

(CH2)2

O
813.

Poly(phenylene sulfide)

S
814.

Poly(pentyloxyethylene)

CH2

CH
(CH2)4

CH3

269

2
Poly(p-phenylene terephthalamide) Kevlar

NH C

CH2

CH2

7
709

94.9

409

224

232

588, 621

96.4

353

283

273

650

114

452

236 Brittle
point, 228

252

589, 706

128

383

344 Vicat
softening
point

334

588, 599

CH3

Poly(propionyloxyethylene)

CH2

6
601

CH
O

817.

5
508, 580 600

Poly(propoxyethylene)

CH2

4
353

C NH

O
816.

3
212

CH
O
C

CH2

CH3

O
818.

Poly(propyl acrylate)

CH2

CH
C O (CH2)2

CH3

O
819.

Poly(propyl chloroacrylate)

Cl
CH2

C
C O CH2
O

CH2

CH3

270

1
815.

1
820.

2
Poly(propylethylene)

CH2

3
85.3

4
363

5
233

6
235

160

366

409
softening
point

437

7
362, 496,
503, 509,
510, 512,
522, 528,
543
287

112

434

253

258

600, 620

131

476

251, 253,
256

275

582, 583

145

407

347 Vicat
softening
point

356

599

CH
CH2
CH2

821.

CH3

Poly(2,4,5-trimethylstyrene)

CH2

CH
CH3

H3C
CH3
822.

Poly(sec-butoxyethylene)

CH2

CH
O
CH CH3
CH2

823.

CH3

Poly(sec-butyl acrylate) conventional

CH2

CH
C O CH CH2
O

824.

CH3

CH3

Poly(sec-butyl chloroacrylate)

Cl
CH2

C
C O CH CH2
CH3

271

CH3

2
Poly(sulfonyl-1,2-cyclohexylene)

4
300

5
401

6
393

7
316

118

300

381

393

316

309

724

413

427

467, 468

309

681

467

453

467, 468

360

940

398

383

467, 468

362

674

< 573
softening
point

537

376

CH CH

SO2

CH2

H2C
H2C
826.

3
118

CH2

Poly(sulfonyl-1,3-cyclohexylene)

CH2
SO 2

CH

CH

CH2

CH2
CH2
827.

Poly(sulfonyl-1,3-phenyleneiminoadipoylimino-1,3-phenylene)

SO 2

NH C

(CH2)4

O
828.

Poly(sulfonyl-1,3-phenyleneiminoadipoylimino-1,4-phenylene)

SO 2

NH C

(CH2)4

O
829.

C NH
O

Poly(sulfonyl-1,3-phenyleneiminoazelaoylimino-1,3-phenylene)

SO 2

NH C

(CH2)7

O
830.

C NH

C NH
O

Poly(sulfonyl-1,3-phenyleneiminocarbonyl-1,4-naphthylenecarbonylimino-1,3-phenylene)

SO 2

NH C
O

C NH
O

272

1
825.

1
831.

2
Poly(sulfonyl-1,3-phenyleneiminododecanedioylimino-1,3-phenylene)

SO 2

NH C

(CH2)10

O
832.

NH C

(CH2)8

NH C

7
467, 468

377

997

385

378

467, 468

343

884

398

388

467, 468

342

621

568 - 583

551

376

360

857

451

420

467, 468

362

591

< 573
softening
point

613

376

(CH2)6

C NH

O
834.

6
371

C NH

Poly(sulfonyl-1,3-phenyleneiminosuberoylimino-1,3-phenylene)

SO 2

5
380

O
833.

4
1111

C NH

Poly(sulfonyl-1,3-phenyleneiminosebacoylimino-1,3-phenylene)

SO 2

3
412

Poly(sulfonyl-1,4-phenyleneimino-2-methoxyisophthaloylimino-1,4-phenylene)

O CH3
SO2

NH C

C NH

O
835.

Poly(sulfonyl-1,4-phenyleneiminoazelaoylimino-1,4-phenylene)

SO 2

NH C

(CH2)7

O
836.

C NH
O

Poly(sulfonyl-1,4-phenyleneiminocarbonyl-1,4-naphthylenecarbonylimino-1,4-phenylene)

SO2

NH C
O

C NH
O

273

2
Poly(sulfonyl-1,4-phenyleneiminocarbonyl-1,4-phenylenemethylene-1,4-phenylenecarbonylimino-1,4phenylene)

SO 2

NH C

CH2

O
838.

NH C

(CH2)10

NH C

(CH2)5

NH C

(CH2)8

NH C
O

842.

401

467, 468

326

721

436

452

467, 468

377

913

444

413

467, 468

343

800

453

429

467, 468

410

707

590

580

376

(CH2)6

C NH
O

Poly(sulfonyl-1,4-phenyleneiminoterephthaloyl-1,4-phenylenecarbonylimino-1,4-phenylene)

SO2

433

C NH

Poly(sulfonyl-1,4-phenyleneiminosuberoylimino-1,4-phenylene)

SO 2

1027

O
841.

412

C NH

Poly(sulfonyl-1,4-phenyleneiminosebacloylimino-1,4-phenylene)

SO 2

7
376

O
840.

6
564

C NH

Poly(sulfonyl-1,4-phenyleneiminopimeloylimino-1,4-phenylene)

SO 2

5
< 573
softening
point

O
839.

4
723

C NH

Poly(sulfonyl-1,4-phenyleneiminododecanedioylimino-1,4-phenylene)

SO 2

3
408

NH C

C NH

274

1
837.

1
843.

2
Poly(sulfonyl-1,4-phenylenemethylene-1,4-phenylene)

SO 2
844.

3
196

4
371

5
497

6
529

7
383

366

839

436

423

131

411

380, 316,
346, 313,
304

319

560, 583,
584, 706,
709

102

311

337 Highly
crystalline
sample
softening
point

328

511

301

1115

297 Brittle
point

270

277, 581

318

1325

201-264

240

593

CH2

Poly(terephthaloyloxy-2-methyl-1,4-phenyleneisopropylidene-3-methyl-1,4-phenyleneoxy)

H3C

CH3
CH3

845.

C O

CH3

Poly(tert-butyl acrylate)

CH2

CH

CH3

C O C

CH3
CH3

O
846.

Poly(tert-butylethylene)

CH2

CH

H3C

CH3

CH3
847.

Poly(tetradecyl acrylate)

CH2

CH

O C O (CH2)13
848.

CH3

Poly(tetradecyl methacrylate)

CH3
CH2

C
CH3

275

O C O (CH2)13

2
Poly(tetradecylethylene)

CH2

3
273

4
1026

5
246

6
266

7
509, 512,
521

CH
(CH2)13
CH3

850.

Poly(tetrathiodecamethylene)

246

1159

197

212

461

851.

Poly(tetrathioethylene)

109

717

249

212

351, 352

852.

Poly(tetrathiomethylene-1,4-phenylenemethylene)

184

848

276

284

461

109

405

256, 228

269

452, 453

109

405

221

269

453

564

1790

331

315

463

SSSS(CH2)10
SSSS(CH2)2
S

853.

CH2

Poly(thio-1,2-cyclohexylene)

CH CH

H2C

CH2

H2C
854.

CH2

CH2

Poly(thio-1,3-cyclohexylene)

CH2
CH
CH
CH2

CH2
CH2

855.

Poly(thio-11-oxoundecamethyleneiminoethylene-1,4-phenyleneethyleneimino-1-oxoundecamethylene)

(CH2)10

C NH
O

(CH2)2

(CH2)2

NH C
O

(CH2)10

276

1
849.

1
856.

2
Poly(thio-1-ethylethylene)

CH2

858.

CH CH2

CH3

Poly(thio-1-methyltrimethylene)

CH2

5
218

6
217

7
330

88.1

301

285

293

466

86.5

388

214

223

462

86.5

333.2

259

260

677

121

504

~223

240

316

139

539

293

258

465

CH3

Poly(thio-1-methyl-3-oxotrimethylene)

4
399

CH
CH2

857.

3
86.5

CH2 CH
CH3

859.

Poly(thio-2,2-dimethylene) Poly(thioisobutylene)

CH3
S

CH2

C
CH3

860.

Poly(thio-2-ethyl-2-methyltrimethylene)

CH3
CH2
S

CH2

C CH2
CH3

861.

Poly(thio-3-methyl-6-oxohexamethylene)

(CH2)2

CH
CH3

(CH2)2

C
O

862.

Poly(thiodifluoromethylene)

45.7

250

155

183

456-458

863.

Poly(thiomethylene)

35.3

205

218

172

454

SCF2

277

SCH2

2
Poly(thioneopentylene)

3
104

4
414

5
233

6
251

7
316, 453

69.5

309

226

225

330, 454,
455

CH3
S

CH2

C CH2
CH3

865.

Poly(thiopropylene)

CH2

CH
CH3

866.

Poly(thiotrimethylene)

69.5

364

~228

191

453, 459,
460

867.

Poly(tridecanolactone)

233

1031

237

226

763

329

1006

293

327

721

49.7

171

304

291

445

159

416
(random),
489 (iso),
365
(sindio)

S(CH2)3

(CH2)12

C O
O

868.

Poly(triethylene glycol p,p'-dibenzoate)


C

869.

C O CH2

O
Poly(trifluoroethylene)

CF2

CH2

O CH2

CH2

O CH2

CH2

CH
F

870.

Poly(trimethylsilyl methacrylate)

CH3
CH2

C
C O
O

CH3
Si CH3
CH3

341 (iso),
382
(random),
400
(sindio) 325 (iso),
436
(sindio)

609

278

1
864.

1
871.

Poly(trithiodecamethylene)

3
227

4
1064

5
203

6
213

7
461

872.

Poly(trithiomethylene-1,4-phenylenemethylene)

165

555

291

297

461

424

1317

319

322

441

458

1427

324

321

441

560

1767

321

317

441

355

1086

328

327

441

389

1201

323

324

441

SSS(CH2)10
S

873.

CH2

CH2

Poly(ureylene-1,4-phenylenemethylene-1,4-phenyleneureylenedecamethylene)

NH C NH

CH2

O
874.

Poly(ureylene-1,4-phenylenemethylene-1,4-phenyleneureylenedodecamethylene)

CH2

(CH2)12

O
CH2

NH C NH

(CH2)6

NH C NH

CH2

Poly(ureyleneoctamethyleneureylene-1,4-phenylenemethylene-1,4-phenylene)

NH C NH

(CH2)8

NH C NH
O

CH2
279

(CH2)18

Poly(ureylenehexamethyleneureylene-1,4-phenylenemethylene-1,4-phenylene)

NH C NH
877.

NH C NH

Poly(ureylene-1,4-phenylenemethylene-1,4-phenyleneureyleneoctadecamethylene)

NH C NH

876.

(CH2)10

NH C NH
875.

NH C NH

2
Poly(vinyl trimethyl silane)

3
115

4
314

5
378-417

6
366

7
693

159

564

286

282

674

182

534

~349

341

278, 628

346

1442

188, 250

240

665

142

393

372

361

665

CH

CH2
H3C

Si

CH3

CH3
879.

Poly[(1,2-diethoxycarbonyl)ethylene]

CH CH
O C

C O

O
H3C
880.

CH2 CH2

CH3

Poly(4-acetoxybenzoyloxyethylene)

CH2

CH
O C CH3

O C
O
881.

Poly[(1-heptoxycarbonyl-1-heptoxycarbonylmethylene)ethylene]
O
C O
CH2

(CH2)6

CH3

C O (CH2)6

CH3

C
CH2
O

882.

Poly[(1-methoxycarbonyl-1-methoxycarbonylmethylene)ethylene]
O
C O CH3
CH2

C
CH2
C O CH3
O

280

1
878.

1
883.

2
Poly[(1-methylcyclohexanoyloxy)ethylene]

CH2

CH3

CH2

5
359

6
353

7
650

165

571

283

289

650

182

534

333

341

278, 628

230

552

345

417

545

CH2
CH2
CH2

CH2

Poly[(2,2-dimethylvaleryloxy)ethylene]

CH2

4
482

CH

O
884.

3
170

CH
O
C O

H3C C

(CH2)2

CH3

CH3
885.

Poly[(2-acetoxybenzoyloxy)ethylene]

CH2

CH
O C
O
O C

CH3

O
886.

Poly[(2-benzoyloxymethyl)styrene]

CH2

CH
CH2

O C
281

2
Poly[(2-butoxyethoxy)methylstyrene]

CH2

O (CH2)3

Poly[(2-ethylhexyloxy)ethylene]

CH2

4
925

5
< 235

6
268

7
552

180

779

207

231

600, 621

164

488

338

336

278, 628

155

443

321

349

278, 580,
628, 653

CH
CH2 O (CH2)2

888.

3
248

CH3

CH
O
CH2 CH CH2
CH2

889.

CH2 CH3

CH3

Poly[(2-methoxybenzoyloxy)ethylene]

CH2

CH2

CH
O
C
O
O CH3

890.

Poly[(2-methylbenzoyloxy)ethylene]

CH2

CH
O
C
O
CH3

282

1
887.

1
891.

2
Poly[(2-nitrosoethyl) methacrylate]

3
125

4
411

5
328

6
304

7
610

164

488

~317

336

278, 628

155

443

324

349

278, 628,
653

155

439

366

353

278, 628

CH3
CH2

C
C O CH2

CH2

N O

O
892.

Poly[(3-methoxybenzoyloxy)ethylene]

CH2

CH
O
C
O
O CH3

893.

Poly[(3-methylbenzoyloxy)ethylene]

CH2

CH
O
C
O
CH3

894.

Poly[(3-nitrobenzoyloxy)ethylene]

CH2

CH
O
C
NO 2

283

2
Poly[(3-trimethylsilylbenzoyloxy)ethylene]

CH2

3
217

4
573

5
353

6
378

7
659

218

752

267

290

489

276

857

325

322

489

CH
O
C
O
H3C

Si

CH3

CH3
896.

Poly[(4-dimethylaminophenyl)methylsilylenetrimethylene]

H3C N CH3

Si

(CH2)3

CH3
897.

Poly[(4-dimethylaminophenyl)phenylsilylenetrimethylene]

H3C N CH3

Si

(CH2)3

284

1
895.

1
898.

2
Poly[(4-methoxybenzoyloxy)ethylene]

CH2

3
164

4
488

5
360

6
336

7
278, 570,
628, 652,
653

155

443

343

350

278, 628,
652, 653

155

439

395

353

278, 628

306

1093

320

280

561

CH
O
C

O CH3

O
899.

Poly[(4-methylbenzoyloxy)ethylene]

CH2

CH
O
C

CH3

O
900.

Poly[(4-nitrobenzoyloxy)ethylene]

CH2

CH
O
C

NO 2

O
901.

Poly[(4-pentadecafluoroheptyl)styrene]

CH2

CH

CF3

285

(CF2)6

2
Poly[(4-phenylbenzoyloxy)ethylene]

CH2

3
214

4
552

5
358

6
388

7
278, 628

200

571

346

350

278, 628

206

691

313

298

651

217

573

408

379

659, 660

142

473

348

300

514

CH
O
C
O

903.

Poly[(4-propionyloxybenzoyloxy)ethylene]

CH2

CH
O

904.

O C

CH2

CH3

Poly[(4-p-toluoylbutyryloxy)ethylene]

CH2

CH
O
C

(CH2)3

CH3

O
905.

Poly[(4-trimethylsilylbenzoyloxy)ethylene]

CH2

906.

CH
O

CH3

Si

CH3

CH3

Poly[(cyclohexylmethyl)ethylene]

CH2

CH
CH2

CH2
CH
CH2

CH2
CH2
CH2

286

1
902.

1
907.

2
Poly[(cyclopentylmethyl)ethylene]

CH2

908.

CH

CH2
CH2 CH
CH2

4
391

5
333

6
317

7
514

131

434

~328, 338

302

591

172

656

243

262

490

314

1292

253255

243

656

177

670

288293

264

656

CH2
CH2

Poly[(heptafluoro-2-propoxy)ethylene]

CH2

3
124

CH
O

F 3C
909.

CF

CF3

Poly[(methyl)phenylsilylenetrimethylene]

CH3
Si

910.

(CH2)3

Poly[(nonadecafluorodecanoyloxy)ethylene]

CH2

CH
O
C

(CF2)8

CF3

O
911.

Poly[(nonafluorovaleryloxy)ethylene]

CH2

CH
O
C

CF3
287

(CF2)3

2
Poly[(pentadecafluorooctanyloxy)ethylene]

CH2

3
259

4
1040

5
258263

6
249

7
656

122

421

315

290

656

142

362

393

392

654

94.9

297

~319,
< 348

319

656, 657

210

652

327

321

656

CH
O
C

(CF2)6

CF3

O
913.

Poly[(pentafluoropropionyloxy)ethylene]

CH2

CH
O
C

CF2

CF3

O
914.

Poly[(tert-butoxycarbonylamino)ethylene]

CH2

CH
NH

CH3

O C O C CH3
CH3
915.

Poly[(trifluoroacetoxy)ethylene]
CH2 CH

O
C
916.

CF3

O
Poly[(undecafluorocyclohexylcarbonyloxy)ethylene]
CH2 CH

CF2
C CF
CF2
O

CF2
CF2
CF2

288

1
912.

1
917.

2
Poly[2-(2-dimethylaminoethoxycarbonyl)styrene]

CH2

3
218

4
597.6

5
342

6
365

7
551

199

610

363

326

650

199

599

323

332

650

199

610

328

326

650

CH
CH3
C O

(CH2)2

N
CH3

O
918.

Poly[2,2,3,3-tetramethylvaleryloxy)ethylene]

CH2

CH
O CH3 CH3
C

CH2

CH3

O CH3 CH3
919.

Poly[2,2,3,4-tetramethylvaleryloxy)ethylene]

CH2

CH
O CH3
C

CH

CH

CH3

O CH3 CH3 CH3


920.

Poly[2,2,4,4-tetramethylvaleryloxy)ethylene]

CH2

CH
O CH3
C

CH2

C
CH3

CH3
289

O CH3

CH3

2
Poly[2,2-difluoro-2-(2-heptafluorotetrahydrofuranyl)ethyl acrylate]

CH2

CH

C O CH2

923.

5
275 Brittle
point

6
283

7
587

261

529

~471
(softening
point)

493

546

192

608

319

316

560

272

1000

250

272

552

F2

F2

Poly[3-(4-biphenyl)styrene]

CH2

4
739

CF2

O
922.

3
209

F2

CH

Poly[3-chloro-2,2-bis(chloromethyl)propyl acrylate]

Cl
CH2

CH

CH2

C O CH2

CH2

CH2Cl

Cl
924.

Poly[4-(1-ethylhexyloxymethyl)styrene]

CH2

CH
CH3
CH2
CH2

O CH (CH2)4

CH3

290

1
921.

1
925.

2
Poly[4-(1-hydroxy-1-methylbutyl)styrene]

CH2

5
~403
softening
point

6
365

7
567

168

375

~438
softening
point

448

567

236

731

~364
softening
point

323

567

CH3

CH2

CH2

CH3

Poly[4-(1-hydroxy-1-methylethyl)styrene]

CH2

4
553

CH

OH C
926.

3
202

CH

OH C

CH3

CH3
927.

Poly[4-(1-hydroxy-1-methylhexyl)styrene]

CH2

CH

OH C

CH3
CH3

291

(CH2)4

2
Poly[4-(1-hydroxy-1-methylpentyl)styrene]

CH2

7
567

185

464

459

399

567

246

665

323 low
viscosity

370

567

CH

CH3

CH2

CH3

Poly[4-(1-hydroxy-3-morpholinopropyl)styrene]

CH2

6
341

CH3

Poly[4-(1-hydroxy-1-methylpropyl)styrene]

HO C
930.

5
356
softening
point

CH3

(CH2)3
CH2

4
642

CH

HO C
929.

3
219

CH

CH2
HO CH

(CH2)2

N
CH2

CH2
O
CH2

292

1
928.

1
931.

2
Poly[4-(1-hydroxy-3-piperidinopropyl)styrene]

CH2

HO CH

(CH2)2

N
CH2

5
327 low
viscosity

6
369

7
567

212

656

319

323

552

240

677.4

314

354

567

CH2
CH2
CH2

Poly[4-(2-hydroxybutoxymethyl)styrene]

CH2

4
688

CH

CH2

932.

3
254

CH

OH
CH2
933.

CH2

CH CH2

Poly[4-(3-morpholinopropionyl)styrene]

CH2

CH

CH2
O

CH3

(CH2)2

293

CH2

CH2
O
CH2

2
Poly[4-(3-piperidinopropionyl)styrene]

CH2

CO

(CH2)2

N
CH2

Poly[4-(4-biphenyl)styrene]

CH2

936.

937.

4
701

5
311

6
353

7
567

261

529

593
(softening
point)

493

546

212

721

293

294

552

272

1067

231

255

552

CH

CH2

935.

3
248

CH2
CH2
CH2

CH

Poly[4-(4-hydroxybutoxymethyl)styrene]
CH2 CH

CH2 O (CH2)4
Poly[4-(octyloxymethyl)styrene]
CH2 CH

CH2

(CH2)7

OH

CH3

294

1
934.

1
938.

2
Poly[4-(sec-butoxymethyl)styrene]

CH2

3
204

4
646

5
313

6
316

7
552

262

630

433

416

297

290

674

446

430

380

273

591

483

462

297

280

1284

210

218

451

CH

CH3
CH2 O CH CH2
939.

CH3

Poly[di(oxy-1,4-phenylene)carbonyl-1,4-phenylene]

C
O

940.

Poly[di(oxy-1,4-phenylene)oxyisophthaloyl]

941.

Poly[di(oxy-1,4-phenylene)sulfonyl-1,4-phenylene]

O
942.

O C

SO 2

Poly[di(oxydimethylsilylene)-1,4-phenylenedimethylsilylene]

CH3
O

Si

CH3

Si

Si

CH3

CH3

295

CH3

CH3

2
Poly[di(oxyethylene)oxy-1,4-phenylenecarbonyl-oxycarbonyl-1,4-phenylene]

(CH2)2

(CH2)2

(CH2)4

C O CH2

O
945.

O CH2

CH2

(CH2)7

C O CH2

CH2

O CH2

CH2

(CH2)2

(CH2)12

O
948.

(CH2)2

O C NH

(CH2)6

290

254

1081

205

235

290

260

916

272

284

349

339

1480

202

229

290

185

758

226

244

290

151

602

244

251

290

C O CH2

CH2

O CH2

CH2

(CH2)3

C O CH2

CH2

O CH2

CH2

Poly[di(oxyethylene)oxymalonyl]

241

O
949.

227

NH C

Poly[di(oxyethylene)oxyglutaryl]

838

Poly[di(oxyethylene)oxydodecanedioyl]

202

O
947.

7
350

Poly[di(oxyethylene)oxycarbonyliminohexamethyleneiminocarbonyl)]

6
316

O
946.

CH2

Poly[di(oxyethylene)oxyazelaoyl]

5
314

Poly[di(oxyethylene)oxyadipoyl]

4
930

C O C
O

944.

3
294

CH2

C O CH2
O

CH2

O CH2

CH2

296

1
943.

1
950.

2
Poly[di(oxyethylene)oxymethylmalonyl]

O CH2

CH2

O CH2

CH2

O C
O

951.

CH

CH2

O CH2

CH2

O C

4
625

5
244

6
269

7
290

305

1344

214

227

290

441

1951

205

226

290

134

521

265

257

290

237

987

226

240

290

219

920

213

238

290

CH3 O

Poly[di(oxyethylene)oxynonylmalonyl]

O CH2

3
168

CH

O (CH2)8

C
O

CH3
952.

Poly[di(oxyethylene)oxyoctadecanedioyl]

O CH2

CH2

O CH2

CH2

O C

(CH2)18

O
953.

Poly[di(oxyethylene)oxyoxalyl]

O CH2
954.

CH2

O CH2

CH2

O C

Poly[di(oxyethylene)oxypentylmalonyl]

O CH2

CH2

O CH2

CH2

O C

CH

O (CH2)4

C
O

CH3
955.

Poly[di(oxyethylene)oxypimeloyl]

O CH2

CH2

O CH2

CH2

O C

C
O

297

(CH2)5

2
Poly[di(oxyethylene)oxypropylmalonyl]

O CH2

CH2

O CH2

CH2

O C
O

CH

CH2

3
185

4
714

5
235

6
259

7
290

271

1163

199

233

290

237

1000

212

237

290

168

680

244

247

290

259

794

325

326

350

179

909

218

197

368

341

804

422

424

444

CH3
957.

Poly[di(oxyethylene)oxysebacoyl]

O CH2

CH2

O CH2

CH2

O C

(CH2)8

O
958.

Poly[di(oxyethylene)oxysuberoyl]

O CH2

CH2

O CH2

CH2

O C

(CH2)6

O
959.

CH2

O CH2

CH2

O C

(CH2)2

C
O

Poly[di(oxymethylene)oxy-1,4-phenylenecarbonyl-oxycarbonyl-1,4-phenylene]

O CH2

O CH2

C O C
O

961.

Poly[di(oxyethylene)oxysuccinyl]

O CH2
960.

Poly[di(oxymethylene)oxy-2,2,3,3,4,4-hexafluoropentamethylene]

OCH2OCH2OCH2(CF2)3CH2
962.

Poly[imino-5-tert-butylisophthaloylimino-(2,5-dimethylhexamethylene)]

NH C

C NH

O
H3C

O
C CH3
CH3

CH2

CH
CH3

(CH2)2

CH CH2
CH3

298

1
956.

1
963.

2
Poly[imino-5-tert-butylisophthaloylimino-(3,4-dimethylhexamethylene)]

3
341

4
804

5
446

6
424

7
444

306

754

436

406

444

306

855

351

358

373

290

770

398

377

406

CH3
NH C

C NH

(CH2)2

CH

(CH2)2

CH3

H3C

CH

CH3

CH3
964.

Poly[imino-5-tert-butylisophthaloyliminohexamethylene]

NH C

C NH

(CH2)6

H3C

CH3

CH3
965.

Poly[iminomethylene(2,5-dimethyl-1,4-phenylene)methyleneiminosuberoyl]

H3C
NH CH2

CH2

NH C

(CH2)6

C
O

H3C
966.

Poly[iminomethylene-1,4-phenylenemethyleneiminocarbonyl(1,3,3-trimethyl-5-oxopentamethylene)]

CH3
NH CH2

CH2

NH C CH CH2

C CH2

CH3

299

CH3

2
Poly[oxy(allyloxymethyl)ethylene]

O CH2

O CH2

5
195

6
215

7
330

82.5

311

259

265

332

139

662

194

210

330

75.5

306

251, 248

247

332, 703

104

479

212

217

330

173

840

190

206

330

87.3

400

211

218

330

CH CH2

Poly[oxy(bromomethyl)ethylene]

O CH2

4
544

CH
CH2

968.

3
117

CH
CH2Br

969.

Poly[oxy(butoxymethyl)ethylene]

O CH2

CH
CH2
O

970.

(CH2)3

CH3

Poly[oxy(chloromethyl)ethylene]

O CH2

CH
CH2Cl

971.

Poly[oxy(ethoxymethyl)ethylene]

O CH2

CH
CH2

972.

O CH2

Poly[oxy(hexyloxymethyl)ethylene]

O CH2

CH
CH2 O

973.

CH3

(CH2)5

Poly[oxy(methoxymethyl)ethylene]

O CH2

CH
CH2 O CH3

CH3

300

1
967.

1
974.

2
Poly[oxy(methyl)phenylsilylene]

3
129

4
471

5
187, 273

6
274

7
748

340

960

331

354

448

293

807

394

363

431

306

691

416
Proposed
structure

443

377

CH3
O

975.

Si

Poly[oxy(methyl)phenylsilyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene]

CH3
O

Si

CH3
O

C
CH3

976.

Poly[oxy-1,3-phenyleneoxy-5-(heptafluoropropyl)isophthaloyl]

O C

O
CF2

977.

CF2

CF3

Poly[oxy-1,4-phenylene-(2-cyano)-2-phenyltrimethylene-1,4-phenylene]

C N
O

CH2

C CH2

301

2
Poly[oxy-1,4-phenylene-2,2-di(ethoxycarbonyl)trimethylene-1,4-phenylene]

CH2

3
335

4
1003

5
327

6
334

7
377

281

562

507

500

331

311

641

484

485

331

CH3

O
C O
O

CH2

C CH2
C O
O
CH2

979.

CH3

Poly[oxy-2-(1-naphthyl)-6-phenyl-1,4-phenylene]

980.

Poly[oxy-2-(2-biphenylyl)-6-phenyl-1,4-phenylene]

302

1
978.

1
981.

2
Poly[oxy-2-(2-chlorobenzoyloxy)trimethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene]
CH3

O CH2

CH CH2

3
406

4
1140

5
339

6
356

7
329

303

634

513

478

331

331

943

338

351

329

CH3

O C
CH2
Cl

982.

Poly[oxy-2-(4-tert-butylphenyl)-6-phenyl-1,4-phenylene]

CH3

CH3

CH3
983.

Poly[oxy-2-(chloroacetoxy)trimethyleneoxy-1,4-phenyleneisopropylidene-1,4-phenylene]

CH3
O CH2

CH CH2
O

C
CH3
303

O C CH2Cl

2
Poly[oxy-2,2-bis(chloromethyl)trimethylene]

3
124

4
420

5
265

6
295

7
273, 300,
339-345

375

1014

383

370

329

451

1212

399

372

329

445

1660

300

268

410

CH2Cl
O CH2

C CH2
CH2Cl

985.

Poly[oxy-2-acetoxytrimethyleneoxy-1,4-phenylenemethyl(phenyl)methylene-1,4-phenylene]

CH3
O CH2

CH CH2

O
C CH3
O
986.

Poly[oxy-2-benzoyloxytrimethyleneoxy-1,4-phenylenemethyl(phenyl)methylene-1,4-phenylene]

CH3
O CH2

CH CH2

O
C CH2
O
987.

Poly[oxy-5-(pentadecafluoroheptyl)isophthaloyloxy-2,2,3,3,4,4-hexafluoropentamethylene]

C O CH2

O
CF2

CF2

CF2

CF2

CF2

CF2

CF2

CF2

CF2

CH2

CF3

304

1
984.

1
989.

2
Poly[oxy-6-phenyl-2-(m-tolyl)-1,4-phenylene]

3
252

4
506

5
492

6
498

7
331

252

506

491

498

331

385

881

~478

437

337

162

372

500, < 433,


418

435

390392

H3C

990.

Poly[oxy-6-phenyl-2-(p-tolyl)-1,4-phenylene]

H3C
O

991.

Poly[oxycarbonyldi(oxy-1,4-phenylene)sulfonyl-1,4-phenyleneoxy-1,4-phenylene]

O C O

SO2

O
992.

Poly[oxycarbonyloxy-1,3-(2,2,4,4-tetramethylcyclobutylene)

H3C

305

CH3
C
O C O CH CH
C
O
H3C
CH3

2
Poly[oxycarbonyloxy-1,4-phenylene(methyl)phenylmethylene-1,4-phenylene]

3
297

4
661

5
449

6
449

7
384, 386

271

679

410

399

388

369

1525

231

242

446

CH3
O C O

994.

Poly[oxycarbonyloxy-4,6-dimethyl-1,2-phenylenemethylene-3,5-dimethyl-1,2-phenylene]

CH2
O C O
O
995.

H3C

CH3 H3C

Poly[oxydi(pentafluorophenyl)silylenedi(oxydimethylsilylene)]

F
F

F CH3
O

Si

Si

F CH3

F
F

CH3
O

Si
CH3

CH3

306

1
993.

1
996.

2
Poly[oxymethylpentafluorophenylsilylenedi(oxydimethylsilylene)]
CH3
CH3
CH3

997.

Si

Si

F CH3

Si
CH3

999.

CH3

Si

CH3

Si

CH3

CH3

Si

Si

CH3

CH3

CH3

CH3

CH3

CH2

CH2

CH2

CH2

CH2

CH3

Si

Si
CH3

Si

Si

Si

Si

CH2

CH2

CH2

CH2

CH2

CH3

CH3

CH3

CH3

CH3

Poly[penta(oxydiethylsilylene)-1,4-phenyleneoxy-1,4-phenylenedimethylsilylene]

CH3

CH3

CH3

CH3

CH3

CH2

CH2

CH2

CH2

CH2

Si

Si

6
213

7
446

538

2491

231

216

372

666

4215

193

158

372

750

4438

208

169

372

CH3

Poly[penta(oxydiethylsilylene)-1,4-phenylenedimethylsilylene]

5
190

CH3

CH3
O

4
1343

Si

F
Poly[oxytri(dimethylsilyleneoxy)(methyl)phenylsilylene-1,3-phenylene(methyl)phenylsilylene)]

CH3

998.

3
286

Si

Si

Si

CH2

CH2

CH2

CH2

CH3

CH3

CH3

CH3

CH3

Si
CH3
307

CH2

CH3

CH3
O

Si

CH3
O

CH3

Si

CH3
O

CH3

Si

CH3

CH3

CH3

Si

Si

CH3

CH3

1001. Poly[tetra(oxydimethylsilylene)-1,4-phenylenedimethylsilylene]

CH3
O

Si

CH3

CH3
O

CH3

Si

CH3

Si

CH3

CH3

CH3

Si

Si

CH3

CH3

1002. Poly[tetra(oxydimethylsilylene)-1,4-phenyleneoxy-1,4-phenylenedimethylsilylene]

CH3
O

Si
CH3

CH3
O

Si

CH3
O

CH3

Si
CH3

Si

6
185

7
372

423

2238

201

189

372

507

2449

221

207

372

126

581

213

217

330

347

848

423

409

297

357

808

453

442

297

Si

CH CH2
CH2

O CH2

CH CH2

1004. Poly[tri(oxy-1,4-phenylene)carbonyl-1,4-phenylene]

5
198

CH3

CH3

1003. Poly[thio-1-(allyloxymethyl)ethylene]

4
2286

CH3

CH3
O

3
423

C
O

1005. Poly[tri(oxy-1,4-phenylene)sulfonyl-1,4-phenylene]

SO2

308

1
2
1000. Poly[tetra(oxydimethylsilylene)-1,3-phenylenedimethylsilylene]

1
2
1006. Poly[tri(oxydimethylsilylene)-1,4-phenylenedimethylsilylene]

CH3
O

Si

CH3
O

CH3

Si

CH3

CH3

CH3

Si

Si

CH3

CH3

1007. Poly[tri(oxydimethylsilylene)oxy(methyl)-2-phenylethylsilylene]

3
351

4
1764

5
211

6
199

7
372

380

2159

171

176

449

344

1901

201

181

449

359

1962

148

183

449

CH2
CH3
O

Si

CH3
O

CH3

Si

CH3
O

CH3

Si

CH2
O

CH3

Si
CH3

1008. Poly[tri(oxydimethylsilylene)oxy(methyl)phenylsilylene]

CH3

CH3
O

Si

CH3

Si

CH3
O

CH3

Si

CH3
O

Si

CH3

1009. Poly[tri(oxydimethylsilylene)oxy(methyl)trimethylsiloxysilylene]

CH3
H3C
CH3
O

Si

Si
CH3

CH3
O

Si
CH3

O
O

Si
CH3

309

CH3

CH3

Si CH3

CH3
O

Si

CH3
O

CH3

Si

Si

(CH2)2

(CH2)2

(CH2)2

C NH (CH2)6
O

7
372

304

1131

260

269

765

204

826

224

247

587

194

767

236

253

589

390

677

583

576

NH C
O

1012. Poly1H,1H-nonafluoro-4-oxahexyl acrylate)

CH2

6
220

CH3

1011. Poly[tri(oxyethylene)oxycarbonyliminohexamethyleneiminocarbonyl]

5
236

Si

CH3

CH3

4
1982

CH3

CH3
O

3
436

CH
C O CH2

(CF2)2

O CF2

CF3

O
1013. Poly1H,1H-nonafluoropentyl acrylate)

CH2

CH
C O CH2

(CF2)3

CF3

O
1014. Polyarylate of phenolphthalein and terephthalic acid

O
C
O
CO

O C

310

1
2
1010. Poly[tri(oxydimethylsilylene-1,4-phenyleneoxy)-1,4-phenylenedimethylsilylene]

1
2
1015. Polyamide of 1,7-bis(aminophenoxy)naphthalene and dicarboxylic acid

NH C

5
412536

6
537

7
717

389

745

522

483

443

711

643

623

396

655

588

605

1016. Polyamide of adipic acid and anilinephthalein

4
564

C NH

C (CH2)4

3
303

C NH
O

NH
C
O
CO

1017. Polyamide of anilinefluorene and terephthalic acid

NH
C

NH C

1018. Polyamide of anilinephthalein and isophthalic acid

C NH

NH
C
O
311

CO

NH

NH C

3
396

4
613

5
633

6
646

7
6

396

613

643

646

579

872

646

664

471

746

643-653

631

O
CO
1020. Polyamide of diphenylphthalidedicarboxylic acid and 1,4-phenylenediamine

C
O

C NH
C

NH

O
O

CO
1021. Polyamide of diphenylphthalidedicarboxylic acid and anilinephthalein

C
O

C NH
C

NH
C

CO

CO

1022. Polyamide of diphenylphthalidedicarboxylic acid and benzidine

C
O

C NH
C

O
O

CO

NH

312

1
2
1019. Polyamide of anilinphthalein and terephthalic acid

1
2
1023. Polyamide of phenolanthrone and terephthalic acid

NH

NH C

3
461

4
689

5
683

6
669

7
6

457

1018

435

449

480

829

600

579

C
O
1024. Polyamide of sebacic acid and anilinephthalein

C (CH2)8

C NH

NH
C
O
CO

1025. Polyamide of 4,4'-diphenyloxidedicarboxylic acid and anilinephthalein

C
O

C NH
O

NH
C
O
313

CO

C O

3
474

4
894

5
573593

6
530

7
6

333

722

457

461

683

474

894

585

530

516

940

543

549

O
C

O
CO
1027. Polyarylate of Bisphenol A and isophthalic acid

CH3
O

O C

CH3

1028. Polyarylate of diphenylphthalidedicarboxylic acid and 4,4'-diphenyleneoxide

C O
C

O
O

CO
1029. Polyarylate of diphenylphthalidedicarboxylic acid and bisphenol A

C
O

C
O
CO

CH3

C O

CH3

314

1
2
1026. Polyarylate of 4,4'-diphenyloxidedicarboxylic acid and phenolphthalein

1
2
1030. Polyarylate of diphenylphthalidedicarboxylic acid and resorcin
O

O C
O

3
390

4
720

5
543

6
542

7
6

455

754

627

604

513

911

587

563

437

817

543

535

C
C

O
O

CO

1031. Polyarylate of phenolanthrone and terephthalic acid

O
C

O C

C
O
1032. Polyarylate of phenolfluorene and 4,4'-diphenyldicarboxylic acid

O
C

O C

1033. Polyarylate of phenolfluorene and isophthalic acid

O
C

O C

O
315

O
C

O C

1035. Polyarylate of phenolphthalein and 4,4-diphenyldicarboxylic acid

O
C

O C

3
437

4
776

5
590

6
563

7
6

465

812

603

573

573

936

583598,
613

612

390

720

543

542

O
CO
1036. Polyarylate of phenolphthalein and diphenylphthalidedicarboxylic acid

O C
C

C
C

O
O

CO

CO

1037. Polyarylate of phenolphthalein and isophthalic acid

O
C
O
CO

O C

316

1
2
1034. Polyarylate of phenolfruorene and terephthalic acid

1
2
1038. Poly(carbonyl-1,4-phehylenesulfonyl-1,4-phenylenecarbonyloxy-1,4-SKHQ\OHQHIOXRUHQ_-9-ylidene-1,4phenyleneoxy)

SO2

C O

O C

5
623

6
580

7
6

383

861

433

445

452

1180

373

383

532

830

687-703

641

1039. Polyester of phenolphthalein and adipic acid

4
933

3
541

(CH2)4

C
O

O
CO
1040. Polyester of sebacic acid and phenolphthalein

C (CH2)8

C O

O
C
O
CO

1041. Polyimide of anilinefluorene and 3,3',4,4'-tetracarboxydiphenyl

N
C

CO

CO

CO

CO

317

3
467

4
840

5
586592

6
556

7
6

558

889

623673

628

550

893

623673

616

493

814

623664

606

CF3
N

CO

CO

CO

CF3

CO

1043. Polyimide of anilineanthrone and 3,3',4,4'-tetracarboxydiphenyloxide

CO

CO

CO
CO

N
C

CO
1044. Polyimide of anilinefluorene and 3,3',4,4'-benzophenonetetracarboxylic acid

CO

CO

CO

CO
CO

N
C

1045. Polyimide of anilinephthalein and 3,3',4,4'-tetracarboxydiphenyloxide

O
N

O
O

N
C
O
CO

318

1
2
1042. Polyimide of 4,4'-diaminodiphenyloxide and 3,3',4,4'-tetracarboxydiphenylhexafluoroisopropylene

3
229

4
398

5
563573

6
576

7
766

237

395

563573

600

766

456

619

723788

737

409

520

789

787

399

854

459

467

683

CO
O

CO

CO

1047. Polyparabanic acid 2

CO
CH2

CO

CO

1048. Polypyromelliteimide of anilinefluorene

CO

CO

CO

CO

N
C

1049. Polypyromelliteimide of anilinephthalein

CO

CO

CO

CO

N
C
O
CO

1050. Polysulfone of Bisphenol A and diphenylene sulfone

CH3
O

C
CH3

SO 2

319

1
1046. Polyparabanic acid 1

320

The above-described effect of influence of chain attachment on Tg of the


polymer is also spread upon polymeric systems, composed of isomers. For example,
polyamidophenyl quinoxalines, analyzed in the study in [83], display the following
structure:
O

NH C
O

C NH
O

PAPQ-1
O

C
NH O

C
O NH

PAPQ-2
In the case of PAPQ-2, polar groups representing phenolquinoxaline cycles are
in tight contact with each other, and the intermolecular interaction occurring at the
sacrifice of these groups is displayed in the frames of the unit; then, the dipoledipole
interaction between units of neighboring chains is not realized.
PAPQ-1 and PAPQ-2 are isomers with the same gross-formula.
Work [83] represent experiments and calculations on the determination of
glass transition temperatures of two pairs of polymers, the structure of which is shown
above. Since strong intermolecular interaction occurring at the sacrifice of separate
polar groups significantly affects Tg, it is obvious that when this interaction takes
place inside the unit, Tg of such polymers must be lower than for polymers with a
completely analogous structure, but with such disposition of polar groups in which
intermolecular interaction manifests itself between units of neighboring chains.
Calculations and experiments have confirmed that this indeed happens in this
way. When the glass transition temperature is calculated for PAPQ-1,
b j = 4bp + 2bh + 2bd + 2 1 bd , and for PAPQ-2 b j = 4bp + 2bh . Consequently,
2

the calculated value of Tg for PAPQ-1 is 265C, and for PAPQ-2 it is 208C.
Appropriate experimental values are 275 and 210C, respectively.
At the same gross-formula of the repeat unit, the glass transition temperature
also depends on the location of separate aromatic fragments. For example, polyimides
PI-1and PI-2 possess identical gross-formulae, but display different locations of
aromatic cycles. Calculating the glass transition temperature for PI-1, it is obtained
that b j = 2bd + 2bp , and in the case of PI-
b j = 2bd + 5bp . In accordance
j

with this, Tg for polyimide PI-1 is 238C, and for PI-


approximately the same values are obtained.

295C. Experimentally,

321

O
C

O
PI-1

PI-2
It should be noted that special attention must be paid to the cases when a
significant deviation in experimental and calculated data on the glass transition
temperature is observed. This may be caused by both the occurrence of a polar group
possessing a special influence on the energy of intermolecular interaction and, vice
versa, exclusion of any polar group from formation of the network of physical bonds
between neighboring chains of the polymer. In the first case, introduction of a new
constant bj is required, but it should always be remembered that the greater is the
number of constants introduced to the calculated scheme, the lower becomes its
predictive power and, in the border case, when every new polymer requires
introduction of a new constant bj, the predictive power of the calculation scheme
becomes equal to zero.
It should be also taken into account that anomalous addition of units to each
other may not weaken, but strengthen intermolecular interaction between neighboring
chains, and then the glass transition temperature will increase.
Let us now turn to the problem of plasticization rules. Let us use two
poly(ether ketones) as polymers:
C

PAEK 1

PAEK 2

O
CH3

H3C
C

O
O

C
O

C
O
As plasticizers, let us choose phthalic acid esters dimethylphthalate (DMP),
dimethylterephthalate (DMTP) and dibutylphthalate (DBP):

These poly(ether ketones) are produced in works [769 771].

322

O
C O CH3
O

O
C O CH3 ,

C O CH3
O

C O CH3

C O

(CH2)3

CH3

C O

(CH2)3

CH3

O
DMP

DMTP

DBP

Dependence of the glass transition temperature for PAEK 1 on the plasticizer


is shown below:
Glass transition temperature, K
Spl, % (mass)
DMP
DMTP
DBP
0
380
380
380
2.5
365
364
369
5.0
355
347
365
10.0
345
341
364
15.0
339
*
364
20.0
330
*
364
Dependence of the glass transition temperature for PAEK 2 on the plasticizer
concentration is the following:
Glass transition temperature, K
DMP
DMTP
0
470
470
2.5
425
417
5.0
408
394
10.0
387
380
15.0
362
*
20.0
353
*
* - dimethylterephthalate is crystallized out of the polymer.
Spl, % (mass)

DBP
470
441
426
415
406
403

IV.4. Influence of plasticization on the glass transition temperature of polymers


Let us analyze the changing effect of the part of the plasticizer introduced into
polymer on the glass transition temperature of the plasticized system. As mentioned
above, there are two fundamental concepts for description of these dependences and
two regularities corresponded to them: the Zhurkov rule of molar parts and the
KarginMalinskii rule of volumetric parts.
According to the Zhurkov rule, decrease of the glass transition temperature is
proportional to the molar part of the plasticizer introduced, i.e. when it is fulfilled, the
dependence of the glass transition temperature on the molar part of the plasticizer
must be linear. The experiments performed indicated that dependence Tg() is not
linear for all studied systems.

Authors data, not published.

323

According to the KarginMalinskii rule, decrease of the glass transition


temperature during plasticization is proportional to the volumetric part of the
plasticizer in the polymer (), i.e. Tg() must be linear. The experimental data for
systems considered give no linear dependence.
Hence, both rules do not describe the real behavior of the given polymer
plasticizer systems.
Let us consider a possibility to describe the dependence of the glass transition
temperature on the part of plasticizer, basing on the notions and calculation schemes,
discussed above. According to the given ideas, the glass transition temperature of a
linear polymer is described by correlation (IV.41).
If the plasticizer molecule interacts with polar groups of the polymer and
screens them, then the specific intermolecular interaction between polymer chains
cannot be displayed. Then, because these groups are excluded from the intermolecular
interaction, constants j must be excluded from equation (IV.41), and this exclusion
must be made in proportion to the molar part of the plasticizer introduced. The
quantity of parameters j excluded depends on the quantity of dipoledipole
interactions displayed by the given polar group of the polymer, as well as on the
number of polar groups contained in the plasticizer molecule. In the general form, we
obtain the following equation:

Tg =

Vi
i

Vi ai + j mbd n
i

(IV.45)

where m is the quantity of parameters bd related to a single polar group of the


polymer; n is the quantity of polar groups in the plasticizer molecule; is the molar
part of the plasticizer.
The value of m is determined according to the rules discussed above. In the
case of poly(arylether ketone) based on phenolphthalein (PAEK 2), m = 2.5; for diane
poly(arylether ketone) (PAEK 1), m = 2. With regard to the plasticizers applied,
dimethylphthalate and dibutylphthalate, for them n = 2. Then, equation (IV.45) is
reduced to the form:
for PAEK 2,

Tg =

Vi
i

Vi ai + j 2.5bd 2
i

(IV.46)

(IV.47)

and for PAEK 1,

Tg =

Vi
i

Vi ai + j 2bd 2
i

Calculations performed by equations (IV.46) and (IV.47) lead to a decrease of


the glass transition temperature of the plasticized sample compared with the original

324

polymer. However, the above-considered reason for decrease of the glass transition
temperature is not unique. The second reason may be trivial and associated with the
so-called effect of dilution. It is meant that a polymer with a relatively high softening
temperature is diluted by the plasticizer with a much lower softening temperature. If
the polymer and the plasticizer are miscible, the system may be considered analogous
to a two-component mixture of polymers, fully miscible with each other (the mixture
possesses general glass transition temperature at any composition). Then the phase
diagram (Tgcomposition) for the mixture of polymer and plasticizer may be
calculated by correlation (IV.59) applied to estimation of the glass transition
temperature of mixtures of miscible polymers. In this case, the correlation is reduced
to the form:

p Vi + pl Vi

i
p
i
pl

Tg =

Vi

i
p

(Tg )p

+ pl

Vi

i
pl

(Tg )pl

(IV.48)

+ 0.06 p pl

where Vi and Vi are Van-der-Waals volumes of the repeat unit of

i
p
i
pl
polymer and the plasticizer molecule, respectively; (Tg)p and (Tg)pl are the glass
transition temperature of the polymer and the softening temperature of plasticizer,
respectively.
Let us consider results of calculation (Figure 42) performed by equations
(IV.46), (IV.47) and (IV.48) for plasticized systems PAEK 1 and PAEK 2. Clearly,
the dependence of the glass transition obtained by equation (IV.48) with regard to the
dilution influence is much deeper than the one calculated by equations (IV.46) and
(IV.47) with regard to screening of polar groups of the polymer by plasticizer
molecules. However, both dependences give no values of depression of the glass
transition temperature, coincident to experimental ones. Hence, taking into account
each mechanism of plasticization separately does not enable adequate description of
the real behavior of the plasticized system. However, if the effects of dilution and
screening are summed up, the obtained dependence of the glass transition
temperature on the plasticizer concentration is close to the experimental one in the
area of plasticizer concentration, when it is miscible with the polymer. Calculated data
are shown below.
Calculated glass transition temperatures for plasticized PAEK 1
Glass transition temperature Tg, K
Cpf, % (mass)
DMP
DBP
Exp.
Calc. (dil. + scr.)
Exp.
Calc. (dil. + scr.)
2.5
365
367
368
366
5.0
355
355
365
352
10.0
345
333
364
327
15.0
339
313
364
305
20.0
330
296
364
285

325

Calculated values of glass transition temperature for plasticized PAEK 2


Glass transition temperature Tg, K
Cpf, % (mass)
DMP
DBP
Exp.
Calc. (dil. + scr.)
Exp.
Calc. (dil. + scr.)
2.5
425
448
441
447
5.0
408
428
426
425
10.0
384
393
415
387
15.0
362
363
406
353
20.0
253
336
403
323

Figure 42. Calculated dependence of glass transition of PAEK-11 (a, b) and PAEK-21 (c, d) on the
content of dimethylphthalate (a, c) and dibutylphthalate (b, d) .

Discussing the PAEK 2DBP system, it may be noted that starting from 7
10% (mass) of dibutylphthalate concentration in the system, clear separation,
displayed in appearance of film dimness, begins. From this point, decrease of the
glass transition temperature is decelerated and, consequently, an abrupt deviation
from the calculated dependence appears (Figure 42d). Using dimethylphthalate as a
plasticizer for PAEK 2, the plasticizer is much better joined with polymer. No signs of
separation are observed in the range of concentrations considered (020% (mass)). In
this case, the calculated dependence of the glass transition temperature on the
plasticizer concentration, obtained by summation of temperature depressions of
dilution and screening effects of polar groups, is close to the experimental one
throughout the length (Figure 42c).

326

Note that the present analysis requires no introduction of any fitting


parameters and is based only on the chemical structure of the plasticizer and the
polymer. It may be mentioned in passing that comparing the calculated and
experimental data, the system composition (the part of plasticizer), at which
miscibility fails (PAEK 2DBP, Figure 42e), may be determined,.
Concerning the case of PAEK 1 and PAEK 2 plasticization by
dimethylterephthalate, in the present case, such analysis cannot be performed
completely. This is associated with the circumstance that dimethylterephthalate is a
crystalline substance, and only the melting point is known for it. The dilution effect
may not be calculated, because, on the one hand, a mixture of crystalline and
amorphous substance exists and, on the other hand, although crystallinity of DMTP is
lost at low concentrations and its glass transition temperature is unknown. In this case,
the only depression of the glass transition temperature associated with screening of
polar groups in the polymer may be estimated. Clearly, the values of depressions
calculated by equations (IV.46), (IV.47) coincide with the analogous values calculated
for dimethylphthalate.
Hence, description of the plasticizer influence on the glass transition
temperature is quite complicated and, for adequate description, both mechanisms of
plasticization must be taken into account.
Let us now turn to description of copolymers.
For copolymers, the relation (IV.41) is transformed to the form:

1 Vi + 2 Vi + ... +

i
1
i
2

Tg =

1 ai Vi + b j + 2 ai Vi + b j + ... +

j
j
i
1
i
2

+ n Vi

i
n

+ n ai Vi + b j

j
i
n

(IV.49)

where 1, 2, , n are the molar parts of the components 1, 2, , n; Vi ,

i
1

Vi , , Vi are the Van-der-Waals volumes of the repeat units of the

i
n
i
2
components

1,

2,

n;

a V + b ,

i
i
j

j
i
1

a V + b ,

i
i
j

j
i
2

a V + b are selections of constants for the components 1, 2, , n

i
i
j

j
i
n
(remember that 1 + 2 + + n = 1).

327

Correlation (IV.49) always leads to non-linear dependences of Tg on the


polymer composition.
If we want to express Tg of the copolymer via the glass transition of
appropriate homopolymers, then using correlations (IV.41) and (IV.49), it may be
written

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n
Tg =
,

Vi
Vi
Vi

i
n
i
1
i
2
+ 2
+ ... + n
+
1
Tg,1
Tg,2
Tg, n

(IV.50)

where Tg,1, Tg,2, , Tg,n are glass transition temperatures of homopolymers, prepared
from the components 1, 2, , n.
In another form, the relation (IV.50) may be presented as follows:
1
1
1
1
=
1 +
2 + ... +
n ,
Tg Tg,1
Tg,2
Tg, n

(IV.51)

where

1 Vi

i
1
;
1 =

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n

1 Vi

i
2
;
2 =

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n

1 Vi

i
n
.
n =

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n
Hence, coefficients 1, 2, , n by their meaning are parts of the Van-derWaals volumes of the units 1, 2, , n.
In the reduced form, relation (IV.49) is presented in the form:

328

k =1

k Vi

Tg =

where

k =n

k ai Vi + b j
k =1
j
i
k

k =n

(IV.52)

k is the molar part of the k-th component;

Vi

i
k

and

a V + b are the Van-der-Waals volume and the selection of constants for

i
i
j

j
i
k
the k-th component, respectively; n is the total number of components in copolymer.
In the reduced form, correlations (IV.50) and (IV.51) are the following:
k =n

Tg =

k Vi

i
k
,

Vi

k =n
i

k T k
g, k
k =1
k =1

(IV.53)

and
1 k =n 1
=
k .
Tg k =1 Tg, k

(IV.54)

Dependences (IV.49)(IV.54) of the glass transition temperature Tg of the


copolymer on its composition are non-linear, and in the case of two-component
copolymers they look as it is schematically represented in Figure 43. These
dependences do not take into account weakening of the intermolecular interaction in
copolymer units, although this weakening must cause a decrease of the glass transition
temperature compared with the additive value.
In practice, dips on the dependences of Tg on composition are deeper than
the ones calculated from equations (IV.49)(IV.54) and equal T = 2030C [78].
That is why a modified relation was suggested in the work [39], which takes this
circumstance into account:

%J =

 ' +  ' + +

' + +  , ' +

 ,

+ +

, (IV.55)

329

+ Q 'L
L
Q

Q , L 'L + M + [ (  ) +  (  ) + + Q ( Q )] 

L
M

where all designations are the same as in the formula (IV.51); 0.03 is an empirical
coefficient which takes into account decrease of strong intermolecular interactions at
copolymerization.

Figure 43. Dependences of the glass transition temperature Tg on composition for two-component
copolymers calculated from equations (IV.49) (curve 1) and (IV.54) (curve 2)

In the reduced form, equation (IV.54) may be presented in the following form:

Tg =

k =n

k =1

k Vi

k =n

k ai Vi + b j + 0.03 k (1 k )
k =1
k =1
j
i
k

k =n

(IV.56)

where k, Vi and ai Vi + b j are the molar part, the Van-der-Waals

j
i
k
i
k
volume and the selection of constants for the k-th component of the copolymer,
respectively.

330

To calculate the glass transition temperatures of copolymers, equations (IV.55)


and (IV.56) require no knowledge of the experimental glass transition temperatures of
homopolymers.
Another variant of the formula for calculation of the glass transition
temperature of copolymers may be deduced by combining equations (IV.41) and
(IV.55):

1 Vi + 2 Vi + ... +

i
1
i
2

Tg =

Vi
Vi
Vi

i
n
i
1
i
2
+2
+ ... + n
+
1
Tg,1
Tg,2
Tg, n

(IV.57)

+ n Vi

i
n

,
+ [1 (1 1 ) + 2 (1 2 ) + ... + n (1 n )] 0.03
where Tg,1, Tg,2, , Tg,n are the glass transition temperatures of homopolymers,
composed of the components 1, 2, , n. In the reduced form, equation (IV.57) can be
written as:
k =n

Tg =

k Vi

i
k
.

Vi

k =n
k =n
i
k
+
0
.
03

k T
k (1 k )
g, k
k =1
k =1
k =1

(IV.58)

Let us now discuss the influence of sequences distribution in polymers on the


glass transition temperature of them.
Imagine that the following sequences exist:
1) Diades AA and BB;
2) Triades AAA and BBB;
where A and B are the components of a copolymer. Let the molar parts of components
A and B be 0.5:0.5.
Now Tg values may be calculated for random copolymers consisting of the
sequences mentioned above.
For the random copolymer consisting of A and B components, we obtain:

A Vi + B Vi

i
A
i
B
Tg =
, (IV.59)

Vi
Vi

i
A
i
B
+B
+ 0.03[ A (1 A ) + B (1 B )]
A
Tg
Tg

( )A

( )B

331

where A and B are the molar parts of the components A and B, respectively;

V and V are the Van-der-Waals volumes of the components A and


i
i

i
A
i
B
B, respectively; Tg,A and Tg,B are the calculated glass transition temperatures for
homopolymers based on the components A and B, respectively.
For instance, for copolymers of ethylene (component A) and vinyl acetate

(component B), it is obtained: Tg,A = 213 K, Tg,B = 301 K; Vi = 34.2 3,

i
A

V = 79.65 3:
i

i
B
(CH2CH2)n

CH2

CH

)m

C O
CH3
B
Then, Tg value for the random copolymer in accordance with equation (IV.59)
is equal to 250 K, when A:B = 0.5:0.5.
Let us now examine the case, when the copolymer composed of the same
components is the random copolymer consisting of diades AA and BB. Since Tg,AA =
Tg,A and Tg,BB = Tg,B, equation (IV.59) is reduced to the following form:

A 2 Vi + B 2 Vi

i
A
i
B
Tg =
. (IV.60)

2 Vi
2 Vi

i
A
i
B
+ B
+ 0.03[ A (1 A ) + B (1 B )]
A
Tg
Tg

( )A

( )B

(It should be remembered since the molar ratio of the components A and B is 0.5:0.5,
the molar ratio of the components AA and BB is the same). Substituting all the values
into equation (IV.60), we obtain Tg = 259 K.
Let us now discuss the case when the random copolymer consists of the triades
AAA and BBB. Then, in accordance with equation (IV.59), we get:

A 3 Vi + B 3 Vi

i
A
i
B
Tg =
. (IV.61)

3 Vi
3 Vi

i
A
i
B
+B
+ 0.03[ A (1 A ) + B (1 B )]
A
Tg
Tg

( )A

( )B

332

Substituting all the parameters into equation (IV.61), we obtain that Tg = 262 K.
Of course, in principle, any copolymer may be constructed of arbitrary
sequences. For example, they may be the following: AAA, ABBA, BBB, ABAB,
BBBA, etc. In this case, the equation for calculating the glass transition temperature is
reduced to the form:

+ ABBA Vi
+ BBB Vi
+
AAA Vi

i
AAA
i
ABBA
i
BBB

Tg =

Vi
Vi
Vi

i
AAA
i
ABBA
i
BBB
+
+ ABBA
+ BBB
AAA
Tg
Tg
Tg
(IV.62)
AAA
ABBA
BBB

+ ABAB Vi
+ AAAB Vi
+ BBBA Vi

i
ABAB
i
AAAB
i
BBBA

Vi
Vi
Vi

i
ABAB
i
AAAB
i
BBBA
+ AAAB
+ BBBA
ABAB
Tg
Tg
Tg

( )

( )

( )ABAB

( )AAAB

( )

( )BBBA

where AAA, ABBA, BBB, ABAB, AAAB, and BBBA are the molar parts of the

,
sequences AAA, ABBA, BBB, ABAB, AAAB, and BBBA; Vi

i
AAA

Vi
Vi
Vi
Vi
,
,
,
,
and

i
BBB
i
ABBA
i
ABAB
i
AAAB

Vi
are the Van-der-Waals volumes of the same sequences, respectively;

i
BBBA
Tg,AAA, Tg,ABBA, Tg,BBB, Tg,ABAB, Tg,AAAB, and Tg,BBBA are the glass transition
temperatures of homopolymers based on the components AAA, ABBA, BBB, ABAB,
AAAB, and BBBA.
Obviously, the molar parts of all the sequences cannot be arbitrary, because
the molar parts of the original components A and B are given. Consequently, the way
of looking at the problem of influence of the sequences distribution is the following:
1) Using the above approach, all the sequences are drawn as homopolymers and Tg
values are calculated for them.
2) The properties of the copolymer containing the given parts of the sequences are
calculated; it should be only remembered that original components are taken in the
given proportions, and the molar parts of the sequences must correspond to the
molar parts of original components.
Let us now consider the influence of chain branchings on the glass transition
temperature.
Branchings of chains are observed quite often and are either a consequence of
side processes proceeding in the course of synthesis of linear polymers, or are created
purposefully for modification of properties. For example, branches may already be
contained in the structure of monomers of the following type:

333

R'
CH2

CH;
R

CH2

C
R

where R is the hydrocarbon radical: CnH2n+1; OCnH2n+1; C O CnH2n +1 ,


etc.
O
In the general case, the number of possible variants of chain branchings may
be quite high. Consider the main among them, schematically represented in Figure 44,
which depicts the backbone and possible types of branches.

Figure 44. Different types of branchings in polymers (see text).

In the simplest (but of the lowest probability) case, branches may possess
equal length and be located in each repeat unit (see Figure 44a). Another variant
corresponds to branches of the same length, regularly distributed along the backbone
in a definite number of units (see Figure 44b). The third type corresponds to the case,
when branches of various length exist, but they are regularly distributed along the
backbone in a definite number of units (see Figure 44c). Then a variant is possible
when branches of different lengths are randomly distributed along the backbone (see
Figure 44d). Two more types are to be considered. According to one of them,
branches of different lengths are randomly distributed along the backbone, and each
branch possesses one more branch of different length on it (see Figure 44e). Clearly,
this general case may include all particular cases, when branches are identical and
regularly distributed along the backbone, and secondary branches are identical or
different, etc. One more principal case consists of a tree-like branching (see Figure
44f). In this case, occurrence of every new branch causes occurrence of one more
branch.
Note that at the present time, such systems have been synthesized in practice
[82], and they are named dendric polymers or dendrimers.
In all cases considered, branches may be of the same chemical origin in
relation to the backbone, or of a different one. The case, when branches are of a
different chemical origin in relation to the backbone corresponds to grafted polymers.

334

In these systems, in most cases, a microphase separation occurs due to


thermodynamic inconsistency of the backbone and the grafted chain. Therefore, each
phase may possess, for example, its own glass transition temperature which, however,
in many cases differs from the glass transition temperature of individual components.
That is why Tg for grafted copolymers may be calculated only in the case of complete
miscibility of the initial and the grafted polymer. However, the reverse task may also
be solved estimation of the composition of each phase by temperatures of relaxation
transitions in it, assuming that each of the phases includes a definite number of
foreign units.
Work [24] analyzes the influence of branchings of different types on the
physical characteristics of polymers.
In homopolymers, branches may be joined directly to the backbone by
substitution of one of the atoms or via a spacer of another chemical origin. Let us
consider all these variants on particular examples. Let us choose polyethylene and its
derivatives as one of the model polymers for this consideration:
Structure I
CH2

CH

(CH2

CH2)m

(CH2)n
CH3
Let us consider the influence of branchings on the glass transition temperature
Tg. Note that if m = 0, we get the case, depicted in Figure 44a, when branches are of
the same length and distributed on each repeat unit. Let us discuss in detail calculation
of this structure. The glass transition temperature is calculated from equation (IV.41).
For the structure I, depicted above,

Vi = (51.3 + 17.1n + 34.2m)3;


i

ai Vi + b j = (185.34 + 80.25n + 160m)103 3K1.


i

Calculating values of

b j , it should be taken into account that every branch


j

requires introduction of a single increment bd. As a consequence, we get


Tg (K ) =

51.3 + 17.1n + 34.2m


103 .
185.34 + 80.25n + 160.5m

(IV.63)

The results of calculation performed in equation (IV.63), are shown in Figure


45. Clearly, if branches locate in every repeat unit (m = 0), then Tg is strongly
dependent on the number of CH2groups in the branch. Tg sharply increases as n
decreases, when n < 5.
For the case m = 0, a series of branched polymers exists, the properties of
which have been studied experimentally. For example, when m = 0 and n = 0,
polypropylene is obtained, for which the calculated value of Tg is 277 K, and the
experimental value is 263 K. If m = 0 and n = 1, we get polybutene-1 with the
calculated value of Tg of 258 K and the experimental one of 248 K. The same

335

convergence typical of the present method is also observed for other polymers at m =
0 and n = 2 and 6.

Figure 45. Dependences of the glass transition temperature Tg on n for structure I. Numbers at curves
denote m values (see text).

If branches are less frequent (m > 0), influence of the number of units in the
branching on Tg decreases and, when branches become rather rare (m = 10), the
branching length does not practically affect Tg (note, by the way, that when m ,
this correlates with a star-like polymer).
Let us now consider the case when branches in polyethylene are randomly
distributed along the backbone, these branches being of different lengths. Assume that
the value of m adheres to the random distribution
F ( m) = 1 e

m mavg

( m 0) ,

(IV.64)

where mavg is the average value of m.


The function of density distribution is of the following form
f ( m) =

1
mavg

m mavg
.

(IV.65)

Assume also that the polymerization degree of branches is distributed by the


Flory law
q(n) = 2nen,

(IV.66)

1
; nw is the weighted mean value of n.
nw
For future calculations assume that the border value n = 10. Then the number
of units in a branch will be determined from the correlation

where =

nbr =

i = nk

i =1

i 2 2 e i .

(IV.67)

336

The formula for calculation of the glass transition temperature Tg deduced on


the basis of equation (IV.41) and with regard to all mentioned above, obtains the
following form for structure I
i = nk

51.3 + 17.1 i 2 2 e i + 34.2


i =1
i = nk

Tg (K ) =

185.3 + 80.25

1
mavg

j 2 2 e j

j =1
j = mk

i 2 2e i + 160.5

i =1

where =

j = mk

103 ,

(IV.68)

je j

j =1

Results of the calculation performed by formula (IV.68) are indicated in


Figure 46. The character of Tg dependence on nw is analogous to Tg dependence on n
with the only difference that the former are more smooth.
Generally, results of the calculations performed [24] for branched
polyethylene indicate that the greatest influence on the glass transition temperature is
cause by short-chain branchings, frequently located along the backbone.

Figure 46. Dependences of Tg on nw for structure I. Numbers at curves denote mavg values (see text).

Let us now turn to the next type of branchings when new branches appear on
the original branch. In this case, the structural formula of branched polyethylene is the
following
Structure II
CH2

CH

(CH2

CH2)n

(CH2)x
CH

(CH2)y CH3

(CH2)z
CH3
This correlates with the scheme in Figure 44e. In this case, the formula for
calculation of Tg, deduced from equation (IV.41), obtains the following form:

337

7J . =

 +  [ + \ + ] + P


 .
 +  [ + \ + ] + P 

(IV.69)

The meaning of designations x, y, z, and m is seen from the structural formula


of branched polyethylene. Note that if x + y + z = 0 and m = 0, this corresponds to a
polymer of the following structure
... CH2 CH ...
CH CH3
CH3
For this polymer, the calculated glass transition temperature is 294 K, and
experimental 302 K.
Analysis of equation (IV.69) indicates that Tg depends only on the sum
(x + y + z) and the value m. Figure 47a shows dependences of Tg on (x + y + z) at
various values of m, i.e. at various frequencies of branches, determined by the value
1/m. Here the influence of branches is approximately the same as in the previous
cases. However, the appearance of secondary branches causes a sharper increase of
the glass transition temperature, which is most clearly displayed at short and frequent
(m = 0) branchings.
Let us consider tree-like branchings. In the case of polyethylene with a trifunctional cross-linked point, the chemical structure of such system is shown below:
Structure III
...
... CH2 CH ...
(CH2 )n
(CH2)n

CH

(CH2 )n
...

CH
..
.

(CH2)n

CH2
CH

(CH2)n

(CH2 )n

CH
..
.

CH

(CH2)n ...
(CH2)n CH ...
..
.
CH

(CH2 )n
...

When determining

b j ,

it should be taken into account that every new

branch causes a necessity of introducing the increment bd, and its influence becomes
most noticeable when the shortest branches exist. If these branches are of the same
length (n =const), the formula for calculation of the glass transition temperature,
deduced from equation (IV.31), obtains the following form:
Tg (K ) =

17.1n + 11.0
103 .
80.25n 15.25

(IV.70)

The dependence of Tg on n, calculated by equation (IV.70), is indicated in


Figure 47b (curve 1). It is clearly observed that if tree-like branchings exist, Tg of the
system changes must more significantly than in all previous cases. When branches
become shortest (n = 1), Tg reaches 160C. This is much higher than Tg of the original
polyethylene. The reason for such a sharp increase of Tg is that at short branchings the

338

chemical structure of the polymer significantly differs from the chemical structure of
polyethylene, the role of dipoledipole interaction increasing.
Consider another variant, when branches are distributed by lengths according
to the Flory law. In this case, the formula for calculation of Tg deduced from (IV.41)
obtains the following form:
i = nk

Tg (K ) =

17.1 i 2 2e i + 110
i =1
i = nk

80.25

2 2 i

103 ,

(IV.71)

15.25

i =1

and the results of calculations performed by equation (IV.71) are shown in Figure 47b
(curve 2). It may be noted that when branches are distributed by lengths, dependence
of Tg on nw is weaker than with branches of the same length. Clearly, comparison is
set at n = nw .
Let us now discuss branchings of the same chemical origin as the backbone,
but joined to it by untying of another chemical structure. As an example, let us
consider the structure of the branched polymer, displayed below:
Structure IV
... CH2 CH (CH2 CH2)m ...
O
CH2
(CH2 )n
CH3
A particular case of this system (m = 0) represents a selection of vinyl ethers
with different values of n. For this case, the formula for calculation of Tg is of the
following form
Tg (K ) =

78.1 + 17.1n
103
289.4 + 80.25n

(IV.72)

Figure 47. Dependences of Tg on (x + y + z) for structure II (a) and dependences of Tg on n (1) and
nw (2) for structure III (b). Numbers at curves denote m values (see text) .

339

and calculations by equation (IV.72) lead to the dependence of Tg on n depicted in


Figure 48. This dependence is analogous to the above-described ones. Therewith,
concordance of the calculated and experimental values of Tg usual for the present
method is also observed in this case.

Figure 48. Dependence of Tg on n for structure IV.

If branches are distributed by lengths and randomly located along the


backbone, the formula for calculation of Tg transforms to:
i = nk

78.1 + 17.1 i 2 2 e i + 34.2


Tg (K ) =

i =1
i = nk

289.4 + 80.25

i =1

2 2 i

i e

j = mk

je j

j =1
j = mk

+ 160.5

103 .
je

(IV.73)

j =1

In the general case, the dependence of Tg on the average number of CH2


groups in branches at various values of m looks as depicted in Figure 49. The
character of these dependences is also analogous to that considered above for
branched polyethylene.

Figure 49. Dependences of Tg on nw for structure IV. Numbers at curves denote m values (see text).

340

In conclusion, let us consider a series of organic glasses based on


polymethacrylates. The general formula of these systems is of the following form:
Structure V
CH3
...

CH2

...

C O CH2

(CH2)n CH3

O
The formula for calculation of Tg for polymethacrylates, deduced from
equation (IV.41), is of the form
Tg (K ) =

113.85 + 17.1n
103 .
355.0 + 80.25n

The dependence of Tg on the length of side branch n is shown in Figure 50. As


the branch length increases, the glass transition temperature abruptly decreases and
transition from the glassy state to the rubbery state is observed.

Figure 50. Dependences of Tg on n for structure V.

In summary, several conclusions on the influence of branchings on the glass


transition temperature Tg can be made.
The first conclusion is that as the length of branches increases, the properties
of polymers change insignificantly and approach the properties of homopolymers of
the chemical structure identical to branches. This is true in the case when branches
located in each repeat unit. If the branches are less frequent and distributed along the
backbone of the polymer is a random way, their influence on properties is more
significant, if their chemical nature differs from that of the backbone.
The highest influence on the physical properties is exerted by short
branchings. For the glass transition temperature, the influence of short branchings is
formally taken into account by introduction of constants bd to the formula (IV.41) in
the case of every new branching. This influence is associated with increase of chain
rigidity and occurrence of an additional intermolecular interaction, when a branch
containing a polar group is joined. It should also be noted that existence of a great
number of short-chain branches brings about a qualitative change of the chemical
structure of the polymer. If short-chain branchings are of the tree-like form, the

341

chemical structure of the system obtained is just roughly similar to that of the initial
polymer.
In conclusion of this Section, let us consider a possibility of estimating
energies of the intermolecular interaction and its components, stipulated by various
types of intermolecular interaction. Basing on the approach developed in refs. [6,
128], we obtain that the total energy of intermolecular interaction <D> is described by
the following correlation:

D =

R Vi
i

0.455 ai Vi + b j

j
i

m,

(IV.74)

where m is the number of atoms in the polymer repeat unit; R is the universal gas
constant; the meaning of the rest of parameters is the same as in equation (IV.41).
For copolymers, we obtain:
k =n

R k Vi

k =n
k =1
i
k
D =
k mk ,
k =n

k =1
0.455 k ai Vi + b j

k =1
j
i
k

(IV.75)

where mk is the number of atoms in the k-th repeat unit.


Substituting equation (IV.41) into formula (IV.75), we obtain:
D =

RTg
0.455

m,

(IV.76)

and for copolymers,


D =

RTg, cop k = n
k mk ,
0.455 k =1

(IV.77)

where Tg,cop is the glass transition temperature of the copolymer.


Energy of dispersion interaction <D>d is determined from the following
expression:

D d=

R Vi
i

0.455 ai Vi

For a statistic copolymer, we obtain:

m.

(IV.78)

342

k =n

R k Vi

k =n
k =1 i
k
Dd=
k mk .
k =n
k =1

0.455 k ai Vi

k =1 i
k

(IV.79)

The energy of the dipoledipole and hydrogen bonds is determined from the
formula:

mR Vi

1
1
i

D dd + h = D D d =
.

0.455 ai Vi + b j ai Vi
j
i

(IV.80)

The ratio of the dispersion interaction in the total intermolecular interaction


equals:

b j
D d
j
= 1+
.
D
a
i Vi

(IV.81)

For statistic copolymers, we obtain:


k =n

k b j

D d
k =1 j
k
= 1+
.
k =n
D

k ai Vi
k =1 i
k

(IV.82)

The ratio of dipoledipole and H-bonding interaction in total intermolecular


interaction may be determined from the following expression:

ai Vi + b j

D dd + h
Dd
i
= 1
=1
D
D

ai Vi
i

b j
=

ai Vi

(IV.83)

which for statistic polymers is reduced to the form:


k =n

( )

k b j k
D dd + h
k
= =1
.
k =n
D
k (ai V )k
k =1

(IV.84)

343

Hence, using the above-described approach, not only the glass transition temperature
of polymers and copolymers may be estimated, but also such important characteristics
as total energy of intermolecular interaction and its components, stipulated by various
types of the intermolecular interaction (weak dispersion interaction, strong dipole
dipole interaction, and hydrogen bonds), may be determined. Running ahead, let us
note that such estimation is also correct for polymer networks, which are analyzed
below.

IV.5. Calculation of the glass transition temperature of polymer networks


It is experimentally determined that the glass transition temperature, as well as
many other properties of network systems, depends on the number of the repeat units
between cross-linked points m in the way as schematically represented in Figure 51. If
the network is not dense, its glass transition temperature depends only slightly on m,
but when the number of the repeating units in chains between neighboring crosslinked points decreases significantly, the glass transition temperature increases
sharply and reaches extremely high values. Tests and calculations indicate that the
glass transition temperature begins increasing, when the number of units in linear
fragments binding cross-linked points becomes smaller than 45. This is the vary
value of m, at which it may be said about the transition from rarefied to frequent
networks.

Figure 51. Schematic representation of dependence of the glass transition temperature Tg of networks
on the number of repeat units m in linear fragments between cross-linked points.

Let us consider two border cases:


1) Network is extremely sparse; this correlates with the case of m 0. In this case,
the properties of networks do not practically differ from the properties of a linear
polymer of the same chemical structure. Consequently, the glass transition
temperature Tg may be calculated from equation (IV.41).
2) Network is extremely frequent and consists of cross-linked points only (the notion
of the network cross-linked point will be discussed in detail below). In this
border case, the network may soften at heating only at the sacrifice of its thermal
degradation by chemical bonds in cross-linked points. The softening temperature
of this network, or more precisely the temperature of onset of intensive thermal
degradation, may be calculated from equation (VII.5).

344

In most cases, real networks respond to the intermediate variant, because they
contain quite long but finite linear fragments between cross-linked points. Calculation
of the influence of both these linear fragments and network cross-linked points on the
glass transition temperature has led to the following equation, which represents the
generalized form of equations (IV.41) and (VII.5).

7J =

UI
,

+ E + . 9

FUOS
OFK
L

D 9

(IV.85)

where 9 is the Van-der-Waals volume of the repeating fragment of the

UI

is a selection of constants for linear chains


network; ai Vi + b j

i
j

l.ch.
L

participating in the repeating fragment of the network; . 9


is a selection

FUOS
of constants for the network cross-linked point.
This equation takes into account both border cases of extremely rarefied and
extremely frequent networks, and intermediate cases. Actually, if a network consists
of linear fragments only, influence of cross-linked points may be neglected and

. 9
0; equation (IV.85) transforms into (IV.41). On the contrary, in a

FUOS
network is the most frequent, i.e. consists of cross-linked points only, equation
(IV.75) transits into (VII.5). In all intermediate cases, both summands from the
denominator of equation (IV.85) act.
L

Figure 52. Schematic representation of networks of different types: (a) network consisting of
fragments of polymer chains of the same chemical structure and possessing tetra-functional
cross-linked point; (b) network consisting of fragments of polymer chains of the same
chemical structure but possessing tri-functional cross-linked point; (c) network consisting
of linear chains of the same chemical structure and cross-links of different chemical
structure possessing tri-functional cross-linked point.

Let us consider now two the most important questions. Question one is
connected with estimation of the Van-der-Waals volume of the repeating fragment of

345

the network. Let us analyze several variants of networks, schematically represented in


Figure 52. The network depicted in Figure 52a possesses a tetra-functional crosslinked point (four chains yield from each cross-linked point, all chains possessing the
same chemical structure). For this network, the Van-der-Waals volume of the
repeating fragment of the network is calculated as
*

Vi = 2(m 2) Vi
+ 4 Vi
+ Vi
,

i
r.f.
i
l.ch.
i
l.ch. i
crl.p.

(IV.86)

where Vi
is the Van-der-Waals volume of the repeat unit in linear chains

i
l.ch.

is the Van-der-Waals volume of the


connecting cross-linked points; Vi

i
crl.p.
*

network cross-linked point; Vi


is the Van-der-Waals volume of the unit of

i
l.ch.
the linear fragment, linked to the cross-linked point (this unit is of somewhat different
chemical structure compared with normal units).
Another variant reflects the case when the network possesses a tri-functional
cross-linked point which links linear chains of the same chemical structure (see Figure
52b). For this network, the formula for calculation of the Van-der-Waals volume of
the repeating fragment obtains the form:
*

Vi = (m 2) 3 Vi
+ 3 Vi
+ Vi
,

2 i
i
r.f.
l.ch. i
l.ch. i
crl.p.

(IV.87)

where designations are the same as in equation (IV.86).


One more type of the network is depicted in Figure 52c. This network is
composed of cross-linked chains of the same chemical nature and cross-linking
bridges of different chemical nature. In this case, the formula for calculation of the
Van-der-Waals volume of the repeating fragment of the network obtains the form:
*

Vi = (m1 2) Vi
+ 2 Vi
+

i
r.f.
i
l.ch.,1 i
l.ch.,1
*

m 2
+ 2
Vi
+ Vi
+ Vi

2 i
l.ch.,2 i
l.ch.,2 i
crl.p.

(IV.88)

346

where Vi
is the Van-der-Waals volume of the repeat unit of linear

i
l.ch.,1
fragments of cross-linked chains, m1 being the average amount of these units, located
*

is the same for the unit


between cross-linked points of the network; Vi

i
l.ch.,1
linked to cross-linked point (possessing different chemical structure compared with

is the Van-der-Waals volume of the repeat unit of the


normal units); Vi

i
l.ch.,2
linear fragment of cross-linking bridges, m2 being the amount of these units;
*

Vi
is the same for the unit linked to the cross-linked point.

i
l.ch.,2
Let us now discuss the notion of cross-linked point of the network. To
formulate this notion, equation (IV.85) was calibrated in work [30] basing on the
experimental data on the glass transition temperatures for numerous and wellcharacterized network systems. Consequently, it was obtained that the following
definition of the cross-linked point must be assumed for better coincidence of
calculated and experimental data: the cross-linked point of a network is a group of
atoms which includes the atom, from which branching of chains takes place, and
neighboring atoms, chemically bonded with it, with their closest substituents. Shown
below are networks of different chemical nature, in which cross-linked points,
according to this definition, are marked by the dotted line. Basing on the present
definition of the cross-linked point and the examples displayed, it is quite easy to
identify the cross-linked point in a network of any chemical nature.
..
.
CH2
CH2
...

CH2

CH2

CH CH2

CH2

... ...

CH2

CH2

C CH2

CH2

CH2

CH2

CH2

...

...

...

CH2

CH CH2

CH CH2

CH CH2

CH

...

...

CH2

CH CH2

CH CH2

CH CH2

CH

...

CH2

...

347

OH
CH2

...
N

CH2

...

OH

CH CH2

O CH2

CH CH2

...

CH CH2

O CH2

CH CH2

...

N
CH2

OH

OH

...
CH3
Si

CH3

CH3
...

Si

...

CH3
CH3

Si
O
Si

CH3
O

Si

...

CH2

CH3

CH2

...

CH3
CH3

O
..
.

..
.

Analyzed below are various variants of networks and examples of calculations


of their glass transition temperatures Tg are given. Let us now dwell on estimation of
value of the molecular mass of the averaged fragment between neighboring crosslinked points. This estimation may be performed on the basis of the chemical structure
of the network and the experimentally determined glass transition temperature. For
this purpose, the experimental value of Tg must be substituted to equation (IV.85), and
the equation must be solved by m. Let us perform this analysis in general for various
types of networks depicted in Figure 52.
Let consider first the network depicted in Figure 52a; four chains yield from
each cross-linked points of this network. Therewith, all linear fragments are of the
same chemical structure of the repeat unit. Then, considering the repeating fragment
of the network, it should be written on the basis of equations (IV.85) and (IV.86):
*

Tg =

+ 4 Vi
+ Vi
2(m 2) Vi

i
l.ch. i
l.ch. i
crl.p.

, (IV.89)

+ 4 ai Vi + b j
+ K i Vi
2(m 2) ai Vi + b j

j
j
crl.p.
i
l.ch. i
l.ch. i

is a selection of constants for the repeat unit of linear


where ai Vi + b j

j
i
l.ch.
*

intercross-linked chains; ai Vi + b j
is the same for border chains linked

j
i
l.ch.

348

to cross-linked points; K i Vi
is a selection of constants for the network

i
crl.p.
cross-linked point; other designations are the same as in formula (IV.86).
The molecular mass of linear chains between cross-linked points Mc may be
determined as
M c = ( m 2) M + 2M * + 1 M crl.p. ,

(IV.90)

where M is the molecular mass of the normal unit; M* is the molecular mass of the
unit linked to the cross-linked point; Mcrl.p. is the molecular mass of the cross-linked
point.
Denote:

Vi
= A;

i
l.ch.

+
2 Vi

i
l.ch.

1
2

=C;
= B ; ai Vi + b j

i
j
crl.p.

l.ch.

Vi

i
*

+
2 ai Vi + b j

i
j

l.ch.

1
2

K i Vi

crl.p.

= D.

(IV.91)

Then, after some transformations, equations (IV.89) and (IV.90) obtain the
form:
m2=
Mc =

B Tg D
Tg C A

B Tg D

Tg C A

(IV.92)

M + 2M * + 1 M crl.p. .

(IV.93)

Let us now analyze a network with tri-functional cross-linked points with the
same chemical structure of chain units yielding from the cross-linked point (see
Figure 52b). For the repeating fragment of this network, it should be written:

P  9


7J =


P  D 9 + E

L

+  9
OFK

+ 9
OFK

+  D 9 + E

OFK
L

FKOS

, (IV.94)

+ . 9

OFK
L

FKOS

where all designations are the same as in equation (IV.89).


Then

 %J 
;
 %J 

2  = 

(IV.95)

349

where
*

; D1 = ai Vi + b j
. (IV.96)
B1 = Vi
+ Vi
+ K i Vi

i
j
i
i
l.ch. i
crl.p.

l.ch.
crl.p.

Therefrom,
B1 Tg D1
M + 2 M * + 2 M crl.p. .
M c = ( m 2) M + 2M * + 2 M crl.p. = 2
3
3 T CA
3
g

(IV.97)

The third variant of the network depicted in Figure 52c consists of linear
chains, cross-linked by chains of a different chemical structure; the cross-linked point
of the network is tri-functional. For this network, the dependence of Tg on the
composition is of the following form:
*

Tg =

(m1 2) Vi
+
+ 2 Vi
+ 2 Vi
3

i
l.ch.,1 i
l.ch.,1
i
crl.p.
*

(m2 2) ai Vi + b j
+ 2 ai Vi + b j
+ K i Vi
+

j
j
crl.p.
i
l.ch.,1 i
l.ch.,1 i

m 2
+ 2
Vi

2 i

, (IV.98)

+ Vi
+ 1 Vi
3

l.ch.,2
i
crl.p.
l.ch.,2 i
*

m 2
ai Vi + b j
+ 2
+ ai Vi + b j

2 i
j
j

l.ch.,2 i
l.ch.,2

where m1 and m2 are the numbers of normal units of cross-linked chains and crosslinking bridges between neighboring cross-linked points. All parameters with index *
relate to chains linked to the cross-linked points. The relation between m1 and m2 is
detected from the molar ratio of components 1 and 2:
m1/m2 = q1/q2,
where q1 and q2 are numbers of moles of components 1 and 2, respectively.
Substituting this relation into equation (IV.98), we get

m1 2 =

B2 2 A1 A2 Tg ( D2 2C C2 )

q
Tg C1 + 1 C2
q
2

q
q
2Tg C1 1 C2 + 2 A1 + 1 A2
2 q2
2 q2

q1
A2
A1 +
2q2

(IV.99)

350

where

A1 = Vi
; A2 = Vi
;

i
l.ch.,1
i
l.ch.,2





=  '
+ '
+ '
+ '
;

OFK 
FUOS
OFK 
FUOS
L

C1 = ai Vi + b j
; C2 = ai Vi + b j
;

i
j
i
j

l.ch.,1

l.ch.,2
*

+ K i Vi
+ ai Vi + b j
D2 = 2 ai Vi + b j
.

j
j
crl.p. i
i
l.ch.,1 i
l.ch.,2
Let us introduce designations:

q
B* = B2 2 A1 A2 + 2 A1 + 1 A2 ;
q2

q
q
D* = D 2 C 2 + 1 C2 ; C* = C1 + 1 C 2 .
q2
q2
Then we obtain
B * Tg D *
.
m1 2 =
Tg C * A *
q
A* = C1 + 1 A2 ;
q2

(IV.100)

(IV.101)

In a network composed of heterogeneous chain fragments, located between


cross-linked points, it is advisable to determine two values of Mc (Mc,1 which is the
molecular mass of the chain fragment between cross-linked points for cross-linked
chains, and Mc,2 which is the same for cross-linking bridges). Therewith,
Mc,1 = (m1 2)M1 + 2M1* + Mcrl.p.,1;
q

Mc,2 = (m2 2)M2 + 2M2* + Mcrl.p.,2 = 2 m1 2 + 2M2* + Mcrl.p.,2,


q1

(IV.102)
(IV.103)

where Mcrl.p.,1 is the molecular mass of the part of the cross-linked point adhered to
cross-linked chains (black circles in Figure 52c); Mcrl.p.,2 is the molecular mass of the
part of the cross-linked point adhered to cross-linking bridges (white circles in Figure
52c). Therewith, the value Mcrl.p.,1 + Mcrl.p.,2 is equal to the molecular mass of the
whole cross-linked point, Mcrl.p.. Then, considering (IV.101), expressions (IV.102) and
(IV.103) become of the following form
M c,1 =
q
M c,2 = 1
q2

B * Tg D *
Tg C * A *

M 1 + 2 M 1* + M crl.p.,1;

B * Tg D *

M + 2 M 2* M 2 + M crl.p.,2 .
Tg C * A * 2

(IV.104)

(IV.105)

351

The average value of molecular mass Mavg of the chain between network
cross-linked points may be calculated from the equation
Mavg = Mc,1 + (1 )Mc,2,

(IV.106)

where = q1/(q1 + q2).


Let us now discuss in detail the procedure of calculating Tg and Mc for all
three variants. The first of them correlated with a network with a tetra-functional
cross-linked point and chains of the same chemical structure yielding from it. As the
simplest example, let us analyze the hypothetical network based on polyethylene

Structure VI
...
...

(CH2)m- 2

...
...

CH2

(CH2)m -2

CH2

...

CH2

(CH2)m-2

CH2

(CH2 )m- 2
(CH2)m -2

CH2

CH2

...

..
.

CH2
CH2

C
...

CH2

...
C

...

CH2

..
.

...

...

(CH2)m -2

CH2

...

..
.

CH2
..
.

For this network, the Van-der-Waals volume of the repeating fragment will be
determined from the relation

Vi = Vi
+ 2 Vi ( m 2) ,

i
r.f. i
crl.p. i
v.

(IV.107)

where Vi
is the Van-der-Waals volume of the cross-linked point (marked

i
crl.p.

by dotted lines); Vi is the Van-der

i
v.
Therewith,

 
 -group.

Numbers of atoms correspond to those given in Table 3.

352

Vi
= VC,1 + 4VC,10 + 8VH,124 = 5.0 + 413.1 + 82.0 = 73.4 3;

i
crl.p.

Vi = VC,10 + 2VH,124 = 13.1 + 22.0 = 17.1 3.

i
v.
Then

Vi
= 73.4 + 217.1(m 2) = 50 + 34.2m.

i
crl.p.

Value ai Vi + b j
= 2(aCVC,10 + aH2VH,124)(m 2) = 2(0.02113.1 +

i
j

l.ch.
19.9822.0)(m 2)103 = (160.39m 320.78)103 3K1.

= KC(VC,1 + 4VC,10) + KH8VH,124 = [1.15(5.0 + 413.1) +


Value K i Vi

i
crl.p.
2.30782.0]103 = 102.92103 3K1.
Substituting these values to equation (IV.89), we get
Tg (K ) =

73.4
103 .
160.39m 217.86

(IV.108)

Relation (IV.108) is true, when m 2. The dependence of Tg on m calculated


from this relation is depicted in Figure 53.

Figure 53. Dependence of Tg on m for the model polyethylene network with tetra-functional crosslinked point

353

If m = 1 or m = 0, the situation changes. When m = 1, the structure of the


network VI looks as follows
..
.
...
...

...

..
.

..
.

CH2
C

C CH2
...

CH2

...

...

CH2

...

...

C
..
.
This network consists of cross-linked pointV RQO\ HDFK &+2-group being
adhered to two neighboring cross-linked points.
Then

Vi = Vi
= VC,1 + 42VC,10 + 82VH,124 = 5.0 + 213.1 + 42.0 =

i
r.f. i
crl.p.

= 39.2 3;

K i Vi
= KC(VC,1 + 42VC,10) + KH82VH,124 = [1.15(5.10 + 213.1) +

i
crl.p.
2.30742.0]103 = 54.35103 3K1.
Since the present network contains no linear fragments, and for it

a V + b
= 0, softening of it may occur only as a result of its thermal
i
i
j

j
i
l.ch.
degradation, i.e. degradation by chemical bonds. Substitution of all parameters into
equation (IV.89) gives
Tg = Td =

39.2
103 = 721 K .
54.34

If m = 0, the structure of the network VI obtains the form


..
.

...
...

...

...

...
...

...

C C
... ...

...

..
.

..
.

..
.

...

...

...

...

..
.
i.e. possesses the ideal diamond structure, for which

...

354

Vi = VC,1 = 5.0 3;

i
r.f.

a V + b
= 0;

i
i
j

i
j

l.ch.

. 9
= KCVC,1 = 1.15103 3K1.

FUOS
L

Substitution of these values into equation (IV.89) gives


Tg = Td =

5.0
103 = 870 K .
5.75

Hence, the calculated temperature of onset of intense thermal degradation of diamond


is ~600C. It is known from literature that diamond is burnt off completely when
heated up to 8501000C.
Let us now consider behavior of the network
Structure VII
...

(CH2)m -2

CH2

CH CH2

..
.

(CH2)m -2

CH2

CH

...

(CH2)m -2

CH2

CH

...

CH2
(CH2 )m- 2
CH2
...

(CH2)m -2

CH2

CH CH2

...
which possesses a tri-functional cross-linked point, marked by dotted lines. For this
network,

9
= VC,6 + VC,10 + 7VH,124 = 9.0 + 313.1 + 72.0 = 62.3 3;

FUOS
L

Vi
= (VC,10 + 2VH,124)(m 2) = (13.1 + 22.0)(m 2) =

i
l.ch.
= (17.1m 34.2) 3;

Vi = Vi
+ 1.5 Vi
= 62.3 + 1.5(17.1m 34.2) =

i
r.f. i
crl.p.
i
l.ch.
= (26.65m + 11) 3;

a V + b
+ K i Vi
= 1.5[(aCVC,10 + 2aHVH,124)(m 2)l.ch. +

i
i
j

i
j
i

l.ch.
crl.p.
+ [KC(VC,6 + VC,10) + 7KHVH,124]crl.p. = {[1.5(0.02113.1 + 219.982.0)

355

(m 2)] + 1.15(9.0 + 313.1) + 72.3072.0}103 = (120.3m 152.76) 103 3K1.


Substituting all obtained above values into equation (IV.94), we get
Tg (K ) =

25.65m
103 .
120.3m 152.76

(IV.109)

The dependence of Tg on m is shown in Figure 54. It is clearly seen that Tg


abruptly decreases as m increases, i.e. as the size of linear fragments between
neighboring cross-linked points increases.

Figure 54. Dependence of Tg on m for the model polyethylene network with tri-functional cross-linked
point.

Equation (IV.109) is true at m 2. If m = 1, we obtain the network of the


following structure:
..
..
.
.
...

CH

CH2

CH CH2 CH

...

CH2
CH CH2 CH CH2 CH ...
..
..
.
.
This network consists only of cross-linked points, which are its repeating
fragments; these fragmentscross-linked points are marked by dotted lines.
For this network
...

Vi = Vi
= VC,6 +1.5VC,10 + 4VH,124 =

i
r.f. i
crl.p.
= 9.0 + 1.513.1 + 42.0 = 36.65 3;

356

a V + b
= 0;

i
i
j

i
j

l.ch.

K i Vi
= KC(VC,6 +1.5VC,10) + KH4VH,124 =

i
crl.p.
= [1.15(9.0 + 1.513.1) + 2.30742.0]103 = 53.41103 3K.
Substituting the values obtained into equation (IV.94), we get
Tg = Td =

36.65 3
10 = 686 K .
53.41

Finally, let us consider structure VII of the network under the condition m = 0.
This network displays the form:
...
...
...

CH CH CH

...

...

CH CH
...

...

CH
...

Here the branching atom is every carbon atom, so it should be written down that

Vi
= VC,6 + VH,124 = 9.0 + 2.0 = 11.0 3;

i
crl.p.

a V + b
= 0;

i
i
j

i
j

l.ch.

K i Vi = K i Vi
= KCVC,6 + KHVH,124 =

i
r.f. i
crl.p.
= (1.159.0 + 2.30732.0]103 = 14.964103 3K1.
Substituting these values of network parameters into equation (IV.94), we
obtain
Tg = Td =

11.0
103 = 735 K .
14.964

Let us now discuss the network of the third type, in which cross-liked chains
are of the same chemical structure, and cross-linking bridges are of a different one. As
an example, let us analyze the real network based on epoxy resin of the following
chemical structure [30]

357

Structure VIII
OH
...

CH2

OH

CH CH2

CH2

CH CH2

O
m

OH

OH
...

CH2

CH CH2
OH

CH2

CH CH2
OH

O CH2

CH CH2

CH2

N CH2

CH CH2

CH2

...

...

The stoichiometric composition of the network conforms to 2 moles of epoxy


oligomer per mole of curing agent, diamine. The cross-linked point of the network is
marked by dotted lines.
For the repeating fragment of the network, we get:

Vi = (VN,144 + 2VC,56 + VC,21 + 4VH,124) + [2VC,39 + 2VC,40 + 2VC,130

i
r.f.
+ 2VC,132 + 2VH,125 + 6VH,124 + 2VC,20 + 4VC,18 + VH,124 + (2VO,130 + 2VC,40
+ VC,39 + VC,132 + 5VH,124 + VH,125 + 2VC,20 + 4VC,18 +4VH,124)m]l.ch.,1 +
(4VC,18 + VC,19 + 4VH,124 + 1/2VC,12 + 1/22VH,124)l.ch.,2 = (0.9 + 214.6 + 10.2 +
42.0) + [212.2 + 216.2 + 22.7 + 25.6 + 24.7 + 62.0 + 211.6 + 412.7 + 42.0) +
(22.7 + 216.2 + 12.2 + 5.6 + 52.0 + 4.7 + 211.6 + 412.7 + 42.0)m] + (412.7 + 8.4
+ 42.0 + 1/212.6 + 1/222.0) = (300.6 + 152.3)m 3;

a V + b
= aC(2VC,39 + 2VC,40 + 2VC,20 + 4VC,18) + aH(2VH,125 +
i i j
i
j

l.ch.,1
VH,124 + 4VH,124) + aO,0(2VO,130) + aO,(2VO,132) + [aC(2VC,40 + VC,39 + 2VC,20
+ 4VC,18) + aH(5VH,124 + VH,125 + 4VH,124) + aO,0(2VO,130) + aO,s( VO,132)]m +
bd2 + bh2 + bm + (bd + bh + bm)m = {0.021(212.2 + 216.2 + 211.6 + 412.7) +
19.98(24.7 + 62.0 + 42.0) + 22.9522.7 + 16.025.6 + [0.021(216.2 + 12.2 + 211.6
+ 412.7) + 19.98(52.0 + 4.7 + 42.0) + 22.9522.7 + 16.05.6]m + 2(55.4) + 2(
139.6) + 16.0 + (55.4 139.6 + 16.0)m}103 = (519.3 + 490.6m)103 3K1;

a V + b
= aC(4VC,18 + VC,19 + 1/2VC,12) + aH(4VH,124 +
i i j
i
j

l.ch.,2
1/22VH,124) + bp = [0.021(412.7 + 8.4 + 1/212.6) + 19.98(40.2 + 2.0) +
+ (25.6)]103 = 175.6103 3K1;

K i Vi
= KNVN,144 + KC(2VC,56 + VC,21) + KH4VH,124 = [2.520.9 +

i
crl.p.
1.15(214.6 + 10.2) + 2.30742.0]103 = 66.1103 3K1.
Substituting all calculated values of network parameters into equation (IV.98),
we get:

358

Tg (K ) =

300.6 + 152.3m
10 3 .
519.3 + 490.6m + 175.6 + 66.1

(IV.110)

If m = 0, we obtain the structure of cured epoxy resin based on diglycidyl ether


of corresponding bisphenol
HO

OH

For this resin, the calculated value of Tg,calc = 395 K; the experimental value is Tg,exp =
396 K.

Figure 55. Dependence of Tg,calc on m for the cured epoxy resin (see text).

Figure 55 depicts the dependence of Tg on m, calculated by formula (IV.110).


When the linear fragment of the network decreases, the glass transition temperature
increases, especially intensive, when m 0.
Let us now consider the influence of structure features of polymeric networks
and their defects on the glass transition temperature Tg.
a) Influence of cross-links distribution in the network on the glass transition
temperature
Let us analyze first the influence of cross-links distribution along the chains
upon the glass transition temperature of the network. As a well-studied model
network, a system based on polystyrene cross-linked by divinylbenzene is chosen.
This network displays the structure:

359

Structure IX

...

(CH2

CH)m -1

CH2

CH CH2

CH

(CH2

CH)m -1

CH2

CH

...

...

(CH2

CH)m -1

CH2

CH CH2

CH

(CH2

CH)m -1

CH2

CH

...

..
This network is a tri-functional one, i.e. three chains yield from every crosslinked point of it (marked by dotted lines). It possesses

'
= VC,7 + 2VC,10 + VC,19 + 5VH,124 = 8.7 + 213.1 + 8.4 + 52.0 =

FUOS
= 53.3 3;

Vi
= VC,7 + VC,19 + 5VC,18 + 6VH,124 + (VC,10 + VC,7 + VC,19 +

i
l.ch.,1
L

5VC,18 + 8VH,124)(m 1) = 8.7 + 8.4 + 512.7 + 62.0 + (13.1 + 8.7 + 8.4 + 512.7 +
82.0)(m 1) = (109.7m 17.1) 3;

Vi
= 2VC,18 + 2VH,124 = 212.7 + 22.0= 29.4 3;

i
l.ch.,2

K i Vi
= KC(2VC,7 + 2VC,10 + VC,19) + KH5VH,124 = [1.15(8.7 + 213.1

i
crl.p.
+ 8.4) + 2.30752.0]103 = 72.9103 3K1;

a V + b
= aC(VC,7 + VC,19 + 5VC,18) + aH6VH,124 + [aC(VC,10 +

i
i
j

j
i
l.ch.,1
VC,7 + VC,19 + 5VC,18) + aH8VH,124](m 1) + bd/2 + bd/2(m 1) = {0.021(8.7 +
8.4 + 512.7) + 19.9862.0 + [0.021(13.1 + 8.7 + 8.4 + 512.7) + 19.9882.0](m 1) +
(55.4)/2 + (55.4)/2(m 1)}103 = (80.2 + 294.0m)103 3K1;

a V + b
= aC2VC,18 + aH2VH,124 + bp/2 = (0.021212.7 + 19.982.0
i
i
j

i
j

l.ch.,2
25.6/2)103 = 67.6103 3K1.

360

Substituting all calculated values of network parameters into equation (IV.98),


we get
Tg (K ) =

64.6 + 109.7m 3
10 .
60.3 + 294.0m

(IV.111)

Formula (IV.111) is true under the condition m 1. The dependence of Tg on


m calculated from formula (IV.111) is depicted in Figure 56 (note that when m ,
we get the calculated value of Tg for linear polystyrene, which equal 373 K).
It is seen that the glass transition temperature sharply increases as the number
of polystyrene units between neighboring cross-links decreases, when the number of
these units (m 1) 0.
When m = 0, the network structure IX is of the following composition
..
..

...

CH2

CH CH2

CH CH2

CH

...

...

CH2

CH CH2

CH CH2

CH

...

..
.
that corresponds to 100% cross-linking. Cross-linked point of this network is marked
by dotted lines. It possesses

'
= 21/2VC,10 + VC,7 + 22/2VH,124 + VH,124 + VC,19 = 21/213.1 +

FUOS
L

8.7 + 22/22.0 + 2.0 + 8.4 = 36.2 3;

= 0, because all atoms in cross-linked


The value of ai Vi + b j

i
j

l.ch.,1
chains are included in cross-linked points composition.

Vi
= 2VC,18 + 2VH,124 = 212.7 + 22.0= 29.4 3;

i
l.ch.,2

361

a V + b
= aC2VC,18 + aH2VH,124 + bp/2 = (0.021212.7 + 19.982.0

i
i
j

i
j

l.ch.,2
25.6/2)103 = 67.6103 3K1;

K i Vi
= KC(21/2VC,10 + VC,7 + VC,19) + KH(21/2VH,124 + VH,124) =

i
crl.p.
[1.15(21/213.1 + 8.7 + 8.4) + 2.307(21/22.0 + 2.0)]103 = 48.57103 3K1.
Substituting these values into equation (IV.98), we get
Tg (K ) =

36.2 + 29.4 3
10 = 565 .
48.57 + 67.6

Let us now consider the influence of cross-linked points distribution along the
chains on the glass transition temperature Tg. The analysis will be performed in terms
of the cross-linking degree . For the fragment of the network
... CH2 CH ...

..
let us determine all values of parameters from equation (IV.98).

Vi = 2VC,18 + 2VH,124 + VC,10 + VC,7 + VC,19 + 3VH,124 = 212.7 + 22.0

i
1
+ 13.1 + 8.7 + 8.4 + 32.0 = 65.6 3;

a V + b + K V = a 2V
C
C,18 + aH2VH,124 + bp/2 + KC(VC,10 + VC,7

i
i
j
i
i

j
i
i
1
+ VC,19) + KHVH,124 = [0.021212.7 + 19.9822.0 1/225.6 + 1.15(13.1 + 8.7 +
8.4) + 2.30732.0]103 = 116.2103 3K1.
Now it should be taken into account that in the repeat unit of polystyrene
...

CH2

...

CH

.. 2*

which yields from the network cross-  


2     


lines) is included into the cross-linked point composition. Then, for this polystyrene
unit, we get:

362

Vi = Vi = VC,10 + VC,7 + VC,19 + 5VC,18 + 8VH,124 = 13.1 + 8.7

ps
i
2 i
+ 8.4 + 512.7 + 82.0 = 109.7 3;
*

a V + b + K V = a (V + V
C
C,7
C,19 + 5VC,18) + aH6VH,124 + bd/2 +
i i j i i
j
i
i
2
KCVC,10 + KH2VH,124 = [0.021(8.7 + 8.4 + 512.7) + 19.9862.0 55.4/2 + 1.1513.1
+ 2.30722.0]103 = 238.1103 3K1.
For a normal polystyrene unit,
...

CH2

...

CH

we obtain Vi = 109.7 3;

i
ps

a V + b = a (V
C
C,10 + VC,7 + VC,19 + 5VC,18) + aH8VH,124 + bd/2 =

i
i
j

j
i
ps
[0.021(13.1 + 8.7 + 8.4 + 512.7) + 19.9882.0 55.4/2]103 = 294.1103 3K1.
When all parameters of equation (IV.98) are calculated, let us analyze three
variants.
1) Uniform distribution of cross-links
For this case, equation (IV.98) obtains the following form:

Tg (K ) =

Vi +

i
1
*

ai Vi + b j + K i Vi + ai Vi + b j + K i Vi +

j
i
j
i
i
1
i
2

+ (1 ) Vi

i
ps

+ (1 2 ) ai Vi + b j

j
i
ps

(IV.112)

where is the cross-linking degree equal to 1/m, which characterizes the part of
linked phenyl groups.
Substituting values of all above-determined parameters of the network into
(IV.112), we get

363

Tg ( K ) =

65.6 + (1 ) 109.7
10 3 .
116.2 + 238.1 + (1 2 ) 294.0

(IV.113)

Formula (IV.113) is true under the condition 0 < 0.5. If > 0.5, no free
units of polystyrene remains, and formula (IV.113) reduces to the form
Tg (K ) =

65.6 + (1 ) 109.7
10 3 .
116.2 + (1 ) 238.1

(IV.114)

When = 0, we obtain the calculated glass transition temperature of linear


polystyrene.
The dependence of Tg on the cross-linking degree determined with the help
of equations (IV.113) and (IV.114) is shown in Figure 57, curve 1.

Figure 57. Dependence of Tg on the cross-linking degree for the polystyrene and divinylbenzene
network: 1 uniform distribution of cross-linked points; 2 irregular distribution of crosslinked points; 3 the most ununiform distribution of cross-linked points. The points denote
Tg values detected experimentally.

Figure 57 indicates that the dependence of Tg on the cross-linking degree


consists of two parts with fast and slow increase of Tg. This is caused by a specific
influence of network cross-linked points upon Tg, which is indicated by the above
analysis.
2) A random distribution of cross-links
It is common knowledge that the function of distribution by distances for a
random field of points fitting a straight line is the following
F (l ) = 1 e

l / l avg

(IV.115)

where l is the distance between two neighboring points, and in the case under
consideration the distance between cross-linked points; lavg is the average distance.
Turning to the polymerization degree m, we get
F (m) = 1 e

m / mavg

(IV.116)

364

If cross-linked points are located alongside each other


...

CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH

...

...

CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH

...

WKH DPRXQW RI


&+2

JURXSV EHORQJLQJ WR VW\UHQH UHVLGXHV PDUNHG E\ GRWWHG OLQHV

in the scheme) and simultaneously included into the cross-linked point composition
will be lower than in the case, when cross-linked points are separated by linear
fragments:
...

CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH

...

...

CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH CH2

CH

...

Consequently, the value of ai Vi + b j + K i Vi decreases by the

value proportional to the amount of cross-links located alongside each other. Then

(1 ) ai Vi + b j increases by the same value.

j
i
ps
Since we are interested in cross-linked points located alongside each other,
F(m) at m = 1 is F (1) = 1 e

1 / mavg

. Then, taking into account that mavg = 1/, we get

F(1) = 1 e.

(IV.117)

Then equation (IV.103) in the second summand in the denominator will obtain
(1 e) = /e, and in the third summand, 1 2 + (1 e) = 1 /e
instead of (1 2).
When substitutions are made, we get

365

Tg (K ) =

Vi +

i
1

ai Vi + b j + K i Vi +
e

j
i
i
1

+ (1 ) Vi

i
ps

+ (1

a V + b + K V +

i
i
j
i
i

j
i
i
2

a V + b
i i j
j
i
ps

(IV.118)

/ e )

Substituting all calculated values of the network parameters into equation


(IV.118), we get
Tg (K ) =

65.6 + (1 ) 109.7
10 3 .

238.1 + (1 / e ) 294.0
116.2 +
e

(IV.119)

Formula (IV.119) is valuable under the condition ( + /e) 1; this conditionally


correlates with 0.66.
If ( + /e) 1, no free units of polystyrene remain, and formula (IV.119)
transits into equation (IV.114).
The dependence of Tg on obtained with the help of (IV.119) and (IV.114) is
depicted in Figure 57, curve 2.
3) The most ununiform distribution of cross-linked points.
This distribution corresponds to the case when a large number of cross-linked
points locates alongside each other and form significant sequences of cross-linked
points linked by linear fragments of polystyrene chains (see scheme in Figure 58).

Figure 58. Hypothetical case of the most ununiform distribution of cross-linked points of network
(scheme).

At this distribution of cross-linked points, the amount of polystyrene units


linked to the cross-linked point becomes small, and it may be written that

366

Tg (K ) =

Vi + (1 ) Vi

i
1
i
ps

ai Vi + b j + K i Vi + (1 ) ai Vi + b j

j
i
j
i
1
i
ps

(IV.120)

Substituting all values of the network parameters into (IV.120), we obtain


65.6 + (1 ) 109.7
10 3 .
116.2 + (1 ) 294.0

Tg (K ) =

(IV.121)

Calculations performed by equation (IV.121) lead to the dependence of Tg on


displayed in Figure 57, curve 3.
This figure also displays experimental data on the dependence of Tg on the
cross-linking degree , taken from ref. [160].
Experimental points fit well curve 2; this testifies that cross-linked points are
distributed in the real network in a random way.
b) Influence of the cross-linked point structure on the glass transition
temperature of network
Let us discuss one more question on the influence of the chemical structure of
the network upon its glass transition temperature. In this case, networks of equal
chemical structure of linear fragments but of different structure of cross-linked points
will be discussed. For this analysis, polydimethylsiloxane networks of the following
chemical structure are suitable [27]:
H3C
...

CH2

CH2

Si
O

H2C
...

...

Si

Si
O

CH3
CH2

CH3
CH2

Si

(O

CH3

CH3

CH3

CH3
Si)n -2
CH3

Si

CH2

...

CH3

H2C

(A)
H3C

...

Si O Si

CH3

CH3

CH2

CH2

CH2

CH2

Si

CH2

CH2 Si O Si CH2
H3C
CH3

CH2

CH3
...

CH3

CH3
(O

Si)n -2
CH3

Si

CH2

...

CH3

(B)

At low values of n (high cross-linked network) synthesis is based on


application of individual compounds with accurately assigned value of n. That is why
at conversion close to 100%, the value of Mc in these networks is determined with
high accuracy.
Prior to analyzing the influence of the structure of cross-linked points on Tg,
let us repeat the definition of the network cross-linked point from the point of view of
its chemical structure: cross-linked point of a network is a group of atoms, including
the atom from which branching begins, plus neighboring atoms chemically bonded
with the former one with their closest substituents (therewith, another end of the

367

branch may also be included into the corresponding cross-linked point, otherwise this
will be a simple branching). Then for considered networks, cross-linked points are of
the structure limited by dotted lines.
Let us determine all values of constants for these systems, included into
equation (IV.85):

Vi
= 3VC,106 + 3VC,109 + 3VC,109 + 3/2[4VC,106 + 2(n 1)3VC,106] +

i
r.f.
21VH,124 + 3/3[12VH,124 + 6(n 2) VH,124] + 3VO,135 + 3/2[VO,135 + (n
2)VO,135] + 3VSi,172 + 3/2[2VSi,173 + (n 2)VSi,172] = (239.85 + 108.15n) 3;

a V + b = a {3V
C
C,109 + 3/2[4VC,106 + 2(n 2)VC,106]} + aH{6VH,124 +

i
i
j

j
i

3/2[12VH,124 + 6(n 2)VH,124] + aO,03/2[VO,135 + (n 2)VO,135] + aSi3/2[2VSi,173


+ (n 2)VSi,172].
Substitution of all values to the present relation gives the value of

a V + b = (176.0 + 721.5n)103 3K1 (therewith,


b j = 0 , because

i
i
j

j
j
i

polymer contains no separate polar groups which induce strong intermolecular


interaction).
For the cross-linked point of the network (A), we get

K i Vi
= KC(3VC,106 + 3VC,109) + KH15VH,124 + KO3VO,135 +

i
crl.p.,A
KSi3VSi,172 = 252.3 3K1.

a V + b , and K V
into

i
i
j

i i
j
i
crl.p.,A
i

equation (IV.85) gives the following dependence of the glass transition temperature Tg
of the network on the polymerization degree:

Substitution of Vi ,

i
r.f.

Tg (K ) =

239.8 + 108.1n
10 3 .
428.3 + 721.5n

(IV.122)

Graphically, this dependence is shown in Figure 59. For networks with tetrafunctional cross-linked pints (scheme B), it can be easily shown that the calculated
dependence of Tg on the polymerization degree n is identical to relation (IV.122),
because each value in the numerator and the denominator of this relation is multiplied
by the same number 4/3.
Let us now discuss the influence of cross-linked points structure of networks
from polydimethylsiloxane on their glass transition temperature. Reasoning from
chemical ideas, possible cross-linked points may be depicted as shown below.

368

Figure 59. Dependence of Tg on 1/n for A and B structures (see text).

To calculate Tg of networks containing various cross-linked points, it is

advisable to determine first the values of ai Vi + b j and Vi for a

j
i
l.f.
i
l.f.
single linear fragment of the network between two neighboring cross-linked points:

Vi = 2VSi,172 + 4VC,106 + VSi,172(n 2) + 2VC,106(n 2)+ VO,135 +

i
l.f.
VO,135(n 2) + 12VH,124 + 6VH,124(n 2);

Vi = (72.1n 0.5) 3;

i
l.f.

a V + b = a [4V
C
C,106 + 2VC,106(n 2)] + aO,0[VO,135 + (n 2)VO,135]

i
i
j

j
i
l.f.
+ aH[12VH,124 + 6(n 2)VH,124] + aSi[2VSi,172 + (n 2)VSi,172];

a V + b = (481n 11.5)103 3K1.


i i j
j
i
l.f.

and ai Vi + b j
for all
The resultant values of Vi

i
j
i
l.f.

l.f.
considered networks are equal.
Displayed below are possible cross-linked points of networks from
polydimethylsiloxane and relations obtained from equation (IV.85) which allow
calculation of Tg in accordance with the number of the repeat units n in linear
fragments connecting these cross-linked points.

369

...
H3C
CH3
...

Si

CH3
H3C

Si

CH3

CH3

Si O

Si

CH3

Si

...

(1)

...

(2)

Tg (K ) =

144.2n + 33.7
10 3
962n + 11.3

CH3

...
CH3
...

Si

CH3
H3C

CH3

CH3

Si O

Si

CH3

Si

Tg (K ) =

108.15n + 52.9
10 3
721.5n + 46.5

CH3

..
.
CH3
...

...

...

CH3

CH3

Si O

Si

CH3

CH2

CH3

CH3

CH2

CH3

Si O

Si ...

CH3

CH3

CH3

CH3

CH3

CH3

Si O

Si

Si

Si

Si

CH3

CH2

CH3
...

...

...

...

CH3

CH3

CH3

CH3

CH3

Si O

Si

CH3

CH3

CH3

CH3

CH3

(4)

Tg (K ) =

72.1n + 58.1 3
10
481n + 61.2

Tg ( K ) =

72.1n + 49.3 3
10
481n + 48.8

CH3

CH3

Si

72.1n + 66.1 3
10
481n + 72.6

CH3
Si ...

Si

Tg (K ) =

...

Si O

Si

(3)

...

Si O

Si ...

CH3

CH3

(5)

According to the relation described above, dependences of Tg of networks on


the number n of polydimethylsiloxane units are shown in Figure 60. It is obvious that
the structure of a cross-linked point affects significantly Tg only if n = 14 ( = 0.25
1). As the distance between cross-linked points increases, the role of the latter
decreases abruptly, and already at n = 10 the glass transition temperature for all
networks approaches Tg of polydimethylsiloxane, which equals 150 K.

370

Figure 60. Dependences of Tg of networked polydimethylsiloxanes on 1/n for different types of crosslinked points. Designations 15 of cross-linked points in the Figure correspond to their
designations in the text.

c) Influence of dangled chains and other defects of network on the glass


transition temperature of polymer networks
Among the multitude of various variants of dangled chains, let us consider two
of them in detail. The first variant is a modification of network I, when a part of crosslink bridges is broken in halves:
...
...

...

C
CH2

(CH2)m -2

...
...

...

CH2
CH2

(CH2)m -2

CH2

..
.

CH2

(CH2)m-2

CH2

...

..
.

CH2
_ -2
(CH2 )m
2

CH3
CH3
_ -2
(CH2 )m
2

...
...

...

CH2
CH2

(CH2)m -2

CH2

CH2

(CH2)m-2

CH2

...

CH2
...
...
In this case, the cross-linked point of the network is kept, but it becomes trifunctional (in the present example, as it is obvious from the scheme, chemical
structure of a cross-linked point also remains unchanged).
The repeating fragment of the present structure consists of the repeating
fragment of tri-functional network and dangled chain. Then for the Van-der-Waals
volume, it may be written down that
...

371

Vi = Vi
+ 1.5 Vi (m 2) + Vi
,

i
r.f. i
crl.p.
i
v.
i
d.ch.

(IV.123)

where Vi
is the Van-der-Waals volume of the network cross-linked point,

i
crl.p.

marked by dotted lines in the scheme; Vi is the Van-der-Waals volume of

i
v.

&+ JURXS Vi
2

i
Therewith,

d.ch.

is the Van-der-Waals of the dangled chain.


Vi
= Vi 2 + Vi
,

i
i
d.ch. i
v.
CH 3

where Vi
is the Van-der-Waals volume of the end CH3

i
CH 3

Values of Vi
and

i
crl.p.

(IV.124)

JURXS

Vi were calculated above;

i
v.

Vi
= VC,13 + 3VH,124 = (17.2 + 32.0) = 23.2 3;

i
CH 3
Prior to further calculations, one more important note should be made.
Principally, there may not appear a situation, when a network possessing no dangled
chains, and a network containing them are of the identical structure. To put it
differently, appearance of dangled chains always implies a change in chemical
structure of the network, which affects the glass transition temperature and other
properties.
So,

Vi = 73.4 + 1.517.1(m 2) + 17.1(m/2 2) + 23.2 = (11.1 + 34.2m) 3;

i
r.f.

a V + b
= 1.5(aCVC,10 + aH2VH,124)(m 2) + (aCVC,10 +
i i j
i
j

l.ch.
aH2VH,124)(m/2 2) + (aCVC,13 + aH3VH,124) + bd = [1.5(0.02113.1 +
19.9822.0)(m 2) + (0.02113.1 + 19.9822.0)(m/2 2) + 0.02117.2 + 19.9832.0 +
(55.4)]103 = (160.4m 336.2)103 3K1.

372

Substituting all values calculated above into equation (IV.85), we get


Tg (K ) =

(11.1 + 34.2m) + (1 )(5.0 + 34.2m)


10 3
(160.4m 336.2 + 102.9) + (1 )(160.4m 217.86)

(IV.125)

5.0 + 34.2m + 6.1


10 3 .
160.4m 15.44 217.86

(IV.126)

or
Tg (K ) =

Equation (IV.126) is valuable up to 5, because at m 2 break of every


linear chain forms two cross-linked points of the network, from which dangled chains
yield.
The dependence of the glass transition temperature Tg on determined by
formula (IV.126) is displayed in Figure 61. It is obvious that Tg grows weakly with ,
the slope of this dependence decreasing as the number of units between cross-links m
increases. At first glance, it may seem that appearance of network defects as dangled
chains must cause a decrease but not an increase of Tg. However, it should be taken
into account that in this particular case, appearance of dangled chains does not
decrease the number of cross-linked points and even change their chemical structure.
Moreover, appearance of every branching requires introduction of the constant bd that
promotes increase of Tg.

Figure 61. Dependences of Tg on (with different m) for tetra-functional polyethylene network with
dangled chains and constant numbers of cross-linked points.

In the case when a cross-linked point loses is original meaning due to chains
break, i.e. it is no more the cross-linked point of the network, the glass transition
temperature obviously decreases. This type of the network is shown below:
CH3
_ -2
(CH2 )m
2
..
..
.
.
CH2
... C CH2 (CH2)m -2 CH2 C CH2 (CH2)m-2 CH2 C ...
...

CH2
..
.

...

373

In the case of this break of network linear fragments a linear chain is formed,
and constants ai and bj instead of Ki should be used in calculation of the former crosslinked point, i.e.

a V + b
= aC(VC,1 + 4VC,10) + aH8VH,124 = [0.021(5.0 + 413.1) +
i i j
i
j

crl.p.
19.9882.0]103 = 320.9103 3K1.
Then for a network fragment between two vertical dotted lines, it should be
given:

Vi = 73.4 + 17.1(m 2) + 17.1(m/2 2)2 + 23.22 = (17.2 + 34.2m) 3;

i
f.

a V + b = {320.0 + 0.021[13.1(m 2) + 13.1(m/2 2)2 + 17.22] +


i i j
j
i
f.
19.98[22.0(m 2) + 22.0(m/2 2)2 + 62.0]}103 = (80.2 + 160.4m)103 3K1.
Substituting all these values and previously calculated parameters of the
network into equation (IV.85), we obtain
Tg ( K ) =

(17.2 + 34.2m) + (1 )(5.0 + 34.2m)


10 3
(80.2 + 160.4m) + (1 )(160.4m 217.86)

or
Tg ( K ) =

5.0 + 34.2m + 12.2


10 3 ,
160.4m + 298.1 217.86

(IV.127)

(IV.128)

Figure 62. Dependences of Tg on (with different m) for tetra-functional polyethylene network with
dangled chains when the formation of them reduce a quantity of cross-linked points.

374

where is a part of the former cross-linked points in the network, from which 2
broken chains yield.
Relation (IV.128) it valuable at 0.25.
The dependence of Tg on calculated by equation (IV.128) is shown in Figure
62. Clearly, the glass transition temperature decreases abruptly as the part of crosslinked points with yielding two broken chains increases. The reason is that the place
of branching loses its role of the cross-linked point, and linear fragment between two
neighbor cross-linked points two-fold increases.
Let us now consider a tri-functional network with dangled chains. The scheme
of this hypothetical network based on polyethylene is shown below:
...

...

(CH2)m -2

CH2

CH CH2

(CH2)m -2

CH2

CH2

CH

...

(CH2)m -2

CH2

CH2

CH

...

CH2
_ -2
(CH2 )m
2

CH3
CH3
_ -2
(CH2 )m
2

CH2
...

(CH2)m -2

CH2

CH CH2

..

.
In this case, when linear chains break, the former cross-linked
point with
yielding a dangled chain loses its significance as a cross-linked point, and is included
to composition of a linear fragment with a branching. For the repeating fragment of
the network bordered by two vertical dotted lines, we obtain:

Vi = 62.3 + 17.1m 34.2 + 17.1(m/2 2) + 23.2 = (25.65m +17.1) 3;

i
f.

a V + b = a [V + 3V
C
C,6
C,10 + VC,10(m 2) + VC,10(m/2 2) + VC,13]

i
i
j

i
j

f.
+ aH[7VH,124 + 2VH,124(m 2) + 2VH,124(m/2 2) + 3VH,124] + bd = {0.021[9.0 +
313.1 + 13.1(m 2) + 13.1(m/2 2) + 17.2] + 19.98[72.0 + 22.0(m 2) + 22.0(m/2
2) + 32.0] 55.4}103 = (24.8 + 120.3m)103 3K1.
Substituting all these and previously calculated parameters of the network into
equation (IV.85), we get:
Tg (K ) =

(17.1 + 25.65m) + (1 )(11.0 + 25.65m)


10 3
(24.8 + 120.3m) + (1 )(120.3m 152.76)

or
Tg ( K ) =

11.0 + 25.65m + 6.1


10 3 ,
120.3m + 177.56 152.76

where is a part of structures with dangled chains.

(IV.129)

(IV.130)

375

Figure 63. Dependences of Tg on (with different m) for tri-functional polyethylene network with
dangled chains .
Figure 64 (right). Schematic representation of different defects of network (see text).

Dependences of Tg on with different m are shown in Figure 63. As the part of


dangled chains grows, Tg abruptly decreases, especially with small m.
Considered above were hypothetical network based on polyethylene with
fragments between cross-linked points and dangled chains of the same chemical
structure. Let us now consider a system with dangled chains of a different structure
compared with backbones of the network. Such a system based on epoxy resin was
synthesized and investigated in ref. [35]. Epoxy resin ED-20 was cured by
methyltetrahydrophthalic anhydride, some part of which was substituted by oleic acid
in order to obtain a network with dangled chains. Results of calculations and
experiments performed for such networks are discussed in detail in refs. [35, 125].
Besides dangled chains, the network may also contain such defects as isolated
loops. Figure 64 displays defects of various types as branchings or isolated loops.
Cross-linked points are the only branching points, in which chains yielding from them
are joined to neighboring chains, i.e. form a cross-link. These structures are marked
by + sign. If a chain forming no cross-link but representing a simple branch yields
from the branching point or forms an isolated loop at the sacrifice of attachment by its
end to the yielding point or another place of the same chain, these places are not also
cross-linked points (marked by sign).
This must be taken into account when equation (IV.85) is used for calculation
of the glass transition temperature. If a branching point is a cross-linked point (in the
sense discussed above), constants Ki must be used for calculation by equation (IV.85).
If a branching point is not the cross-linked point, for the same atoms constants ai and
bj must be substituted into equation (IV.85).
Using equation (IV.85), the amount of defects in the real network structure
may be estimated by comparing calculated and experimental values of Tg. Let us
perform such detailed analysis on the example of network based on methyl
methacrylate and 1-methoxy-1-vinyl-1-sila-2,7-dioxa-4,5-(1,2-carborano) cycloheptane copolymer. Chemical structure of this polymer is shown below [67]:
CH3
CH2

CH2

C O CH3
O

O
x

Si

CH2
C

CH

II

O CH3
O
CH2

C
B10H10

376

Under the thermal effect upon the copolymer, the cycle in structure II transforms into
a linear structure, and a network of the following chemical structure is formed:
CH3
CH2

CH2

CH

C O CH3

Si

C C CH2

O CH2

B10H10

CH3

II
According to the definition of the cross-linked point given above, its chemical
structure in the present network is the following:
...

CH2

...

...

CH
Si

...

O
.
Let us calculate values of

Vi , ai Vi + b j
i

and

K i Vi

for the

present network.
For structure I,

Vi = VC,10 + VC,13 + VC,1 + VC,48 + VC,41 + 8VH,124 + VO,129 + VO,139

i
I
= 96.8 3.
For structure II,

Vi = VC,10 + VC,41 + VC* + 2VC,40 + 2VC,117 + 10VH,124 + 10VH,128 +

i
II
3VO* + VSi,174 + 6VB,185 + 4VB,184 = 208.8 3.
Increments of the Van-der-Waals volume for atoms marked by * sign are absent in
Table 3, so they are displayed below:
C
Si

1.54
1.88

1.54

1.08

H ' = 8.0 3;
&

Si

1.64

1.50

C ' = 1.95 3;
2

C
O
O

1.64
1.64

Si

1.64

1.88

3
C '6L = 30.1 .

377

Let us now calculate values of

ai Vi + b j
i

for the ideal network with the

structure shown above.


For structure I,

a V + b = a (V
C
C,10 + VC,13 + VC,1 + VC,48 + VC,41) + aH(8VH,124) +
i i j
i
j

I
aO,sVO,129 + aO,s VO,139 + 3bd = 255.8103 3K1.
For structure II,

a V + b = a (2V
C
C,40 + 2VC,117 + VC,41) + aH(7VH,124 + 10VH,128) +

i
i
j

j
i
II
aB(6VB,185 + 4VB,184) + bd = 401.0103 3K1.

Let us now calculate K i Vi


for cross-linked point of the network

i
crl.p.
with structure displayed above.

d
K i Vi
= KC(VC,10 + VC*) + KH(3VH,124) + K O
(3VO*) + KSi(VSi*) =

i
crl.p.
69.9103 3K1.
Let us consider in the general form dependence of the glass transition
temperature on chemical structure of the network taking into account that it possesses
defects in the structure and branchings displayed in Figure 64. As noted above, if a
branch forms no cross-link, then, though chemical structure of the cross-linked point
in the branching point is the same as in the case of the cross-link formation, this crosslinked point loses its role, and constants ai and bj instead of Ki must be used for it in
calculation by equation (IV.85). The equation (IV.85) is rewritten in the form:

Tg =

Vi
i

a V + b
+ y K i Vi
+ (1 y ) ai Vi + b j

i
i
j

j
j
i
crl.p.
i
l.ch.
i
crl.p.

, (IV.131)

where y is a part of branching points acting as cross-linked points. Comparing


calculated and experimental glass transition temperatures, the value of y may be found
from the relation, deduced from equation (IV.131) .

378

y=

ai Vi + b j

j
l.ch. i
crl.p.

Vi ai Vi + b j

Tg, exp i

K i Vi
+ ai Vi + b j

j
i
crl.p. i
l.ch.

where Tg,exp is the experimental glass transition temperature.


If to form a network copolymers are used, as it takes place in this case, the
situation becomes more complicated. Assume that one of components of copolymer
does not form a network. Then in the ideal case (100% of cross-linking)

Q 9 + Q 9

 Q D 9 + E

+ Q D 9 + E

,
L

,,


+ . 9

FUOS ,,

L

(IV.132)

(n is the part of structure II).


If a part of branching points are not cross-linked points, then

Q 9 +

,
7J =

 Q D 9 + E + Q D 9 + E
+

,
OFK

+ Q 9

,,


+ \ . 9
+  \ D 9 + E

FUOS

,,


(IV.133)

where y is the part of active cross-linked points.


To calculate the part of active y and inactive (1 y) cross-linked points,

equation (IV.133) will be used. First, let us calculate the value of ai Vi + b j :

j
i

a V + b
= aC(VC,10 + VC*) + aH(3VH,124) + aO,0(2VO*) +

i
i
j

j
i
crl.p.,II
aO,sVO* + aSiVSi* = 490.4103 3K1.

379

Substituting found values of Vi ,

i
I

a V + b ,

i
i
j

j
i
II
(IV.133), we get

K i Vi
,

i
crl.p.

Tg (K ) =

Vi ,

i
II

a V + b ,

i
i
j

j
i
I

a V + b

i
i
j

j
i
l.ch.

96.8 + 112.1n
10 3 .
255.8 420.5ny + 635.6n

into

equation

(IV.134)

Experimental dependences of Tg on composition of component II, taken from


ref. [20] are shown in Figure 65. It also displays calculated dependences of Tg on
concentration of the second component with different parts of acting cross-linked
points y. Calculation was performed by equation (IV.134). The Figure displays that,
for example, if the part of component II is 20%, the amount of active cross-linked
points is ~58% and of inactive 42%. At the amount of this component of 10%, the
part of active cross-linked points is ~55%, etc. Hence, although networks obtained in
ref. [20] are high cross-linked ones, but are significantly defective. If 20% of structure
II is contained by the copolymer, the part of the sol fraction is 18.5%. Consequently, a
part of inactive cross-linked points is contained in the branched copolymer, being
washed off the system. The remaining part of inactive cross-linked points (42 18.5 =
23.5%) is in the network composition and forms a structure of branchings or isolated
loops type.

Figure 65. Dependences of Tg on proportion n of 1-methoxy-1-vinyl-1-sila-2,7-dioxa-4,5-(1,2carborano)cycloheptane; the proportion of active cross-linked points: 1 (1), 0.9 (2), 0.8 (3),
0.7 (4), 0.6 (5), 0.5 (6), 0.4 (7). The points denote the experimental values.

Let us give one more example of estimation of the network structure based on
comparison of the calculated and experimental glass transition temperatures Tg [105].
Consider ebonites as an example of this network, because various types of networks
may be realized in them, and because essential but conflicting experimental data on
their structure and mechanical properties have also been obtained. Let us calculate Tg
of polyisoprene-based ebonite. Let us assume that the amount of sulfur participating
in the reaction corresponds to a single atom of sulfur per repeat unit of isoprene.

380

Several variants of formation of a network, in the composition of which structures of


four types may participate, will be considered.
..
..
.
.
1)

...

CH2

CH C CH2
S
..
.

3) ...

CH2

...

2)

...

CH2

CH C CH2
S

CH3

...
;

CH3

S
..
.

CH C CH2

..
.
4) ...

... ;

CH2

CH C CH2

... .

CH3

(S)n CH3
..
.

Structure 1 contains monosulfide bridges, structure 2 disulfide bridges,


structure 3 polysulfide ones, and structure 4 represents original units of
polyisoprene. For structure 1, all atoms are included into cross-linked points of the

network. That is why for it ai Vi + b j = 0. Then, we obtain:

j
i

Vi = 2VC,10 + VC* + VC** + VC,13 + 8VH,124 + VS,157 = 92.2 3;

i
1

K i Vi = KC(2VC,10 + VC* + VC** + VC,13) + KH8VH,124 + KSVS,157 =

i
1
136.9103 3K1.
For constants of the Van-der-Waals volume for several atoms included in the
structure of networks 14 are absent in Table 3, they are shown below:
S
C

1.76
1.54

S
1.54

1.54

C ' = 10.2 3; C


H
S

2.10

1.76
1.54

1.54

1.54

C ' = 6.1 3;


C
2.10

S '$ = 18.9 3.

For structure 1, the glass transition temperature Tg calculated from equation


(IV.85) coincides with temperature of the onset of intense thermal degradation Td
(because all atoms are included into the cross-linked point of the network): Tg = Td =
673 K.
Consequently, if sulfur is completely consumed for polyisoprene vulcanization
in a manner that monosulfide bonds are formed only, then Tg must coincide with the
temperature of onset of intense thermal degradation, and it equals to 400C. In
practice, ebonites possess the glass transition temperature of 60120C [84]. That is

381

why it must be assumed that besides monosulfide bridges, di- and, probably,
polysulfide ones must occur. Let us calculate Tg for a network system possessing
various types of the bridges between polyisoprene chains.
For structure 2,

Vi = 2VC,10 + VC* + VC** + VC,13 + 8VH,124 + VS,164 = 111.1 3;

i
2

K i Vi = KC(2VC,10 + VC* + VC** + VC,13) + KH8VH,124 + KSVS,164 =

i
2
172.8103 3K1.
For structure 3,

Vi = 2VC,10 + VC* + VC** + VC,13 + 8VH,124 + VS,164 + (n 2)VS*;

i
3

Vi = (73.3 + 18.9n) 3;

i
3

K i Vi = 172.8103 3K1;

i
3

a V + b = (n 2)a V ;
S
S*
i
i
j

j
i
3

ai Vi = (141.7n 283.5)103 3K1;

i
3
For structure 4,

Vi = 2VC,10 + VC,16 + VC,15 + VC,13 + 8VH,124 = 81.5 3;

i
4

a V + b = a (2V
C
C,10 + VC,16 + VC,15 + VC,13) + aH8VH,124 + bd + b
i i j
j
i
4
= 400.7103 3K1.
Let us deduce a relation binding Tg of ebonite to molar parts of bridges
containing mono-, di- and polysulfide groups. Let 1, 2, , n be molar parts of
cross-linked points (bridges) containing mono-, di- and n-sulfide groups. At assigned
stoiciometric relation between sulfur S and polyisoprene, occurrence of a single
disulfide bond leads to unlinking of one polyisoprene unit. Occurrence of a trisulfide
bond causes appearance of two unlinked polyisoprene units, etc. Consequently, the
part 1 of unlinked polyisoprene units will be equal to

1 = 2 + 2 3 + ... + (n 1) n =

i=n

(i 1) i .

i =2

(IV.135)

382

Then, on the basis of the above considerations and equation (IV.135), we get

 9L +  9L +  + Q 9L +
L

L

L


7J =

 . L 9L +  . L 9L +  + Q . L 9L +
L
L

L



L =Q
+ 9L L L 
L
 L =


L =Q


+ D L 9L + D L 9L + E M L L 

L 
M
L
 L
 =
i=n

Therewith, it is taken into account that

= 1 1 2 n =

i =2

i=n

(i 1) i

(IV.136)

i .

i =1

Let us consider some particular cases. Assume that the network is formed only
of mono- and disulfide bonds. Then

1 Vi + 2 Vi + (1 1 2 ) Vi

i
1
i
2
i
4
Tg =
. (IV.137)

1 K i Vi + 2 K i Vi + (1 1 2 ) ai Vi + b j

j
i
1
i
2
i
4

Taking into account 2 = 1 1 2, finally, we get

(1 2 2 ) Vi + 2 Vi + 2 Vi

i
1
i
2
i
4
Tg =
. (IV.138)

(1 2 ) K i Vi + 2 K i Vi + 2 ai Vi + b j

j
i
1
i
2
i
4
Equation (IV.138) binds Tg of the ebonite network with the molar part 2 of
disulfide bridges. Substituting numerical values of all increments into this equation,
we obtain
92.2 + 8.2 2
103.
(IV.139)
136.9 + 299.7 2
The dependence of Tg on 2 is shown in Figure 66. As the part of disulfide
bonds increases, Tg smoothly decreases and at 2 = 0.5 (disulfide bridges exist only)
becomes equal 60C.
Tg =

383

Figure 66. Calculated dependence of Tg for ebonites containing only mono- and disulfide cross-links
on the part 2 of disulfide links.

Taking into account that literature data exist on the formation of bridges only
with a limited number of sulfur atoms (the average value is assumed to be 2.6 [84]),
let us consider the question of the influence of the part of bridges of different length
(up to three sulfur atoms) upon Tg.
Designate the number of linked units as mlnk = m1 + m2 + m3, where m1, m2 and
m3 are the numbers of units linked by mono-, di- and trisulfide bridges, respectively.
The total number of units both linked and free equals mlnk + m2 + 2m3 = m1 + m2 + m3
+ m2 + 2m3 = m1 + 2m2 + 3m3. Then
m1
;
m1 + 2m2 + 3m3
m3
3 =
;
m1 + 2m2 + 3m3

1 =

m2
;
m1 + 2m2 + 3m3
m2 + 2m3
4 =
.
m1 + 2m2 + 3m3

2 =

Divide numerators and denominators of these expressions by m3

 =

2 2
2 2 + 2 2 +  ;


 =
2 2 + 2 2 +  ;

 =

2 2
2 2 + 2 2 +  ;

2 2 + 
 =
2 2 + 2 2 +  .

(IV.140)

Substituting equations (IV.140) into equation (IV.136), we get

%J . =

2 2 + 2 2 +  +  2 2 + 


 .
2 2 + 2 2 +  +  2 2 + 

(IV.141)

Let us now consider a triangle (Figure 67), analogous to the Gibbs triangle, on
the sides of which parts of bridges with various bonds are plotted, and isolines
represent levels of equal Tg. Clearly, Tg of ebonites may change in an extremely wide
range at the existence of mono-, di- and trisulfide bridges. The maximum softening
temperature (monosulfide bonds only exist), coincident with the temperature of

384

Figure 67. Triangular diagram characterizing the influence of composition of network of ebonites
on Tg; the numbers at isolines denote Tg.

intensive degradation in this case, equals 400C. The minimal value Tg = 10C is
typical of the network containing trisulfide bridges only.
A case may also be imagined when sulfuric bridges may be of different
lengths, and a definite distribution by length exists in the real system. Such variants
were discussed in ref. [105], where the glass transition temperature of ebonites was
calculated as the value dependent on the function of distribution by lengths of cross
bridges composed of sulfur atoms. Consequently, it has been concluded that the
bridges may not contain a large number of sulfur atoms. So the statement set in
literature that the bridges are mostly formed of disulfide chains is true.

Chapter V. Temperature of transition into the viscous flow


state for amorphous polymers

V.1 Estimation of temperature of transition into the viscous flow


state of polymers
Let us note once again that, in dependence on temperature, amorphous
polymer may exist in three physical states: glassy, rubbery and viscous flow. Methods
of estimation of the glass transition temperature, Tg, of polymers based on chemical
structure of the repeat unit are discussed above. Estimation of not only glass transition
temperature Tg, but also temperature of transition into the viscous flow state Tf, which
determine the temperature range of the rubbery state, by the chemical structure of
polymer is extremely desirable. Therewith, the value of Tf is necessary for processing
of polymers.
The position of Tf in relation to Tg of the polymer is closely linked with the
molecular mass of the chain M.
An attempt to estimate Tf was made in ref. [96]. It is based on application of
an expression that sets the dependence of zero (Newtonian) shear viscosity 0 on the
molecular mass M [177]:

0 =

0.5
1.5
RT 2 M 3 M e 1 M e
1
+
K

,
15
M e2 M 3 M

(V.1)

where K = b 2 N 02 k BT 2 M 2 (b is the segment length; N0 is the number of


segments composing the present macromolecule; is the friction coefficient of the
segment, including external and internal friction of the initial macromolecule; kB is
the Boltzmann constant); Me is the mass of the macromolecule, starting from which
entanglements of chains affect significantly the melt viscosity; R is the universal gas
constant.
According to refs. [175, 176], equation (V.2) holds in the range of Me < M <
5Me.
It may be shown that the value of K does not depend upon M. Actually, b2 =
2
n0a , where n0 is the number of repeat units in a segment; a is the typical size of the
repeat unit; = 0n0 (0 is the friction coefficient for a single repeat unit). Then
K = b 2 N 02 k BT 2 M 2 = 0 n0 n0 a 2 N 02 k BT 2 M 2 =

(V.2)

= 0 (n0 aN 0 )2 k BT 2 (M a n0 N 0 )2 = 0 a 2 k BT 2 M a2 .

Hence, the value of K depends upon characteristics of the repeat unit only and

is independent of the molecular mass; therewith, a ~ Vi

the Van-der-Waals volume of the repeat unit.

23

, where

Vi
i

is

386

Equation (V.2) is used in ref. [84] for determination of the temperature of


transition into the rubbery state Tf. Displaying no details of analysis, let us write down
the final expression for estimation of Tf:
Tg
Tf

=1+

1
ln
C

B
3
1
0.5

X 1.5 ( AX ) 1 +

(V.3)

where C is a constant (C = 26); X = M/Ms, M is the molecular mass of the polymer, Ms


is the molecular mass of the mechanical segment of the macromolecule, i.e. the
molecular mass starting from which the glass transition temperature becomes
independent of M; A = Ms/Me; B = (A0.5 1)3 + 1/3.
As the data in ref. [177] indicate, for most polymers (polystyrene, poly-methylstyrene, polyvinylacetate, polyisobutylene, polyisoprene) A 2.4. For
polyethylene only, A 5.26.8, and for polybutadiene, A 4.3. Calculations for
different values of A give the following constants B:
A = 2.4 B = 0.5;
A = 4.3 B = 1.57;
A = 5.2 B = 2.43;
A = 6.8 B = 4.48.
Substituting values of parameters A, B and C into equation (V.3), for the
general case we obtain:
Tg
Tf

=1+

1
ln
26

B
.
3

1.5
0.5

M
M
1

2.4
1 +
M
M
3

s
s

(V.4)

To use equation (V.4), it is necessary first to estimate the value of Ms. It may
be calculated with the help of equation (V.1) with regard to the condition that Ms =
NsM0, where Ns is the polymerization degree of a mechanical segment of
macromolecule; M0 is the molecular mass of the repeat unit of the polymer.
The value of Ns (or Ms) determines the transition temperature to the rubbery
state of the polymer, when Tg is yet independent of the molecular mass. However,
when N < Ns (or M < Ms), the polymer also transits into the glassy state, but then Tg
will depend on the molecular mass. In this case, Tg = Tf (but therewith, it should be
taken into account that Tg is lower than the true glass transition temperature).
Expression (V.4) enables Tf to be also determined for this case.
It is obvious that the border transition temperature into the viscous flow state
is limited by the temperature of onset of intense thermal degradation Td (see Chapter
VII for details). From this point of view, not at every molecular mass a polymer may
transit into the viscous flow state. Using the expression (V.4) and assuming that Tf =
Td, the maximal polymerization degree (or the maximal molecular mass) may be
determined, at which the polymer may be transited into the viscous flow state (Table
20). However, temperature ranges of the viscous flow state, obtained in this case, are
by 1520% higher than really observed ones. This is due to the fact that due to

387

polydispersity of synthetic polymers, a part of chains transits into the viscous flow
state (actually, a superposition of the viscous flow and the rubbery states is observed),
which causes overload of longer chains.
Table 20
Maximum degree of polymerization Nmax at which the polymer can be transferred from the
glassy state to the viscous flow state without degradation
Polymer
Tg/Td [84]
Xmax
Ns
Nmax
Polyethylene
0.575
112
3136
28
Polyisobutylene
0.420
72
144
10368
Polystyrene
0.612
18
320
5760
Polyvinylacetate
0.716
227
2043
9
Poly(methyl methacrylate)
0.506
40
307
12280
Poly(vinyl chloride)
0.630
15
230
3450

To illustrate the above-said, let us analyze some experimental data displayed


in ref. [109], in particular, by polyisobutylene and polyvinyl chloride. Analysis of the
data on polyisobutylene with the help of expression (V.4) indicates that at N > 10 400
the polymer is unable to be transited into the viscous flow state. All the rest molecules
with higher polymerization degree would inevitably degrade (under the influence of
either mechanical stresses or temperature) when tried to be transited into the viscous
flow state. The latter causes the fact that long molecules disrupt due to their overload.
Broken chains may be found after thermomechanical analysis by determination of the
molecular-mass distribution.
The value of Tfcalc at N = 10 400 found with the help of expression (V.4)
equals to 181C, whereas Tfexp = 120C. Reduction of Tf ( Tfexp Tfcalc = 61C) is
stipulated by the fact that, according to the data from [109], longer molecules are
~50-fold overloaded because of shorter molecules transition into the viscous flow
state. If the experiment is conducted at the load of 0.07 MPa, long molecules in the
range of Tf are affected by the load of ~3.5 MPa.
Hence, if we want to obtain Tf independent of more or less disregarded
attendant factors as polydispersity with the help of thermomechanical analysis, it is
necessary to introduce additional limits for the polymerization degree of images,
therewith, setting the condition that the value of T = Tfexp Tfcalc is smaller than the
limits of the experimental error. The polymerization degree N, up to which
polydispersity does not affect Tf, may be estimated from the empirical expression:
NTg
N sTf

2.3.

(V.5)

The influence of polydispersity must be displayed in dynamic mechanical


tests. In this case, asymmetry of the temperature dependence of the mechanical loss
tangent in the area of the -transition (at T < T) is also associated with transition of
shorter chains with N < Ns into the viscous flow state. A polymer possessing no lowmolecular component (N > Ns) must display a symmetric maximum of losses.

388

V.2 Dependence of Newtonian viscosity on molecular mass of


polymer in a wide range of its change
It is common knowledge that the reptation DoiEdwards model has allowed
substantial developments in the rheology of concentrated solutions and melts of
polymers. Already in initial works by DoiEdwards [225], the following law for
Newtonian viscosity 0 was obtained:

0 ~ (M/Me)3 at M Me,

(V.6)

where M is the molecular mass of the polymeric chain; Me is the molecular mass
starting from which physical entanglements between chains affect molecular motion
of the chain.
This formula was deduced only with regard to reptational motions in the
chain. Further on, with regard to fluctuation of the tube contour, in which reptational
motion proceeds, Doi [226] succeeded in refining dependence (V.6) for M/Me >> 1.
The form of the expression deduced by Doi
M

0 ~
Me

M
1 1.47 e
M

0.5 3

(V.7)

enabled the dependence of 0 on M/Me to be described for M/Me >> 1. However,


expression (V.7) in the range of values M/Me < 20 indicates an abrupt deviation with
experimental data [226].
More accurate calculation of fluctuations of the tube contour has enabled Lin
[174, 175] to deduce an expression:
3
1.5
0.5 3
M
1M
Me
1 1.47
0 ~
+ e ,
3 M
M
M e

(V.8)

which describes the behavior of 0 down to M/Me ~ 5 or even 1, if the value of Me is


corrected in each of these areas.
As indicated in the work by Lin [174], Doi had insufficiently utilized the
possibilities of the theory. Moreover, as it will become clear below, if summation in
appropriate expressions deduced in ref. [227] is performed completely, then the
dependence 0 = f(M/Me) may be obtained within the framework of the DoiEdwards
theory. This dependence describes efficiently the experimental data in the total area of
M/Me variation, i.e. from unit to as large as one likes value (in the area preceding the
chain rupture at shear).
As indicated in ref. [96], calculated estimation of the glass transition
temperature of polymers in the range Me < M < Ms requires taking into account not
only the chemical structure of the repeat unit of the polymer, but also its viscosity
depending on the molecular mass.
Obtaining reliable analytical dependences of viscosity on molecular mass is of
the utmost importance for proteins when calculating the glass transition temperature

389

(specificity of proteins is so that, in contrast to synthetic polymers, many of them have


the molecular mass in the range Me < M < Ms).
In contrast to protein macromolecules, synthetic polymers usually possess a
wide molecular-mass distribution. The presence of low-molecular fractions causes a
change of the glass transition temperature of the system and spectra of relaxation
times of it. Hence, for both proteins and synthetic polymers, a significant role is
played by entanglements between chains (the notion of entanglement is interpreted
below). By virtue of the fact that the molecular mass of protein macromolecules is not
large enough, entanglements affect the glass transition temperature. In this respect,
proteins occupy a special place among high-molecular compounds.
Taking into account the polymerization degree is also important for many
heat-resistant polymers, which possess no rubbery state due to low polymerization
degree N < Ncr, where Ncr is the critical polymerization degree, and at heating transit
directly to the viscous flow state.
An attempt was made in ref. [772] to deduce an analytical expression, which
allows description of the dependence of Newtonian viscosity in the whole range of
changes of the molecular mass Me M Mcr. Following this work and according to
refs. [226, 227], let us write down an expression for the relaxation shear modulus:
G (t ) =

4nkTL
A (t ) B (t ) C (t ) .
5a

(V.9)

Here

A = 1+

N cr

tp 2
,

exp

p =1

 
95

,
 =    + [  ]H[S

5= 5

C =

N0

(V.10)

tp 2
,

2 n 2 exp

n =1

where A, B, C are relaxation times for motions of different types: A is


relaxation time of the linear chain between neighboring entanglements; B is
maximal relaxation time of the whole chain; C is time of reptational motion of
chain.
In accordance with ref. [227] expression (V.10) may be rewritten in
following form:

 
95 

 =  + [  ]   H[S

5= 5


the
the
the
the

(V.11)

((E) is the function of transformation of a chain length element (affinor), which


depends on the stretch ratio). Then, transferring from summation to integration, under
the condition

390

N0

2 p2 = 1

p =1

it may be written down


N0

2 p2

p =1

N0

dp

2
1 p

(V.12)

Let us consider the range of such times t when A(t) = 1. With regard to
expression (V.11), let us write down that
,
BC = C + [(E) 1] B
where
=
B

p2 n2
exp
4 p 2 n 2 t + .
C
p, n
B
64

(V.13)

In future discussion, we will be interested in the maximal Newtonian, or zero


shear viscosity 0 which, according to ref. [226], is determined as

0 = G (t )dt ,
0

and in this case equal

QN7/



(  ) %   
+ &   .
 =
D
5  3 S Q S  + Q %
3 Q

&

(V.14)

Because according to ref. [227],


2

N
N
B = 2 0 A , C = 6 0 A ,
Ne
Ne
then

B 1 Ne
=
.
C 3 N0

(V.15)

Let us now calculate sums included in expression (V.14) with regard to


condition (V.15). In accordance with transition (V.12), let us write down

391

N N

0 0
1
1
dpdn
4 p 2n 2 2 1 N e 2 p 2n 2 2 1 N e 2 .
n
n
p +
p +
1 1
3 N0
3 N0

64

Transiting to new variables x = 1/p and y = 1/n, let us calculate the integral
0.5

1/ N0

1 Ne

xN 0
0 .5
2
3 N 0
1 Ne
y dy

x arctg
.
= 1 +
1 Ne 2
1 Ne 2
3 N0
x + N0
x
y2 +
3 N0
3 N0

Taking into account that

(V.16)

1 Ne 2
x << 1, integral (V.16) may be presented
3N 0 N 0

in the following form:


1/ N0

1 Ne
y 2 dy

= 1 +
1 Ne 2
3 N0
x
y2 +
3 N0

0 .5

1 Ne

x arctg
3 N0

0.5

x.

(V.17)

With regard to expression (V.17), let us calculate the integral


1/ N0 1/ N0

1.5
0 .5
0.5

1 Ne
1 Ne
7 1 3N 0
x 2 y 2 dy
. (V.18)
arctg

=
+
1 N e 2 12 4 N e
3 N 0
3 N0
x
y2 +

3 N0

Analogously, the sum at C may be calculated, which equals


8

N0

1
= .
3
1 n

2n4
n

dn

(V.19)

Substituting values of sums (V.18) and (V.19) into expression (V.14), we


obtain:
=

  1    1 
QN7/


0 DUFWJ  10 +  . (V.20)
(  ) +





D
  10
 1
 1

As indicated in ref. [227], (E) may be presented as

 =  +


+


( ),


where are components of the strain tensor of the chain.

392

If a polymer is considered incompressible = 0 , then the relaxation

process B is displayed in the non-linear area only.


With regard to volumetric compressibility (which, in fact, was made by Lin) in
L
1
B (B is a constant, which may be found from the condition
the form of =
L0
3
of equality of the viscosity calculated by expression (V.20) and the Rous viscosity at
M = Me); fluctuations of the chain contour will contribute to the linear viscosity. But
taking into account fluctuations of the chain contour length, according to ref. [226],
causes a change of coefficient at C, which will be taken into account by constant C.
Then expression (V.20) may be presented in the following form:
0 =

1 .5
0 .5
0 .5

1 Ne
1 Ne Ne
4nkTL0 L 7 1 3 N 0



arctg
+ C . (V.21)
C B +
15a
3 N0
3 N 0 N 0
L0 12 4 N e

~
Considered in ref. [226] are fluctuations of the contour length L (t ) , according

2 1/ 2
~
to which values of fluctuations are L = L (t ) L
and are determined from the

correlation:

L L(Me/M0)1/2.
The same work has indicated that the average contour length of the chain is
determined as follows:
~
L (t ) L .
The analogous meaning of L(t) is indicated in ref. [174]. According to this work, the
contour length of the chain is determined from the same correlations.
One more important point of the work [174] is that at time, t equal to the
correlation time of fluctuations L(t) and designated as B, a part of tube, which is still
in the stress state, shortens to length L0. Further on, expression for G(t) in ref. [174] is
presented in the form:
G (t ) =
0.5

TL
t
t
B B
,
+ C C
5a
B
C

L M e
=
.
L M
Because according to refs. [226] and [174]

where B =

393

L M e
=

L0 M

0.5

0.5

M
, C ~ 1 1.47 e
M

and taking into account that Ne/N0 = Me/M, expression (V.21), finally, may be
presented in the following form:

0 =

M
RT 2

KM e
15
Me

M
M
M
,
+ f P
+ f D
1.28 f L
Me
Me
M e

(V.22)

where fL is the Lin function, fL(M/Me) = 1/3 (Me/M)1.5; fD is the Doi function,
fD(M/Me) = [11.47(Me/M)0.5]3; fP is an additional function obtained in work [767],

1 ( 
!







 0 ) =  0 DUFWJ 0 , is density.




The numerical coefficient at fL(M/Me) is calculated under the condition that at

M/Me = 1, 0 from expression (V.22) coincides with the value of 0R from the Rouse
formula, i.e.

0R =

RT 2
KM .
36

Correlation (V.22) is true at M/Me 1. In the case when M/Me < 1, in equation
(V.22) it should be taken M/Me = 1 in order to transit to the Rouse expression.
Let us consider two functions (M/Me) which take into account influence of
the molecular mass (polymerization degree) on Newtonian viscosity 0:
M M
=

P
Me Me
and

M
M
M

+ f P
+ f D
1.28 f L
Me
Me
M e

(V.23)

L(M/Me) = [fL(M/Me) + [1 (Me/M)0.5]3](M/Me)3,

where P(M/Me) is the function deduced in ref. [767]; L(M/Me) is the function
deduced in refs. [174, 175].
Figure 67 represents calculation results for the ratio P(M/Me)/L(M/Me)
depending on X = M/Mc in the range of variation of X: 1 X 40. A difference
between the results suggested in ref. [767] and the results by Lin is observed in the
area of high values X ~ 10. But if results obtained by the Lin formula for L and by
the Doi formula for D are compared, it is found that this deviation is much higher. At
X ~ 10 D/L = 0.676, and P/L = 0.935, i.e. the Lin formula overestimates in the
area of large values of X. Equation (V.23) enables data to be improved, obtained by
the Lin formula in the range of high X, and approximate them to the experimental
data.

394

Figure 67. Dependence of P/L on X.

Basing on expression (V.22), the calculation scheme for determination of the


transition temperature to the viscous flow state Tf, suggested in ref. [96], may be
improved. Using the same expressions presented in ref. [96], the expression for Tg/Tf
will have the following form:
Tg
Tf

= 1+

1
1

ln B X 1.5 (AX )0.5 1.47 + 0.427 + 1.08(AX )1.5 1

C
9 AX

1
arctg

3 AX

0.5

(V.24)

where A = Mc/Me, X = M/Mc, C = 26 [96], B = (A0.5 1.47)3 + 0.427 + 1.08A1.5(1


1/9A)arctg(3A)-0.5. When A = 2.4 [230], B =1.805. If the ratio Tg/Tf is determined at
1/A = X = 2.4-1, then Tg/Tf = 1.08 and the glass transition temperature Tg* of polymer,
which possesses M < Mc, is Tg* = Tg/1.08 = 419 K. For example, if Tg = 180C, then
Tg* = 146C. The maximal shift of Tg at N = Ne will represent Tg = 34C.
Calculations by the Lin formula indicate Tg/Tf = 1.066, Tg* = 152C and Tg = 28C.
Expressions obtained for 0 = f(N/Nc) allow a significant refinement of sought
for dependence of 0 in the area of Ne N Nc. However, when they are used for
determination of Tg* in the area of Ne N Nc, they refine Tg* insignificantly
compared with the Lin expression [174, 175]. That is why, in future consideration,
determining Tg* or Tf at N > Nc, more simple Lin expression may be used.
Basing on the approach considered, irregular polymeric systems may be
subdivided into three classes according to the type of transition from the glassy state
into the viscous flow state.
1. Substances of the polymeric type (polymers in their classic meaning), which
display Tg independent of the polymerization degree.

395

2. Polymeric systems of the EdwardsDoi type, in which a significant role of the


reptation motion is played and Tg depends on the polymerization degree N,
therewith Ne N Nc.
3. Substances of the Rouse type oligomeric systems with the polymerization
degree N < Ne.
In the case of the Rouse substance, the Newtonian viscosity of a melt may be
presented in the following form:

0R =

N A a a 2
RT 2
N.
KM =
36
36M a

In accordance with the principle of determination of typical temperatures


which was used by us in ref. [96], the Rouse glass transition temperature will be
determined from the condition of equality of the coefficients of viscosity at a
temperature change

R a (TR ) N R
= 1,
E E (TE ) N e

(V.25)

where R and E are densities of polymer melts, included in the Rouse and the Lin
formulae, respectively; a and E are friction coefficients of chains from the same
formulae, respectively.
Suggesting that R = E, correlation (V.25) transforms to the following
equation:

%#
=
%

0
# .

(V.26)

Taking into account that (T) ~ exp(E*/RT) and that, according to ref. [96],
E* = CTg, expression (V.26) may be presented in the form:
N
TR
1
= 1 + ln e .
TE
C
N
Here C = C

Tg

(V.27)

Tg

= 1.08 and C 28, N < Ne.


TE TE
Let polymer possess Nc = 360, Ne = 150, and NR = 75. Then
,

TR =

Tg
TE
=
.
1.025 1.067

Let Tg = 453 K, then TR = 409 K = 136C, Tg = 180C, TE = 146C.


The value of B (and of A, as well) may be determined as the cross point of the
tangent appropriate to the Rouse law and the tangent appropriate to the EdwardsDoi
law for low-molecular fractions of the current polymer.

396

As indicated above, a significant role in determination of Tg in the range of


N < Nc is played by the parameter A = Nc/Ne. An additive scheme of determination of
Nc basing on their chemical structure of the repeat unit of the polymeric chain was
suggested in ref. [96]. Ref. [767] gives the additive scheme for determination of Ne.
Attempts to determine Ne were already made long ago since the reptation
model had appeared. In a series of works [230, 231] Ne was estimated on the basis of
topological ideas considering cross-linked points as a consequence of chain
entanglements. Analytical expressions obtained are complicated for analysis and do
not solve the final task setting of a correlation between Ne and the chemical
structure of the chain. Moreover, it has been already indicated in works [233, 234]
that topological cross-linked points are unable to make a significant contribution to
the thermodynamic properties of polymers because of low probability of their
appearance. The latter experimental data [229] indicate that Ne significantly depends
on temperature.
Basing on the results of works [229, 233, 234], let us suggest the following
scheme of Ne determination. First, all cross-linked points in polymer are of the
physical type and appear due to capture of one chain by a potential pit of another
chain at thermal motion of them relative to each other. Hence, a network with
physical cross-linked points appears in the polymer. Secondly, a stress = Eel
appears in between cross-linked points of the chain, i.e. any fixation of parts of the
chain causes occurrence of stress between fixation points due to thermal motion (this
fact is known in the theory of the rubbery state). And, thirdly, the reptation motion
inside a tube of diameter d appears when a cross-linked pointoscillator formed by
captured unit of a neighboring chain loses its stability under the effect of applied
stress.
The points formulated above may be presented in the following form:
Ee max

d 2
= f max ,
4

(V.28)

where Ee is the elasticity modulus of the network, Ee = 3RT/Me; Me is the molecular


mass of the chain between physical cross-linked points, Me = Ne; is the molecular
mass of the repeat unit; d is the diameter of the tube in which the reptation proceeds;
fmax is the maximal force at which the cross-linked pointoscillator loses its stability,
and reptation motion starts.
Monograph [28] indicates solution of the problem of stability loss by an
oscillator under the effect of high intensities and, according to these data,

3 Tm
f max = ak BT 1 1
e T
4

max = const1/ar0.

(V.29)
(V.30)

Substituting expressions (V.29) and (V.30) into expression (V.28), we obtain:


Ne = d2/(Cb2),
where

(V.31)

397

3 Tm
C = 1 1
const 2 .
(V.32)
4e T

Expression (V.31) is analogous to that obtained in ref. [96]. Moreover, it


enables the meaning of parameter C to be defined concretely. The dependence of C
on T completely indicates experimentally measured dependences of Ne on temperature
[96]. Using the Beaman rule Tg/Tm = 2/3, expression (V.32) may be presented in the
form:

9 Tg
C = 1 1

8e T

const .
2

(V.32)

In fact, Tg/Tm may change in a wide range depending on the chemical structure
of the repeat unit, and this fact may be taken into account with the help of an
appropriate additive scheme [96].
Using formulae (V.31) and (V.33), and the expression for Ne, as well, deduced
in work [96], Nc/Ne may be presented in the form:
1F
O
 7J
= FRQVW7J   
.
(V.34)
H 7
1H
G

The value of const is constant for all polymers, l03 = Vi , d2 is the

i
r.u.
cross-section square of the repeat unit.
This analysis makes it possibility to describe all three states of the polymer,
the temperature of transition from one state to another, as well as the area of N < Nc
via typical temperatures. This allows approaching from new positions the
interpretation of results of acoustic spectroscopy for polydispersity polymers,
influence of short chains (N < Nc) at T < T (temperature of -transition), for example,
-transition.

Chapter VI. Melting point of polymers

The melting point is determined as the temperature at which a polymer transits


from the crystalline state into the viscous flow state. In contrast to low-molecular
substances, in which this process proceeds in a jump-like manner, in the case of
polymers melting is observed in a temperature range. This happens due to
polydispersity of polymeric chains, their branching and imperfection of crystallites
formed. The equilibrium and experimental melting points are distinguished. The
equilibrium melting point Tm0 = H m S m , where Hm is the melting enthalpy, Sm
is the melting entropy. The equilibrium melting point is determined by the point of the
phase equilibrium between a monocrystal of the polymer and its melt. Since perfect
monocrystals are difficult to obtain from the polymer, the equilibrium melting point is
determined by methods of extrapolation, for example, by extrapolation of the
dependence of the experimental melting point on the size of crystallites or on the
molecular mass of the polymer.
The melting point Tm is a physical characteristic and is most difficult to
calculate. The case in point is calculation based on the chemical structure of the repeat
unit of the polymer. Let us discuss two approaches to solution of this problem. One of
them is based on estimation of the relation between the glass transition temperature Tg
and the melting point Tm. It should be noted that, according to the Beaman rule [132],
Tg/Tm 2/3. However, the detailed analysis of a large range of polymers of extremely
various structures has indicated [172] that this relation varies in a wide range,
although for a broad group of polymeric systems it is ~2/3.
The equation expressing the relation between glass transition temperature Tg
and melting point Tm is deduced in ref. [42] basing on the experimental data,
according to which the coefficient of molecular packing of a crystalline polymer at
the melting point is approximately equal to the coefficient of molecular packing of a
amorphous polymer of the same structure at the glass transition temperature, i.e.
melting of a crystalline polymer and transition of an amorphous polymer from the
glassy state into the rubbery state occurs when the same part of the empty volume is
reached.
The relation for Tg/Tm is presented in the following form [42]:

Vi

Tg
i
=
A
Tm ( i Vi ) + j

(VI.1)

Here i = (k0 kg)/ki (ki is the partial coefficient of packing of the i-th atom); j are the
constants taking into calculation the contribution of strong intermolecular
interactions; A = kg/(k0 kg) = 10.418; the meaning of the rest of designations is the
same as in equation (IV.39). Values of i and j are shown in Table 21.

399

Table 21
Values of parameters i and i of various atoms and types of intermolecular interaction
Atom or type of intermolecular interaction
Designation
i
i, 3
Silicon
0.0840

Si
Carbon
0.08685

C
Hydrogen
0.0740
H

Oxygen in the backbone


0.0621

O,b
Oxygen in a branch
0.0963
O,s

Nitrogen in the backbone


0.0212

,b
Dipoledipole interaction
-0.0727

d
Hydrogen bond
-0.0188

h
p-Substitution
-0.100

Hydrogen bonds in polyamides:

Aromatic polyamides
0.422

Even aliphatic polyamides


0.392

Evenodd aliphatic polyamides


0.392

Odd aliphatic polyamides


0.467

Aliphaticaromatic polyamides containing an even


0.445

number of CH2-groups
Aliphaticaromatic polyamides containing an odd
0.544

number of CH2-groups
Aliphaticaromatic polyamides based on iso-,
0.445

terephthalic acids and p, m-phenylenediamines


containing CH2-groups only

In many cases, calculations performed by equation (VI.1) indicate good


coincidence with experiment. To increase the accuracy of calculations, it is advisable
to know group contributions ( i Vi ) + j of polymeric groups, from which
i

many polymers are composed. These group contributions are obtained by summing up
values of i Vi for atoms participating in the current group. Then values of j typical
of each polar group and type of substitution in aromatic cycles, and small
readjustments, which increase accuracy of the calculation, as well, are added to the
sum. The readjustments are deduced by comparing experimental and calculated
values of Tm for a series of polymeric standards. Values of ( i Vi ) + j for a
i

series of atomic groups are shown in Table 22.


Table 22

Corrected values of quantities

( i Vi ) + j for a series of basic fragments (See text)


i

Chemical structure of a fragment and neighboring atoms bonded to it

&&+2&
1

C......

CH

.......C

( i Vi ) + j
i

2
1.434
2.795

CH3
C H3
C ......

C
C H3

.......C

4.302

400

C......

CH

.......C

2
4.228

C2H5
C......

CH

.......C

3.652

O CH3
C......

CH

.......C

2.702

&&&O2&

3.770

Cl

C......

CH

.......C
7.889

CH3
C......

.......C
9.177

C......

.......C

1.576

O
C......

CH

.......O

3.064

CH3
CH3
C ......

.......O

4.623

CH3

C......

CH

.......O
8.175

C......

.......O
6.745

C......

.......O

2&+22

6.645

1.972

401

.......O

O......

7.028

O......

.......O

6.928

.......O

9.826

CH3
O......
CH3

&2&
&2&
&1+&
S......

.......C

S......

.......C

S......

.......S

0.211
0.168
0.153
6.567

6.467

6.671

6.571

S......

S......

.......S

.......N

6.724

6.624

S......

.......N

S......

.......O

S......

.......O

6.849

6.749

402

&6&
1

C
N

C......

2
1.338

12.000

O
C......

.......C

O
C

.......C
O

9.455

C
O

C......

.......C
13.290

CF3
C......

.......C

5.997

CF3
.......C

C......

.......C

C......

6.463

6.263

NH

C......

.......C

8.630

&&+22
&2&
&&+21
1.54
C......

C O

1.50

.......C

1.703
0.130
1.573
2.083

O
1.54
C......

C O

1.37

.......C

2.039

O
C......

.......N

1.856

O
C......

.......N

6.619

403

O......

6.519

.......N

1.37

C......

1.54

NHC

.......C

2.403

O......

6.802

.......N

6.776

.......N

N......

N......

C......

6.902

.......N

O......

6.676

.......N
O

N
C

.......C

16.646

O
C
O......

9.882

.......C

C
O

O
C
C......

9.600

.......C

.......C

C
O
O......

NH
N

8.912

404

O
C

9.704

S......

.......C

C
O
NH

S......

.......C

8.734

N
C H3
O ......

Si

.......O

5.986

6L 2 6L

0.042

C H3

O......

.......O

2.541

O
C......

.......C

2.049

CH3
N......

.......N
10.685

C......

.......C
10.172

& &)2 &

2.048

Application of these values enables the melting point to be calculated with


high accuracy. This is indicated by Table 23, which displays experimental values of
Tm for a series of polymers, as well as the calculated values obtained with the
application of atomic and group contributions.

Table 23
Experimental and calculated values of melting point Tm for a series of polymers
Tm, K
Tm, K
Tm, K
calculated by
calculated by
Polymer
experimental
equation (VI.1)
equation (VI.5)
1
2
3
4
Polyethylene
410
410

Polypropylene
449
451
421
Polybutene
405
413
418
Poly-4-methylpentene-1
508
500
504
Poly(ethylene adipate)
323
334
321

405

1
Poly(ethylene oxide)
Poly(propylene oxide)
Poly(tetramethylene oxide)
Poly(tetramethylene adipate)
Poly(trimethylene adipate)
Polystyrene
Poly--vinylnaphthalene
Poly(hexamethylene adipamide)
Poly(hexamethylene sebacamide)
Polyamide-6
Polyamide-11
Polyvinylisobutyl ether
Poly(methyl methacrylate)
Polyvinylethyl ether
Poly(dimethyl siloxane)

2
339
340; 348
338; 309
332
311
513; 523
633
539
499; 488
488; 499
467; 455
438
433
417
234

3
348
356
311
314
315
509
625
539
494
496
481
497
465
398
234

4
314
342
349
334
311

Melting points of copolymers cannot be described by a simple relation,


deduced from equation (VI.1) using the rule of additivity.
Another approach [29] is based on consideration of the repeat unit of the
polymer as a set of anharmonic oscillators.
According to ref [114], the free energy of an anharmonic oscillator is:
F F0 +

m 2 2 3
a a a y 2 ,
0
2
3

(VI.2)

where

=
=

=
F0 = kT ln 2sh
; y2 =
cth
; =
;
0 2m
m
2kT

2kT
(m is the mass of the atom; is the elasticity coefficient of the oscillator; is the
coefficient of anharmonicity of the oscillator; a is the value which characterizes the
displacement of the equilibrium point of the harmonic oscillator relative to zero).
The value of a in expression (VI.2) is deduced from the condition of the
minimum of the free energy by a, i.e. when
m2a a2 y20 = 0.

(VI.3)

The condition of stability loss of this system is considered, and the critical
temperature Tcr, at which stability is lost, is determined from the expression:
Tcr =

where

=
,
A +1
k B ln
A 1

(VI.4)

406

2m m 2
A=
.
2
The critical temperature Tcr is a phase transition temperature, i.e. of melting in
this case.
If the both LennardJones potential, with the help of which parameters of the
equation (VI.2) are determined, and the expression for the coefficient of volumetric
expansion are used, the following correlation for estimation of the melting point may
be obtained [29]:

K i Vi
1
= i
,
Tm
Vi

(VI.5)

where Ki = 18.52R/(zDi) (R in the universal gas constant; z is the coordination


number; Di is the energy of intermolecular interaction occurred at the expense of the
i-th atom).
The number of values of Ki is determined by the number of atoms composing
the repeat unit. But because some of atoms participate in the composition of polar
groups with strong dipoledipole interaction, hydrogen bonds, etc., the latter may be
taken into account by addition of a part of energy strong intermolecular interaction,
stipulated by contribution of the i-th atom, to the energy of dispersion interactions Di.
Then
K i,1 ~

1
1
;
; K i,2 ~
Di
Di + Did

.  =
L

'


; etc.,
+ ' K

(VI.6)

where Did is the contribution of the i-th atom into the dipoledipole interaction;

'K is the contribution of the i-th atom into hydrogen bonding; etc.
Computerized calculations performed according to equation (VI.5) by the
method of least squares have indicated [29] that for polymers, containing atoms of
carbon, hydrogen and oxygen in the repeat unit, for satisfactory calculation of the
melting point of a series of polymers basing on chemical structure of the repeating
L

d
unit, it is enough to know DH, DH
, D0, and D0d parameters (Table 24); values of Di
were calculated at z = 4.

Element
Hydrogen
Oxygen

Table 24
Numerical values Ki, Di and Di for hydrogen and oxygen atoms
Di, kcal/mol
Designation
Ki,x103, deg1
Di, kcal/mol
KH
10.42
0.88

10.03
0.92
0.04
d
KH
16.5
0.56
KO

13.3
0.70
0.14
d
KO

407

Table 23 displays calculated values of Tm for a series of polymers containing


atoms of carbon, hydrogen and oxygen in the repeat unit. In spite of quite good
convergence of the calculated and experimental data, it should be noted that the
present method is of limited application.

Chapter VII. Temperature of onset of intense thermal


degradation of polymers

Let us consider the characteristic of thermal stability of polymers, determined


with the help of thermogravimetric analysis. The temperature dependences of the
substance mass at continuously increasing temperature are determined with the help
of this method (thermogravimetric curves). It is common knowledge that for most of
polymers, thermogravimetric curves possess the shape schematically represented in
Figure 68. Estimating the thermal stability of a polymer, let us use the temperature of
onset of intense thermal degradation Td determined by the point of intersection of
tangents to two branches of the thermogravimetric curve (see Figure 68).

Figure 68. Schematic representation of the thermogravimetric curve and the method of evaluation of Td
value.

The correlation, indicated in ref. [88], which estimates the temperature of


onset of thermal degradation Td is deduced on the basis of consideration of valency
bound atoms as a set of anharmonic oscillators which form the repeat unit of the
polymer. At this point, approach is the same as for estimation of the glass transition
temperature Tg and the melting point Tm. But in this case, the energy of chemical
bonds but not the energy of intermolecular interaction is taken into calculations,
although the latter also displays a significant influence on the energy of dissociation
of chemical bonds. For example, it is common knowledge that the energy of
dissociation of CC (carboncarbon) bonds changes in the range from 30 to 90
kcal/mol [64] depending on which group contains carbon atoms, i.e. according to their
valence surrounding. The same also relates to other pairs of valence bonded atoms
(CO, CS, CN, etc.).
When a polymer is heated up, its volume changes, this change being
composed of two parts: increase of the free volume and change of chemical bonds
lengths. Analysis of these changes has led to the following dependence of the
temperature of onset of intense thermal degradation Td on the parameters of the
chemical structure of the polymer [88]:

409

xi Vi
1
= 42 i
,
Td
Vi

(VII.1)

where xi is the partial coefficient of volumetric expansion of the i-th atom, appeared
at the sacrifice of the change chemical bonds lengths; Vi is the Van-der-Waals
volume of the i-th atom; Vi is the Van-der-Waals volume of the repeat unit of
i

the polymer.
Therewith
3R
,
xi =
2d
0 i

(VII.2)

where is the coefficient of anharmonicity, =

1 3
2 d 3

; is a force constant,
d0

; is the potential of chemical interaction. All values are given for the i-th
d 2
atom, chemically bonded with other atoms.
To estimate the energy of dissociation of chemical bonds, let us use the Morse
potential

( d ) = E e a ( d d 0 ) 1 1 ,

(VII.3)

where d is the distance between chemically bonded atoms; d0 is the equilibrium


distance; E is the energy of dissociation of chemical bonds. Taking into account
expression (VII.3),

xi =
Then
Td =

3 R

.
4 ad 0 E i

(VII.4)

Vi
i

Ki Vi

(VII.5)

where
K i = 31.5

R
.
(ad 0 E )i

(VII.6)

In the case of dissociation by CH bonds, the values of constants are d0 =


0.108, Eavg = 396 J/mol, a = 0.0266 nm1.
Values of Ki are shown in Table 25.

410

Table 25
Conventional symbols and numeric values of constants Ki
Designation of
Atom or group
Note
Ki103, 3K1
constant
1
2
3
4
Carbon
KC
1.150

1.920
Suitable for carbon atom participating
Carbon
K Cd
in a polar group
Hydrogen
KH
2.307

d
0.556
Suitable for hydrogen atom
Hydrogen
KH
participating in a polar group
Oxygen
KO
0.058

d
1.572
Suitable
for
oxygen
atom participating
Oxygen
KO
in a polar group
d
2.520
Suitable for nitrogen atom
Nitrogen
KN
participating in a polar group
d
Nitrogen
0.411
Suitable
for nitrogen atom
K N,c
participating in a heterocycle
Sulfur
KS
1.900

6.300
Suitable for sulfur atom participating
Sulfur
K Sd
in a polar group
Fluorine
KF
1.360

d
Chlorine
2.500
Suitable
for
chlorine
atom
K Cl
participating in a polar group
h
3.450
Suitable in the presence of hydrogen
Hydrogen bond at the
K OH
bond at the sacrifice of OH-group
sacrifice of OH-group
h
2.200
Suitable
in the presence of hydrogen
Hydrogen bond at the
K NHCO
bond at the sacrifice of NHCO-group
sacrifice of NHCO-group

Because atoms may participate in the composition of polar groups possessing


a specific intermolecular interaction, their contribution to thermal stability will be
different from the contribution of the same atoms possessing only weak Van-derWaals interaction. For example, polymers may contain the following polar groups:
CH3;

O
NH C O
O

C O

C OH;

O
;

OH;

C N;

NH C

O
N

C
O

O
O
;

Cl;

O CH2

When Td is calculated by equation (VII.5), taking into account of the


intermolecular interaction is performed in the following way.
If the atom considered is not a participant of the polar group, its parameter Ki
is multiplied by the appropriate Van-der-Waals volume. But if the atom is included
into the polar group, its parameter Ki is designated as K ih or K id (h is the hydrogen
bond, d is the dipoledipole interaction), and K ih or K id is multiplied only by the
appropriate Van-der-Waals volume, i.e. contribution of the atom into the Van-derWaals interaction is not taken into account as the significantly weaker value.
For representatives of various classes of polymers, examples of calculation of
Td values are shown in Table 26.

411

Table 26
Calculated and experimental values of temperatures of onset of intense thermal degradation Td
in inert medium for a series of polymers
K i Vi 10 3 , Vi , Td, K Td, K
Chemical structure of the repeat unit
i
i
Exper.
Calc.
3
3/K
1
2
3
4
5
48.6
34.2
713
704
CH2CH2
81.4

51.5

673

673

186.1

115.6

623;
633;
643

621

372.6

231.2

593;
653

621

241.0

166.1

663

688

CH

177.6

109.9

633

618

CH3

164.3

96.4

600
623

588

213.1

126.15

563

591

CH3
CH2

CH

(CH2)5

NH

O
C

(CH2)4

C NH

(CH2)6

NH

(CH2)2

O C

CH2

CH2

C
C O CH3
O

CH2

C
O
C
O

CF2CF2
CH2

CH

70.8

54.9

773

775

95.0

54.1

570

569

606.0

396

683

653

C N
C

C HN

NH
C
O
C
O

412

C O

570.0

389.5

728

683

777.0

513.0

733;
728

660

527.0

381.0

753

723

668.0

437.4

663;
653;
658

655

O
C
O
C
O

O
C
O

O
O

CH3
O

C
CH3

O
O C

Equation (VII.5) allows not only estimation of the temperature of onset of


intense thermal degradation of the polymer at its heating in an inert medium, but also
makes it possible to solve a series of other problems. For example, for some time it
was not clear whether intense thermal degradation may be started at temperatures
below the glass transition temperature Tg. To put it differently, must the polymer at
heating first transit from the glassy state into the rubbery state or into the viscous flow
state in order to make possible thermal degradation, or may the latter also proceed in
the glassy state?
For many polymers, the temperature of onset of intense degradation is indeed
above the glass transition temperature or the melting point. However, in a number of
cases, degradation starts at temperatures below the glass transition temperature or the
melting point. The latter is stipulated by the fact that stability of a chemical bond
depends crucially on dipoledipole interactions and hydrogen bonds. Let us discuss
this in more detail.
If at the same backbone a polymer possesses polar groups in its composition
which occupy a significant Van-der-Waals volume compared with other atoms, not
participating in the composition of these groups, this polymer possesses a lower
temperature of thermal degradation. Let us show it on the example of polystyrene
CH2

CH

In the case of polystyrene, existence of a bulky polar phenyl group leads to the
following expression for calculation of Td by expression (VII.5):

413

Td (K ) =

109.9
d
K C 22.2 + K H 6.0 + K Cd 17.5 + K H
10

103 = 618

(deduction of this expression requires the data from Table 3 to be used, which indicate
Van-der-Waals volumes of atoms).
Formally, if polystyrene is considered as a substituted polyethylene, in which
one atom of hydrogen in every unit is substituted by phenyl, it may be concluded that
such substitution causes a decrease of Td. This reduction may start before the glass
transition temperature is reached. Polymers with bulky side substituents containing
polar groups behave In this manner. Polymethylidenephthalide is one of these
polymers:
CH2 C
O
C
O
The formula for calculation of the temperature of the onset of intense
degradation of polymethylidenephthalide, deduced from (VII.5), is of the following
form:
Td ( K ) =

126.15
d
d
K C 21.6 + K Cd 83.3 + K H 4 + K H
8 + KO
9.25

103 = 592 .

The presence of a bulky polar grouping in a side chain induces high glass
transition temperature, equal to 390C for this polymer. This value of Tg appears as
the result of both calculations and experiments [55]. The latter should be discussed in
more detail, because in the cases in which the intense thermal degradation starts
before the glass transition temperature is reached, direct determination of Tg becomes
impossible. In this connection the following method of determination of Tg for
polymethylidenephthalide was used in refs. [55]: a series of methylidenphtalide
copolymers with styrene, methyl methacrylate and methyl acrylate of different
compositions was synthesized. The glass transition temperature was determined for
these copolymers. Figure 69 represents dependences of Tg on the composition of these
copolymers. As the concentration of methylidenephthalide increases, Tg grows first,
but when Tg reaches values at which the intense thermal degradation starts, this
growth terminates. In this area of high concentrations of methylidenephthalide,
softening of copolymers proceeds at the sacrifice of their decomposition. At the same
time, extrapolation of dependences of Tg on composition to the 100% concentration of
methylidenephthalide results in Tg = 390C for polymethylidenephthalide. Hence, the
onset of intense degradation of this polymer is at 300C and, consequently, for
polymethylidenephthalide, the onset of this process is at much lower temperatures,
than the glass transition temperature.
On the contrary, in the case of weakly polar groups of polymers (polyethylene,
polydimethylsiloxane, etc.), the glass transition temperature and the melting point are
much lower than the temperature of thermal degradation. A case may be selected (on
the example of polyheteroarylenes), when the glass transition temperature and the
temperature of thermal degradation will be practically coincident.

414

Hence, when analyzing the influence of the chemical structure of polymers on


their thermal characteristics and when predicting the polymer properties, it should be
taken into account that one and the same groupings may cause the opposite influence
on various thermal characteristics.

Figure 69. Dependences of Tg on the molar part of styrene (1), methyl methacrylate (2) and methyl
acrylate (3) for the copolymers of these components with methylidenephthalide.

Of special attention is the case when degradation of a polymer starts from


dissociation by end groups. If this dissociation does not lead to formation of the same
end groups, as mentioned above, calculation of the value must be performed on the
basis of the chemical structure of the repeat unit of the polymer. In this case, if
dissociation of end groups, which proceeds easier than decomposition in the polymer
backbone, leads to formation of the same end groups, the calculation must be
performed on the basis of the chemical structure of end groups only. In this case,
polyformaldehyde decomposition is typical:
HO CH2

(CH2

O)n ...

Calculation of Td by the structure of end groups (marked by dotted lines) leads


to the following value:
Td,1 ( K ) =

39.4
d
h
4 + K OH
10.3
K Cd 25.1 + K H

103 = 447 .

This value coincides well with the experimental value of Td, determined by the
first change of the mass of polyformaldehyde under conditions of thermogravimetric
analysis (Figure 70). This analysis was performed for a sample stabilized by

&+ &+ 2 JURXSV LQWURGXFHG LQWR SRO\PHU LQ  FRQFHQWUDWLRQ ,W LV


2

common knowledge [69] that introduction of the units mentioned prevents premature
decomposition of polyformaldehyde proceeding by the end groups. That is why,
calculation of temperature of the second decrease of the sample mass is conducted by
structure of the repeat unit. It should be noted preliminarily that polyformaldehyde is
a polar polymer, chains of which possess a strong intermolecular interaction. Then

415

Td,2 (K ) =

Figure 70.

32.5
d
h
K Cd 25.1 + K H
4 + KO
3.4

103 = 583 .

7KHUPRJUDYLPHWULF FXUYH IRU SRO\IRUPDOGHK\GH VWDELOL]HG E\  RI

&+ &+ 2
2

groups. (The experiment was performed in the medium of argon at the heating rate of 5
deg/min).

This value coincides also well with the experimental value, determined from
the thermogravimetric curve (see Figure 70).
Hence, analysis of the influence of the chemical structure on thermal
degradation of a polymer makes it possible to perform scanning by various polar
groups existing both on the ends of macromolecules and in the repeat units.
Therewith, it may occur that the temperature of decomposition of these groups is
lower than the onset temperature of intense thermal degradation of the whole
polymer. It is advisable to perform further scanning with regard to chemical
transformations of these groups at polymer heating. Clearly, such calculation analysis
is unable to substitute completely experimental studies of thermal degradation of
polymers, which is extremely complicated.
In the case of copolymers, equation (VII.5) obtains the form:

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n
Td =
,

1 K i Vi + 2 K i Vi + ... + n K i Vi

i
1
i
2
i
n

(VII.7)

where 1, 2, , n are the molar parts of components 1, 2, , n; Vi ,

i
1

Vi , , Vi are the Van-der-Waals volumes of these components;

i
n
i
2

K i Vi , K i Vi , , K i Vi are the series of constants for

i
n
i
1 i
2
components 1, 2, , n. In the reduced form, correlation (VII.7) becomes

416

k =n

Td =

k Vi

k =1
i
k
.
k =n

K i Vi
k

k =1
i
k

(VII.8)

If it is desirable to express the temperature of onset of intense thermal


degradation via analogous values for homopolymers composed of components 1, 2,
, n, then, according to equation (VII.5) and expression (VII.7), it may be written
that:

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n
Td =
,

Vi
Vi
Vi

i
n
i
1
i
2
+2
+ ... + n
1
Td,1
Td,2
Td, n

(VII.9)

where Td,1, Td,2, , Td,n are temperatures of onset of intense thermal degradation of
homopolymers synthesized from components 1, 2, , n.
In the reduced form, expression (VII.9) is the following:
k =n

Td =

k Vi

i
k
.

Vi

k =n
i

k T k
d, k
k =1
k =1

(VII.10)

In a different form, correlation (VII.7) looks as follows:


1
1
1
1
= 1
+ 2
+ ... + n
,
Td
Td,1
Td,2
Td, n
where

1 Vi

i
1
;
1 =
k =n

k Vi
k =1
i
k

(VII.11)

417

2 Vi

i
2
;
2 =
k =n

k Vi
k =1
i
k

n Vi

i
n
.
n =
k =n

k Vi
k =1
i
k

Chapter VIII. Optical and opto-mechanical properties of


polymers

VIII.1 Refractive index


The refractive index is the most important optical property of polymers. This
characteristic is directly associated with the dielectric constant of the substance. In the
general case, three molecular processes contribute to the statistic dielectric constant:
orientation of constant momenta in a field, the relative displacement of positive and
negative ions inside the molecule and the displacement of electrons in relation to
nuclei. These three processes describe orientational, atomic and electron polarization,
respectively.
In the area of sound frequencies, the dielectric constant is usually independent
of frequency. As frequency increases, orientational polarization falls behind the field
and finally has no effect on the dielectric constant. Usually, this transition proceeds in
the area of radio frequencies. At much higher frequencies, usually in the infrared
range, another type of transition is observed, when the frequency of influence
approaches self frequencies of ion oscillations or exceeds them. In the optical range,
the contribution of orientational polarization is negligible; the contribution of atomic
polarization may also be neglected. In this range, not the dielectric constant but
refractive index n is measured, the square of which equals under the condition that
both these values are determined at one and the same frequency. The refractive index
somewhat changes with frequency in the optical range which is associated with a
gradual approximation to self frequencies of oscillation of electrons existing in the
ultraviolet range.
The simplest method, in which an attempt is made to take into account
electrostatic interactions between molecules, is the method of the local Lorentz field.
The Lorentz calculation of the local field intensity is also valid at optical frequencies
under the same conditions, as in the statistic case, if the wavelength of the alternative
field is long compared with the lattice spacing. As applied to polymers, this means
that the wavelength must be long compared with the size of the elementary cell (or
approximately comparable with the average distance between atoms of neighboring
macromolecules).
A change of the refractive index n in dependence on density at the current
frequency with the Lorentz correction adheres to the following law:
n2 1 M

= R.
n2 + 2

(VIII.1)

This expression represents the LorenzLorentz formula, which is the optical


analogue to the ClausiusMossotti formula (see below).
In equation (VIII.1), M is the molecular mass (of the repeat unit, in the case of
polymers); is density; R is molecular refraction.
In the case of glassy polymers, the value of is calculated by the formula
(II.6), substitution of which into equation (VIII.1) gives

419

n2 1
n2 + 2

N A Vi
i

kavg

= R = mi Ai ri ,

(VIII.2)

where NA is the Avogadro number; kavg is the average coefficient of molecular


packing (for bulky monolith bodies, kavg = 0.681; for films, kavg = 0.695); mi is the
number of atoms of the i-th type in the repeat unit; ri is the specific refraction of
atoms of the i-th type; Ai is the atomic mass; Vi are increments of Van-der-Waals
volumes of atoms participating in the repeat unit of polymer.
Molecular refraction R is the additive value and is composed of refractions Ri
of separate atoms and constants for types of chemical bonds (double, triple). Some
values of Ri, required for calculations, are shown in Table 27 composed according to
the data by Eisenlore for the wavelength of 0.5893 m (D).
To calculate the refractive index n of copolymers, equation (VIII.2) has the
form:
n2 1
2

n +2

kavg (1R1 + 2 R2 + ... + n Rn )



N A 1 Vi + 2 Vi + ... + n Vi


i
1
i
2
i
n

(VIII.3)

where 1, 2, , n are the molar parts of the components 1, 2, , n; R1, R2, , Rn


are the molar refractions of homopolymers based on components 1, 2, , n;

Vi , Vi , , Vi are their Van-der-Waals volumes.

i
n
i
1 i
2
Table 27
Atomic refractions of series of atoms in organic compounds according to the data by Eisenlore
Atoms, atomic groups and features of structure
Symbol
RD, cm3/mol
Carbon
RC
2.418
Hydrogen
RH
1.100
Oxygen:
RO
In OH
1.525
In ethers *
1.643
RO<
In CO
2.211
RO=
Chlorine
RCl
5.967
Bromine
RBr
8.865
Iodine
RI
13.900
Double C=C bond
R=
1.733
2.398
Triple CC bond
R
Nitrogen:
In primary amines
2.322
R
H 2 NC

In secondary amines

RHN(C)2
R

2.502

N(C)3
In tertiary amines
2.840
RCN=C
In imides (tertiary)
3.776
RNC
In nitriles **
3.118
* As well as in -oxides, no constant being introduced for the tertiary oxide cycle.
** Values of nitrogen atomic refraction in imides and nitriles include constants for double and triple
carbonnitrogen bonds.

420

In the reduced form, correlation (VIII.3) may be written down as follows:


k =n

n2 1
n2 + 2

k avg
=

k =n

k Rk

k =1

N A k Vi

k =1
i
k

(VIII.4)

Equations (VIII.1) and (VIII.2) allow sufficiently accurate estimation of the


refractive index of polymers and copolymers on the basis of their chemical structure.
However, in the case of polymers and copolymers with low glass transition
temperatures, the value of n is often somewhat underestimated. This is associated with
the fact that the coefficient of molecular packing k for such systems is somewhat
lower than the average value kavg. That is why for more accurate estimation of the
refractive index, it is desirable to take into account the temperature dependence of k,
described by equations (II.14) and (II.15). Substitution of these dependences into
expression (VIII.2) gives
n2 1
n2 + 2

Rkg

[1 + G (T Tg )]N A Vi ,

(T < Tg);

(VIII.5)

(T > Tg),

(VIII.6)

n 1
2

n +2

Rkg

[1 + L (T Tg )]N A Vi ,
i

where kg = 0.667.
Consequently, using equation (VIII.4) for copolymers, we obtain:
k =n

n2 1
n2 + 2

kg
=

k Rk

k =1

k =n

1 + G T Tg N A k Vi

k =1
i
k

)]

, (T < Tg);

(VIII.7)

, (T > Tg).

(VIII.8)

k =n

n2 1
n2 + 2

kg
=

k Rk

k =1

k =n

1 + L T Tg N A k Vi

k =1
i
k

)]

Table 28 displays refractive indices for a series of amorphous polymers.


Clearly, the value of n depends on the chemical structure of the polymer and increases
at transition from aliphatic polymers to aromatic ones. It is also clear that for
20
polymers existing in the rubbery state at room temperature (T > Tg), values of nD
calculated with regard to the temperature dependence of the coefficient of molecular
packing, i.e. with application of expression (VIII.5), correlate much better with

421

20
experimental values of nD
. Therewith, values of the glass transition temperature Tg
determined by equation (IV.41) and values of the coefficients of volumetric expansion
G and L determined by correlations (III.8) and (III.5) were used for calculation.
Table 28
20
20
Molar refractions R, experimental n D, exp and calculated n D, calc values of the refractive
indices for a series of amorphous polymers
20
20
Ri , n D
nD
, exp
Polymer
, calc *
Tg,calc, K
i
cm3/mol
1
2
3
4
5
282
20.126
1.479
1.49
/
1.47
CH2 CH

C O CH3
O
CH

CH2

24.744

1.469

1.49 / 1.47

267

33.980

1.466

1.50 / 1.48

242

24.744

1.490

1.490

377

29.362

1.485

1.50 / 1.49

338

38.598

1.483

1.50 / 1.49

287

44.233

1.568

1.56

378

33.343

1.591

1.60

376

37.961

1.587

1.59

401

C O C2H5
O
CH2

CH
C O C4H9
O
CH3

CH2

C
C O CH3
O
CH3

CH2

C
C O C2H5
O

CH3
CH2

C
C O C4H9
O
CH3

CH2

C
C O
O

CH2

CH

CH3
CH2

422

CH2

CH

2
38.210

3
1.610

4
1.62

5
356

15.497

1.467

1.49 / 1.47

253

20.126

1.467

1.49

301

34.986

1.578

1.58

353

69.983

1.585

1.58

431

117.662

1.610

1.60

582

Cl

CH2

CH
O CH3

CH2

CH
O C

CH3

CH2

CH
O

CH3
O

O C

CH3
C

C O

O
O
C
O
C
O

* Column 4 indicates two values of ncalc: the first of them is determined by equation (VIII.2), and the
second by equation (VIII.6), i.e. with regard to the temperature dependence of the coefficient of
molecular packing k.

More comprehensive data on polymers, for which refractive indices were


measured, are shown in Table 28, the experimental data in which are borrowed from
Polymer Handbook [453].
Table 28
Average refractive indices of polymers
Polymer
1
Cellulose
Cellulose nitrate
Cellulose triacetate
Cellulose tripropionate
Ethyl cellulose
Methyl cellulose (low viscosity)
Natural rubber
Poly(acrolein)
Poly(acrylic acid)

n (exp)
2
1.54
1.5 1.514
1.47 1.48
1.48 1.49
1.479
1.497
1.519
1.52
1.529
1.527

n1 (calc)
3
1.49 (am);
1.54 (cryst)
1.52
1.48
1.49
1.47
1.48
1.51
1.51
1.49

423

1
Poly(acrylonitrile)
Poly(allyl mathacrylate)
Poly(N-allyl methacrylamide)
Poly(N-benzyl methacrylamide)
Poly(n-butyl methacrylate)
Poly(benzyl methacrylate)
Poly(bornyl methacrylate)
Poly(2-bromoethyl methacrylate)
Poly(p-bromophenyl methacrylate)
Poly(2-bromo-4-trifluoromethylstyrene)
Poly(1,2-butadiene)
Poly(1,3-butadiene)
Poly(butene) (isotactic)
Poly(butyl acrylate)
Poly(sec-butyl -bromoacrylate)
Poly(sec-butyl -chloroacrylate)
Poly(2-tert-butyl-1,3-butadiene)
Poly(tert-butyl methacrylate)
Poly(butylmercaptyl methacrylate)
Poly(N-butyl methacrylamide)
Poly(o-chlorobenzyl mathacrylate)
Poly(2-chloro-1-(chloromethyl)ethyl methacrylate)
Poly(2-chlorocyclohexyl methacrylate)
Poly(o-chlorodiphenylmethyl methacrylate)
Poly(2-chloroethyl -chloroacrylate)
Poly(2-chloroethyl methacrylate)
Poly(1-(o-chlorophenyl)ethyl methacrylate)
Poly(chloroprene)
Poly(o-chlorostyrene)
Poly(m-cresyl methacrylate)
Poly(o-cresyl methacrylate)
Poly(cyclohexyl -bromoacrylate)
Poly(cyclohexyl -chloroacrylate)
Poly(cyclohexyl -ethoxyacrylate)
Poly(cyclohexyl methacrylate)
Poly(p-cyclohexylphenyl methacrylate)
Poly(1-decene)
Poly(2-decyl-1,3-butadiene)
Poly(2,3-dibromopropyl methacrylate)
Poly(1,3-dibromopropyl methacrylate)
Poly(2,6-dichlorostyrene)
Poly(2-diethylaminoethyl methacrylate)
Poly(1,1-diethylpropyl methacrylate)
Poly(2,3-dimethylbutadiene) (methyl rubber)
Poly(1,2-diphenylethyl methacrylate)
Poly(diphenylmethyl methacrylate)
Poly(p-divinylbenzene)
Poly(dodecyl methacrylate)
Poly(2-ethoxyethyl acrylate)
Poly(2-ethoxyethyl methacrylate)
Poly(3-ethoxypropyl acrylate)
Poly(ethyl acrylate)
Poly(ethyl -chloroacrylate)

2
1.52
(1.5187)
1.5196
1.5476
1.5965
1.483
1.5680
1.5059
1.5426
1.5964
1.5
1.5000
1.5154
1.5125
1.4631
1.466
1.542
1.500
1.5060
1.4638
1.5390
1.5135
1.5823
1.5270
1.5179
1.6040
1.533
1.517
1.5624
1.5541.558
1.6098
1.5683
1.5707
1.542
1.532
1.4969
1.5066
1.5575
1.4730
1.4899
1.5739
1.5270
1.6248
1.5174
1.4889
1.525
1.5816
1.5933
1.6150
1.4740
1.471
1.4833
1.465
1.4685
1.502

3
1.49
1.51
1.53
1.57
1.49
1.56
1.53
1.53
1.58
1.55
1.53
1.51
1.50
1.48
1.53
1.52
1.51
1.50
1.56
1.52
1.57
1.53
1.53
1.59
1.53
1.52
1.56
1.54
1.62
1.55
1.55
1.55
1.53
1.51
1.52
1.56
1.49
1.50
1.56
1.53
1.63
1.52
1.51
1.51
1.58
1.58
1.61
1.49
1.46
1.48
1.47
1.47
1.51

424

1
Poly(ethyl methacrylate)
Poly(ethylene) (density 0.914 g/cm3)
(density 0.94 0.945 g/cm3)
(density 0.965 g/cm3)
Poly(ethylmercaptyl methacrylate)
Poly(2-fluoroethyl methacrylate)
Poly(4-fluoro-2-trifluoromethylstyrene)
Poly(2-(heptafluorobutoxy)ethyl acrylate)
Poly(heptafluorobutyl acrylate)
Poly(2-heptyl-1,3-butadiene)
Poly(hexadecyl methacrylate)
Poly(2,2,3,3,4,4-hexafluorobutyl acrylate)
Poly(n-hexyl methacrylate)
Poly(2-hydroxyethyl methacrylate)
Poly(iminoadipoyliminohexamethylene) (Nylon 6,6)
Poly(iminoadipoyliminotetramethylene)
Poly(imino(1-oxohexamethylene)) (Nylon 6)
Polyisobutene
Poly(isobutyl methacrylate)
Polyisoprene
Poly(2-isopropyl-1,3-butadiene)
Poly(isopropyl methacrylate)
Poly(p-isopropylstyrene)
Poly(methacrylonitrile)
Poly(p-methoxybenzyl methacrylate)
Poly(2-methoxyethyl acrylate)
Poly((N-2-methoxyethyl)methacrylamide)
Poly(4-methoxy-2-methylstyrene)
Poly(o-methoxyphenyl methacrylate)
Poly(methoxypropyl acrylate)
Poly(o-methoxystyrene)
Poly(p-methoxystyrene)
Poly(methyl acrylate)
Poly(methyl -bromoacrylate)
Poly(methyl -chloroacrylate)
Poly(methyl methacrylate)
Poly(N-methyl-methacrylamide)
Poly(4-methyl-1-pentene)
Poly(1-methylcyclohexyl methacrylate)
Poly(2-methylcyclohexyl methacrylate)
Poly(3-methylcyclohexyl methacrylate)
Poly(4-methylcyclohexyl methacrylate)
Poly(o-methylstyrene)
Poly(-naphthyl methacrylate)
Poly(-naphthyl methacrylate)
Poly(m-nitrobenzyl methacrylate)
Poly(2-nitro-2-methylpropyl methacrylate)
Poly(nonafluoropentyl acrylate)
Poly(octafluoropentyl acrylate)
Poly(oxycarbonyloxybis(1,4-(3,5-dichlorophenylene)))
Poly(oxycarbonyloxy1,4-(2,6-dichloro)phenylene-isopropylidene-1,4(2,6-dichloro)phenylene)
Poly(oxycarbonyloxy1,4-phenylene-1,3-dimethylbutylidene-1,4phenylene)
Poly(oxycarbonyloxy1,4-phenylene-1-methyl-butylidene-1,4-phenylene)

2
1.485
1.51
1.521.53
1.545
1.547
1.4768
1.46
1.390
1.367
1.5000
1.4750
1.392
1.4813
1.5119
1.53
1.53
1.53
1.5051.51
1.477
1.521
1.5028
1.4728
1.554
1.52
1.552
1.463
1.5246
1.5868
1.5705
1.471
1.5932
1.5967
1.472
1.480
1.5672
1.517
1.4893
1.5398
1.459
1.465
1.5111
1.5028
1.4947
1.4975
1.5874
1.6410
1.6298
1.5845
1.4868
1.360
1.380
1.6056
1.6056

3
1.50
1.49
1.52
1.54
1.54
1.47
1.51
1.40
1.39
1.50
1.50
1.41
1.49
1.49
1.52
1.52
1.52
1.49
1.50
1.51
1.51
1.50
1.58
1.49
1.54
1.46
1.51
1.57
1.54
1.46
1.57
1.57
1.46

1.5671

1.57

1.5745

1.57

1.54
1.51
1.49
1.52
1.51
1.52
1.52
1.52
1.52
1.60
1.59
1.59
1.57
1.53
1.37
1.38
1.61
1.60

425

1
Poly(oxycarbonyloxy1,4-phenylene-1-propylbutylidene-1,4-phenylene)
Poly(oxycarbonyloxy1,4-phenylene-sec-butylidene-1,4-phenylene)
Poly(oxycarbonyloxy1,4-phenylenebutylidene-1,4-phenylene)
Poly(oxycarbonyloxy1,4-phenylenecyclohexylidene-1,4-phenylene)
Poly(oxycarbonyloxy1,4-phenylenediphenyl-methylene-1,4-phenylene)
Poly(oxycarbonyloxy1,4-phenyleneethylidene-1,4-phenylene)
Poly(oxycarbonyloxy1,4-phenyleneisobutylidene-1,4-phenylene)
Poly(oxycarbonyloxy1,4-phenyleneisopropylidene-1,4-phenylene)
Poly(oxy-2,6-dimethylphenylene)
Poly(oxydimethylsilylene) (poly(dimethyl siloxane))
Poly(oxyethylene) (high molecular weight)
Poly(oxyethyleneoxymaleoyl) (poly(ethylene maleate)
Poly(oxyethyleneoxysuccinoyl) (poly(ethylene succinate))
Poly(oxyethyleneoxyterephthaloyl)
(amorphous)
(poly(ethylene
terephthalate))
Poly(oxymethylene)
Poly(oxy-1-oxopentamethylene)
Poly(oxypropylene)
Poly(pentachlorophenyl methacrylate)
Poly(pentadecafuorooctyl acrylate)
Poly(pentafluoropropyl acrylate)
Poly(pentafluorovinyl propionate)
Poly(1-phenyl-n-amyl methacrylate)
Poly(phenyl -bromoacrylate)
Poly(phenyl methacrylate)
Poly(1-phenylallyl methacrylate)
Poly(1-phenylcyclohexyl methacrylate)
Poly(1-phenylethyl methacrylate)
Poly(2-phenylethyl methacrylate)
Poly(N-(2-phenylethyl)methacrylate)
Poly(2-(phenylsulfonyl)ethyl methacrylate)
Poly(n-propyl methacrylate)
Poly(propylene) (density 0.9075 g/cm3)
Poly(styrene)
Poly(tetradecyl methacrylate)
Poly(tetrafluoro-3-(heptafluoropropoxy)propyl acrylate)
Poly(tetrafluoro-3-(pentafluoroethoxy)propyl acrylate)
Poly(tetrafluoro-3-(trifluoromethoxy)propyl acrylate)
Poly(2-(1,1,2,2-tetrafluoroethoxy)ethyl acrylate)
Poly(tetrafluoroethylene)
Poly(2,2,2-trifluoro-1-methylethyl methacrylate)
Poly(trifluorochloroethylene)
Poly(2-trifluoroethoxy)ethyl acrylate)
Poly(trifluoroethyl acrylate)
Poly(trifluoroethyl methacrylate)
Poly(trifluoroisopropyl methacrylate)
Poly(trifluorovinyl acetate)
Poly(3,3,5-trimethylcyclohexyl methacrylate)
Poly(undecafluorohexyl acrylate)
Poly(vinyl acetate)
Poly(vinyl alcohol)
Poly(vinyl benzoate)
Poly(vinyl butyl ether)
Poly(viny sec-butyl ether) (isotactic)
Poly(vinyl butyral)
Poly(vinyl chloride)

2
1.5602
1.5827
1.5792
1.5900
1.6539
1.5937
1.5702
1.5850
1.575
1.43
1.4563
1.4840
1.4744
1.5750

3
1.57
1.57
1.57
1.58
1.61
1.58
1.57
1.58
1.59
1.41
1.45
1.46
1.46
1.54

1.48
1.465
1.4495
1.608
1.339
1.385
1.364
1.5396
1.612
1.5706
1.5573
1.5645
1.5487
1.5592
1.5857
1.5682
1.484
1.5030
1.59
1.592
1.4746
1.346
1.348
1.360
1.412
1.35 1.38
1.4185
1.42 1.43
1.419
1.407
1.437
1.4177
1.375
1.485
1.356
1.4665
1.49 1.53
1.5775
1.4563
1.4740
1.48 1.49
1.54 1.55

1.42
1.47
1.46
1.61
1.35
1.39
1.37
1.55
1.59
1.56
1.56
1.56
1.55
1.55
1.57
1.56
1.50
1.51
1.60
1.49
1.36
1.36
1.37
1.41
1.31
1.45
1.42
1.42
1.41
1.44
1.45
1.40
1.52
1.36
1.49
1.50
1.58
1.48
1.48
1.50
1.57

426

1
Poly(vinyl chloroacetate)
Poly(vinyl decyl ether)
Poly(vinyl dodecyl ether)
Poly(vinyl ethyl ether)
Poly(vinyl-2-ethylhexyl ether)
Poly(vinyl formal)
Poly(vinyl hexyl ether)
Poly(vinyl isobutyl ether)
Poly(vinyl methyl ether)
Poly(vinyl methyl ether) (isotactic)
Poly(vinyl octyl ether)
Poly(vinyl pentyl ether)
Poly(vinyl phenyl sulfide)
Poly(vinyl propionate)
Poly(vinylcarbazole)
Poly(vinylfuran)
Poly(vinylidene chloride)
Poly(vinylidene fluoride)
Poly(vinylnaphthalene)
Poly(1-vinyl-2-pyrrolidone)
Poly(N-vinylphthalimide)
Poly(2-vinyltetrahydrofuran)
Poly(2-vinylthiophene)

2
1.512
1.4628
1.4640
1.4540
1.4626
1.50
1.4591
1.4507
1.467
1.4700
1.4613
1.4581
1.6568
1.4665
1.683
1.55
1.60 1.63
1.42
1.6818
1.53
1.6200
1.55
1.6376

3
1.54
1.48
1.48
1.47
1.49
1.50
1.48
1.48
1.47
1.47
1.48
1.48
1.67
1.48
1.64
1.56
1.57
1.38
1.63
1.55
1.60
1.53
1.65

VIII.2 Stress-optical coefficient


Polymeric materials have been used widely in the photoelasticity method of
investigation of stresses. This method is based on the phenomenon of birefringence,
found by D. Brster, which appears in transparent optically isotropic materials under
deformation and is stipulated by anisotropy of the refraction index in the mutually
perpendicular directions.
In the case of glassy polymers, the first consequence of applied load is the
variation of interatomic distances and valence angles in a polymeric chain. These
changes determine the instaneous elastic deformation. Elastic deformation is
associated with the mobility of atoms composing chains of macromolecules within a
statistical segment of a macrochain. Occurrence of the birefringence and its value at
deformation of polymers existing in the glassy state are generally stipulated by the
dislocation of electronic shells of atoms and electronic clouds forming chemical
bonds, as well as by distortion of the valence angles that leads to anisotropy of
polarizability of elementary units of macromolecules.
In the glassy state, birefringence may also be associated sometimes with the
elastic orientation of optically anisotropic macromolecules or their parts (for example,
mobile side methyl groups in polyacrylates and fluoride groups in poly(methacrylate)
esters near their equilibrium state. Therewith, the so-called elastic component of
birefringence occurs, which reaches its maximum almost immediately after the load
application. In the case of the ideal elastic body, total birefringence would be
controlled by elastic deformation, because under these conditions elastically deformed
polymer would exist in the equilibrium state. However, it should be noted that the
behavior of real polymeric substances differs from the elastic one. The change of
deformation and values of birefringence with time is typical of them even in the
glassy state.

427

For optically sensitive polymeric materials that are in the glassy state, the
experimental Wertheim law is applicable up to certain stress levels. This law connects
the optical retardation at any point of the polymeric model existing in the planestress state with the difference of the main principal stresses 1 and 2 acting in the
plane model in the same point, and thickness of the model d:

or

= C(1 2)d

(VIII.9)

n = n1 n2 = C(1 2)d,

(VIII.10)

where C is the stress-optical coefficient; n is the birefringence; n1 and n2 are


refractive indices along and perpendicular to the optical axis.
Introducing the light wavelength into equation (VIII.10), a correlation is
obtained which connects birefringence with the fringe order:
C d
m = ( 1 2 ) ,

(VIII.11)

where m is the fringe order for a given wavelength .


Coefficient C is usually considered to be independent of the wavelength, but
as more accurate investigation has indicated, it is not true. In practice of the
photoelasticity method, to characterize he optical sensitivity of polymeric materials,
another value the stress-related value of material strip 10.0 is used. This value
equals

10.0 =

.
C d

(VIII.12)

In the case of a linear dependence between stress and strain, birefringence may
be expressed via strain. Then, the dependence is reduced to the following form:

= C(1 2)d,

(VIII.13)

where C is the strain-optical coefficient of the material; 1 and 2 are the principal
strains.
Both optical coefficients are connected with each other by the correlation:
C E

C = =
,
1
1 + 0.0

(VIII.14)

where is the Poisson ratio; E is the elasticity modulus; 10.0 is the strain-related
value of material strip.
Since the value of birefringence of a polymer in the glassy state, in a definite
range, is proportional to the anisotropy of polarizability of the elementary unit bonds,
the increase of polarizability anisotropy of macromolecules and, consequently, the
optical sensitivity of the polymer may be achieved by introduction into the starting
monomer or oligomer of groups with high polarizability anisotropy (such as aromatic

428

cycles of types of benzene, naphthalene, anthracene, carbonyl groups and any


molecular groupings containing double or triple bonds, i.e. groupings containing
mobile -electrons) to the molecule.
At the present time, there exist two approaches to numerical estimation of the
stress-optical coefficient C for polymers based on the chemical structure of the
repeating unit of them. The first approach suggested in ref. [36] is empirical.
On the basis of numerous experimental data, treated in the cited reference, a
correlation has been suggested that connects C with the parameters of the chemical
structure of the repeat unit of the polymer:

C =

Ci
i

N A Vi

+,

(VIII.15)

where Ci are constants which characterize contributions of every atom and type of
intermolecular interaction into the stress-optical coefficient (Table 29); Vi is the
i

Van-der-Waals volume of the repeat unit composed of volumes of atoms participating


in this unit; NA is the Avogadro number; = 0.3544104 cm2/kG is the universal
parameter.
Table 29
Values Ci characterizing contributions of each atom and type of intermolecular interaction to the
stress-optical coefficient
Atom or type of intermolecular interaction
Ci103,
Symbol
MPa1 cm3/mol
CC
Carbon
2.0492
CH
Hydrogen
0.5227
CO,b
Oxygen in the backbone
3.1980
CO,s
Oxygen in the side group
0.7568
Nitrogen in the backbone
7.1750
C,b
Nitrogen in the side group
1.3030
C,s
CCl
Chlorine
3.4760
CS
Sulfur
0.7900
Cd
Dipoledipole interaction *
1.6000
Ch
Hydrogen bond
6.2100
p-Substitution of aromatic cycles **
Cp
1.7000
* Coefficient Cd is applied to every group of any chemical nature; if two identical groups locate at the
same atom, a single coefficient Cd must be introduced. For phenyl group, Cd = 2.15103.
** Coefficient Cp is introduced in the case of p-substitution of aromatic cycles; the number of
coefficients Cp introduced equals to the number of aromatic cycles substituted in the p-position.

Table 30 shows values of stress-optical coefficient C for a series of polymers


in the glassy state. The value of C changes is extremely wide range in relation to
chemical structure of polymer from low negative values for poly(methyl
methacrylate) and poly--methylstyrene to extremely high positive values typical of
aromatic polymers. High optical sensitivity of these polymers is indicated by their
structure: the presence of a large number of condensed cycles characterized by a
significant anisotropy of polarizability. As follows from Table 30, compounds
saturated by nitrogen and sulfur atoms and aromatic cycles possess the highest values

429

of C. A significant influence on the optical sensitivity is also caused by the type of


substitution of aromatic cycles. Para-substitution promotes increase of C.
The presence of a large number of polar C=O-groups leads to a decrease of
optical sensitivity. Possessing high negative anisotropy of polarizability, this group
decreases the total positive effect. The fact that the C=O-group is the carrier of the
negative effect has been observed in ref. [206] and confirmed in ref. [100] in the study
of the opto-mechanical properties of polymers displayed in Table 30.
Table 30
Experimental and calculated values of stress-optical coefficients C for a series of glassy polymers
Polymer
C106,
C106,
Mpa-1 exper. MPa-1 calc.
1
2
3
10.7
10.65

CH2

CH

CH3
CH2

3.3

2.80

2.0

4.57

15.7

22.30

7.3

14.80

22.9

16.70

C
C

CH3

CH3
CH2

CH2

CH

CH3

CH2

CH
Cl

Cl
CH2

CH

Cl

430

CH3
CH2

2
24.5

3
10.10

57.5

43.9

111.0

105.0

161.0

160.0

90

77.3

150.0

143.0

93

83.0

C
C O CH2
O

CH2

CH
N

CH3
O

O C

CH3

C
C

C
C

C HN

NH
C

O
C
O

C
C

C
O

S
O

C
C

C O
O

O
C
O
C
O

431

C NH

2
77.4

3
66.4

62.0

60.1

NH
C

C NH

NH
C

Using equation (VIII.15), the contribution of each group to the stress-optical


coefficient may be estimated quantitatively. This contribution is characterized by the
value of Ci N A Vi , with the higher value of C corresponding to the lower
i

absolute value of Ci
i

N A Vi , because this relation is negative. A sequence of


i

separate groups is formed due to the contribution

Ci
i

N A Vi 107 to the stressi

optical coefficient:

C O
O

O;

145.2

157.5

166.8

C
;

223.4
C

323.3

336.0

C
;

315.9

O
345.7

348.3

486.4

The calculation scheme considered for determination the stress-optical


coefficient enables the expected optical sensitivity of the polymer to be estimated and
is of a definite meaning for synthesis of polymers applicable in the photoelasticity
method of investigation of stresses.
For copolymers, equation (VIII.15) obtains the following form:

432

1 Ci + 2 Ci + ... + n Ci

i
1
i
2
i
n
+ ,
C =


N A 1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n

(VIII.16)

where 1, 2, , n are the molar parts of the components 1, 2, , n; Ci ,

i
1

Ci , , Ci are series of constants for the components 1, 2, , n;

i
2
i
n

Vi , Vi , , Vi are Van-der-Waals volumes of the

i
n
i
1 i
2
components 1, 2, , n; NA is the Avogadro number.
In the reduced form, equation (VIII.16) obtains the form:
k =n

C =

k Ci

k =1 i
k
+ ,
k = n

N A k Vi

k =1 i
k

(VIII.17)

where k, Ci and Vi are the molar part, the series of constants and the

i
k
i
k
Van-der-Waals volume of the k-th element, respectively.
If it is desired to express the stress-optical coefficient for copolymers via
stress-optical coefficients of the components, equation (VIII.16) changes to:

1 (C ,1 ) Vi + 2 (C ,2 ) Vi + ... +

i
1
i
2

C =

1 Vi + 2 Vi + ... +

i
1
i
2

+ n C , n Vi

i
n

+ n Vi

i
n

(VIII.18)

where C,1, C,2, , C,n are stress-optical coefficient of the components 1, 2, , n.


In the reduced form, equation (VIII.18) obtains the following form:

433

k =n

C =

k (C , k ) Vi

k =1

k =n

k Vi
k =1
i
k

(VIII.19)

where C,k is the stress-optical coefficient of the k-th component.


Let us now consider a semi-empirical method of estimation of the stressoptical coefficient C suggested in ref. [91]. The matter is that the calculation scheme
for determination of C, described above, which enables C of polymer to be
determined with high accuracy based on chemical structure of the repeating unit, does
not link C with other opto-mechanical thermal parameters (elasticity modulus, glass
transition temperature, etc.). Let us first perform general analysis.
Let us consider a thin plate lying in the XOY plane. Because the light
wavelength is much less than transverse sizes of the sample, it may be considered as
an infinite plate. The light beam falls along the Z axis. Since the plate is thin, then
according to [86]
U ZZ =

(U XX + U YY ) ,
1

(VIII.20)

where UZZ, UXX and UYY are displacements along the corresponding axes; is the
Poisson ratio.
When the plate is undistorted, its material represents an isotropic dielectric
with the dielectric constant 0. When the plate is deformed, the optical symmetry of
the medium changes and, consequently, the substance becomes optically anisotropic
and may be described by introducing dielectric constant tensor ik. According to [85]

ik = 0ik + a1Uik + a2Ullik,

(VIII.21)

where Uik is the strain tensor; a1 and a2 are opto-elastic constants.


We shall be interested in the value of ZZ which, in accordance with (VIII.21),
may be presented in the following form:

ZZ = 0 a1

(U XX + U YY ) + a2 1 2 U XX + 1 2 U YY =
1
1
1

1 2

(U XX + U YY ).
= 0 + a2
a1

1
1

(VIII.22)

Let YY = 0 (uniaxial deformation). Then UXX = UYY and

ZZ = 0 + [a2(1 2) a1]UXX;
E
U XX 2U XX = EU XX ,
XX =
1 2

(VIII.23)
(VIII.24)

434

where E is the elasticity modulus.


From the system of equations (VIII.23) and (VIII.24), the dependence of ZZ
on YY may be obtained:

ZZ = 0 +

1
[(1 2 )a2 a1 ] XX .
E

(VIII.25)

2
, and 0 by n02 (where nZZ and n0 are refractive indices
Substituting ZZ by nZZ
in the Z direction and of isotropic substance, respectively), the stress-optical
coefficient C may be found from expression (VIII.25) which, according to the
definition, will be equal:

C =

nZZ
1
=
[(1 2 )a2 a1 ].
XX = 0 2n0 E
XX

(VIII.26)

Using for a1 and a2 the expressions deduced in ref. [50], according to which

1
a1 = 6 ( 0 1)( 0 + 2 ) ;

T 3

2
a2 = 3 + ( 0 1)( 0 + 2 ),

T 3
where is the polymer density, we obtain the final expression:
C =

)(


1 2
2
2
( + 1) n0 + 2 n0 1 3 .
2n0 E 3
T

(VIII.27)
(VIII.28)

(VIII.29)


Let us estimate the term by equation (VIII.29). Let us rewrite
T

as

T
=
.

T T T

(VIII.30)

When T << Tg, it may be assumed that in a definite temperature range the

= const < 0. Consequently,


coefficient of volumetric expansion G = const, and

expression (VIII.29) may be reduced to the following form:


C =


1
+ * ,
2n0 G E T

which is an analogue of equation (VIII.15). In expression (VIII.31)

(VIII.31)

435

* =

(1 + ) ( n02 1) ( n02 + 2 )
3n0 E

where is the Poisson ratio; n0 is the refractive index; E is the elasticity modulus.
Because according to the above-said, in the case of amorphous polymers,
0.0962
(see Section IV.3), and for isotropic dielectric, according to ref. [85],
G =
Tg

i Vi

0 = i

Vi

, expression (VIII.31) obtains the final form:

C =

15.6Tg
n0 E

Ci Vi

Vi

+ * ,

(VIII.32)

where Ci = /T; Vi are the increments of Van-der-Waals volumes of atoms.


Values of Ci are practically independent of temperature and stress, as well as
change weakly with temperature and stress in the optical range of frequencies [57,
106].
To determine C, we will base on the following reasons. Because * is the
order zero value, and C is the first order value by smallness, Ci may be presented as
Ci = Ci,0 + Ci, where Ci,0 is the order zero component, and Ci is the component of
the first order by smallness. Then determination of C from expression (VIII.32) will
be reduced to solution of the system of equations

Ci Vi
i

Vi

+ = 0 ;

Ci Vi
i

Vi

= C ,

(VIII.33)

* n0 E
C n E
where C = 0 ; =
.
15.6Tg
15.6Tg
The second equation in the system (VIII.33) is used in ref. [91] for calculation
of constants Ci for various atoms and types of intermolecular interaction. Calibration
of the method was performed on the basis of experimental values of C, n0, E and Tg
for a series of amorphous polymers studied well by solving the excessive system of
equations, composed on the basis of correlation (VIII.32). Consequently, a selection
of constants Ci shown in Table 31 was obtained. Knowing these constants, the value
of C may be determined first from expression (VIII.33), and then C may be
calculated from the correlation
C =

C 15.6Tg
n0 E

(VIII.34)

436

Table 31

Values of constants Ci for different atoms and types of intermolecular interaction


Element or type of molecular interaction
Designations
Ci106, MPa-1
Carbon
0.005
CC
Hydrogen
0.118
CH
Oxygen in the backbone
2.660
CO,b
Oxygen in the side group
0.700
CO,s
Nitrogen in the backbone
16.620
CN,b
Nitrogen in the side group
0.640
C,s
Sulfur
0.740
CS
Dipoledipole interaction
6.470
Cd
Hydrogen bond
66.040
Ch
Para-substitution
0.730
Cp
Metha-substitution
3.410
Cm

Let us perform this analysis in detail for poly(phenyl quinoxoline)

C
C

N
N

N
N

C
C

which possesses the maximal stress-optical coefficient. For this polymer,

Ci Vi

= CC(4VC* + 4VC,21 + 4VC,19 + 2VC,20 + 20VC,18) + CH,20VH,124 +

CO,bVO,131 + 4CN,sVN,151 + 2Cd = 0.005(411.1 + 410.2 + 48.4 + 211.6 +


2012.7) + (0.118)202.0 + 2.662.1 + 40.646.1 + 26.47 = 27.4 MPa1 cm3/mol;
Vi = 462.5 3.
i

27.4
= 0.0592.
462.5
Substituting the value of C , Tg = 563 K, n0 = 1.62 and E = 1900 MPa into
expression (VIII.34), we obtain
Then C =

C =

0.0592 15.6 563 3


10 = 169 B.
1.62 1900

The semi-empirical approach to estimation of the stress-optical coefficient C


of polymers, developed in ref. [91], also enables the elasticity modulus of the linear
glassy polymer to be calculated. Using correlation (VIII.34) for the determination of
C , it may be indicated that

437

Table 32
Opto-mechanical properties of a series of heat-resistant aromatic polymers
Structure of the repeating unit
Tg, K
C106, MPa-1
C Vi ,
Calc.
Exper.
i
MPa-1cm3/mol
1
2
3
4
5
563
27.4
169.0
161.1

E, MPa
Calc.
Exper.

n
Calc.

6
1810

7
1900

8
1.62

3120

3100

1.62

2017.0

120.7

111.2

N
C

788

O
C
437

438

C O

2
593

3
12.940

4
88.1

5
92.7

6
2200

7
2190

8
1.60

543

10.600

72.8

72.8

1640

1630

1.63

630

7.813

80.2

77.4

1260

1260

1.64

C
O
C
O
O

O
O

C
O

C
O

C
O

NH

NH
C

438

439

Ci Vi 15.6T
g

C 15.6Tg i

=
E=
.
C n0

Vi C n

0
i

(VIII.35)

Substituting equation (IV.41) into correlation (VIII.35) for calculation of Tg of


linear polymer, finally, we obtain:

Ci Vi 15.6

E=
.

C n0 ai Vi + b j

j
i

(VIII.36)

The results of calculation of E for representatives of different classes of heatresistant polymers are shown in Table 32. It should be noted that the elasticity moduli
of glassy polymers at temperatures below Tg differ insignificantly from each other (for
example, two-fold difference must not be considered large, because it may be
indicated as a result of tests at different deformation rates, for samples of different
shape, for samples of the same polymer with different prehistory of production, etc.).
That is why we must proceed with caution when estimating the elasticity modulus of
linear glassy polymers.
The stress-optical coefficient C is a fundamental characteristic of materials
applied to the photoelasticity method of stress investigation. According to this
method, a model of the full-scale construction is prepared from transparent optically
sensitive polymers to which appropriate loads are the applied. Consequently, a
birefringence appears in the material of the model, and one may evaluate the stress
strain state on the basis of the construction of the pattern of fringes.
In the case of bulky models, the method of deformation freezing is quite
efficient. The essence of this method is that the model is heated up to a temperature at
which the model material transits into the rubbery state. The model is then loaded and
cooled under load down to room temperature, which for usual stress-optical materials
is approximately 80100C below the glass transition temperature. Strains occurring
at loading of the model and optical anisotropy are frozen. Further on, the model is
sawn into thin plates which are then studied.
Paying no attention to other varieties of the photoelasticity method of stress
investigation, let us note that successful development of these methods is possible
only by creation of new polymeric materials with the required opto-mechanical
properties. Solution of various problems requires materials of two types, elastic and
viscoelastic. Investigation of stress fields in stratified elastic media by the method of
dynamic photoelasticity requires polymeric materials with different moduli possessing
the ratio of elasticity moduli from 2 to 10 and elastic properties at both static and
dynamic loading. Such materials have appeared recently in the practice of dynamic
photoelasticity [45].

440

Figure 71. Stress relaxation for specimens ED-20 + MTHPA (1), polyisocyanurate network (2) and
viscoelastic material (3) the composition of which is indicated in Figure 72

A difficulty in obtaining such materials is that all polymeric glasses


independently of the chemical structure possess nearly identical elasticity moduli of
~103 MPa. Seemingly, to obtain materials with different moduli, polymers may be
synthesized which at the experimental temperature (room, for example) would exist in
the zone of transition from the glassy state to the rubbery state. Because the elasticity
modulus decreases abruptly in this zone, a material can always be selected possessing
an acceptable elasticity modulus. However, materials in the transition zone possess
extremely viscoelastic behavior, whereas the photoelasticity method of stress
investigation require elastic materials for solving this task. These two incompatible
requirements may be fulfilled in polymer networks of a definite chemical structure by
synthesizing high cross-linked networks which, possessing extremely high rubbery
modulus, would possess low glass transition temperatures (significantly below the
room temperature). The current state of the problem in the sphere of prognosis of
properties of such systems enabled the structure of high cross-linked networks to be
predicted, which would fulfill these two conflicting requirements [46, 47].
When the above-considered approach is used, the glass transition temperature
and the elasticity modulus are predicted, and network carbofunctional organosilicon
polyisocyanurates with various but short lengths of the linear fragment were chosen as
a particular object of investigation [45, 46]:
R
CH3
O C

CH3

N
C

C O

CH3

O
R

where
CH3
R:

NH C O CH2
O

CH2

O CH2

Si

CH3
O

CH3 n

Si CH2
CH3

O CH2

CH2

O C NH
O

441

A significant property of these materials is that despite the values of moduli


typical of the zone of transition from the glassy state to the rubbery state, they display
rubbery, as glass or rubber, but not viscoelastic behavior as usual materials existing in
the transition zone.
Let us compare the relaxation behavior of the given materials and materials
(elastic and viscoelastic), which are usually applied to the photoelasticity method
[47]: epoxy oligomer ED-20 cured by polysebacic acid anhydride (viscoelastic
material) and oligomer ED-20 cured by methyltetrahydrophthalic anhydride (Tg =
115C, elastic material). Existence of the transition zone (from the glassy state into
the rubbery state) of a viscoelastic material in the temperature range from 5 to 34C
enables, changing the test temperature, comparison of the relaxation properties of
these polymers to be performed at identical values of the initial elasticity modulus.
Analyzing the data shown in Figure 71, it may be concluded that the curve of
stress relaxation for a polyisocyanurate network is similar to the relaxation curve of a
glassy polymer. The absolute value of stress decrease, when the part of low relaxation
rate for the sample of network polyisocyanurate is reached, is approximately the same
as for the glassy material (ED-20 + MTHPA) and significantly lower than for a
viscoelastic polymer.
Figure 72 indicates dependences of mechanical I(t) = (t)/ and optical D(t) =
m(t)/d creep on time: m(t) is the order interference fringes at the moment of
measurement; is the stress in the sample; d is the sample thickness in the
translucence direction.
Figure 72b indicates that the compliance of the viscoelastic material smoothly
increase after loading. However, in contrast to a typical viscoelastic polymer, for
silicon-containing polyisocyanurate networks (see Figure 72b) the creep process
rapidly attenuates.

Figure 72. Mechanical I(t) and optical D(t) creep curves of network polyisocyanurates at n = 6.2 (see
the structural formula in the text), = 1.4 MPa (a) and viscoelastic polymer based on epoxy
oligomer ED-20 cured with the mixture of polysebacic anhydride (27.3 %) in the presence of
ozelaic acid (13.8%) (b)

442

Hence it may be noted that the elastic behavior, complicated only by weak
signs of viscoelasticity, is typical of carbofunctional organosilicon polyisocyanurates
synthesized in Refs. [45, 46].
Table 33 indicates the values of equilibrium elasticity moduli, obtained by
approximation of relaxation curves of stresses attracting new relaxation memory
functions (see below) which, besides high correlation coefficients, yield a series of the
physical characteristics of the material (number of microdefects, initial entropy of the
system, etc.). This Table also indicates values of the stress-related value of material
strip 10.0 . As observed from the Table, stress-optical sensitivity n increases with the
value of the linear fragment in the sequence of polyisocyanurates differing by the
amount of dimethylsiloxane units between network cross-linked points. Obviously,
despite a decrease of the concentration of groups with high polarizability anisotropy
(isocyanurate cycle, aromatic rings), optical sensitivity increases due to an increase of
macrochain mobility, which depends on both the concentration of cross-linked points
in the network and the amount of dimethylsiloxane units possessing a low potential
energy of SiCSi bond rotation [52].
Table 33
Opto-mechanical properties of macrodiisocyanates networks
Static loading
Dynamic loading
N
1.2
2.5
3.8
5.4
6.2

E, MPa

1.0
O , MPa (at 22C)

Cp, m/s

Ed, MPa

1.0
O , MPa

866
283
149

43

1.26
1.14
1.08
0.80
0.69

1980
1700
1270
1000
800

4822
3454
1897
1140
727

2.6
2.3
2.1
1.8
1.6

The dynamic characteristics of the opto-mechanical properties of polymers


may differ significantly from statistic ones due to the influence of the time factor. For
example, under the effect of short-term pulse loads the processes associated with
recording of the optical pattern of fringes in the model last from several microseconds
to hundreds of microseconds. In this case, usual quasi-static creep and stress
relaxation tests may not reflect the essence of phenomena proceeding in the polymeric
material under dynamic influence.
Analysis of photograms indicates that longitudinal and transverse waves
spread in a rod from network polyisocyanurates at a constant rate. Values of the rate
of longitudinal wave spreading Cspr and dynamic elasticity modulus Ed calculated by
2
, where is the material density, are indicated in Table 33. It
the equation Ed = Cspr
is obvious that these materials have a wide range of values of the longitudinal wave
rate Cspr (500 2000 m/s) and dynamic elasticity modulus Ed (300 5000MPa).
Hence, optically sensitive materials based on network polyisocyanurates with
short chains between neighboring cross-linked points of the network differ
significantly from the traditionally used materials which have practically equal values
of the dynamic modulus different static elasticity modulus. This creates problems in
their application to solution of some dynamic tasks.
To estimate the viscoelastic properties of network polyisocyanurates under the
conditions of dynamic influence, let us consider pulses of fringe patterns m(t) in
different cross-sections l of rods. For comparison, values of pulses m(t) are depicted
on a single graphic (Figure 73) with a time displacement which takes into account the

443

rate of wave spreading in materials. In contrast to a typical viscoelastic polymer (see


Figure 73b), an insignificant change of the shape of pulses and their duration is
typical of samples of network polyisocyanurates with n = 1, , 9. This change is
observed at an increase of the distance passed in the rods and testifies their low
viscoelasticity under pulsed loading. Consequently, optically sensitive network
polyisocyanurates are suitable for the study of the stressstrain state of stratified
media by the dynamic photoelasticity method.

Figure 73. Change of impulses of pattern of fringes m(t) in different sections l of network
polyisocyanurates rods at n = 2.5 and l = 50 (1); 150 mm (3) (a) and viscoelastic polymer
at l = 60 (1); 100 (2); 160 mm (3) (b), the composition of which is indicated in Figure 72.

Considered above were elastic polymeric materials. However, to solve a series


of tasks, optically sensitive materials with viscoelasticity are required. Clearly, the
viscoelastic behavior is typical of the transition zone from the glassy to the rubbery
state.
The currently available photoelasticity methods of creep modeling (the
photocreep method) are based on a special selection of materials which, besides high
optical sensitivity, would possess a clearly expressed viscoelastic behavior. To solve
this problem, it is necessary to obtain polymer networks existing at test temperature in
the transition area from the glassy to the rubbery state.
Therewith, the viscoelasticity of materials increases as the test temperature
approaches glass transition temperature Tg. Therefrom, it is required to obtain a series
of polymeric materials with assigned Tg, differently spaced in relation to the test
temperature.
To predict the thermal and physical characteristics of newly synthesized epoxy
anhydridoacidic compounds used for solving the present task, the above-described
calculation scheme of determination of Tg was used for polymer networks in ref. [35].
To obtain optically sensitive polymers differing by their thermal and optomechanical indices, to synthesize model materials, epoxy oligomer ED-20, curing
agent methyltetrahydrophthalic anhydride, cocuring agents dicarboxylic azelaic
acid and monocarboxylic oleic acid in various stoichiometric ratios were used in order
to regulate the cross-linking frequency and the molecular mass of linear fragment Ms.

444

Figure 74. Dependence of the stress-related values of material strip reduced to the specific stress m/
on loading time t. Formula of compositions is: 1) ED-20 1 mole, methyltetrahydrophthalic
anhydride (MTHPA) 1.4 mole, oleic acid 0.6 mole; 2) ED-20 1 mole, MTHPA 0.2
mole, azelaic acid 0.9 mole; 3) ED-20 1 mole, MTHPA 1.6 mole, oleic acid 0.4 mole

Consequently, a set of polymer networks was obtained, the glass transition


temperatures of which are close to room temperature. Obviously, materials based on
these networks possessed clearly expressed viscoelasticity. Such polymers possess the
ability to creep in an extremely wide range of absolute values of compliance and
process rates. This also causes the clearly indicated time dependence of the stressrelated values of material strip reduced to specific stress (Figure 74).

Chapter IX. Dielectric constant of polymers and organic


solvents

Calculations of the dielectric constant of polymers from their chemical


structure is an important problem both from the point of view of directed synthesis of
polymers with the given dielectric constant and for estimation of the polarity
(magnetic momentum) of the repeat unit of the polymer that is also of a significant
meaning for predicting polymer solubility in organic solvents. That is why it is also
worthwhile to perform quantitative estimation of the dielectric constant for organic
liquids which are solvents for polymers. It must be noted immediately that the
problem of calculation of the dielectric constant of organic liquids is more
complicated than of polymers. This is associated with the circumstance that the
coefficient of molecular packing for amorphous polymers is approximately identical
and depends only slightly on the chemical structure of the polymer. As mentioned
above, in the first approximation, the coefficient of molecular packing for glassy
amorphous polymers at room temperature is estimated by the value 0.681. In more
accurate approximation, the coefficient of molecular packing is approximately
identical for all polymers at their glass transition temperatures Tg; this value is kg =
0.667. As shown below, this enables accurate calculations of dielectric constant to
be performed for polymers at room temperature.
For organic liquids, their coefficient of molecular packing significantly
depends on the chemical structure; the coefficient of molecular packing for organic
liquids varies in the range from 0.45 to 0.705. This circumstance makes it difficult to
calculate the density (or molar volume) of a liquid, which is necessary for further
calculation of the dielectric constant.
It is common knowledge that the relationship between polarization vector P
and the vector of electric field intensity E in vacuum and in a dielectric is of the
following form: D = E + 4P = E, where D is the vector of electric field density. The
theory leads to the following expression for the dielectric constant in the case of nonpolar dielectrics:

= 1+

ni i
i

1 ni i i

where ni is the concentration of atoms, ions or molecules of the i-th type; is the
polarizability of these structural elements; i is the factor taking into account the
dipoledipole interaction.
It is common knowledge that the dielectric constant of a homogeneous
dielectric indicates how many times the value of electric field intensity E decreases
inside the dielectric compared with the value of external field E0:
E
E= 0.

446

Macroscopic field E is the vector sum, E = E0 + Ei, of the external field E0 and
the field Ei stipulated by the substance polarization P in the external field:
Ei = 4P.
In weak fields, polarization is proportional to the macroscopic field P = E,
where is the macroscopic susceptibility of the substance.
Therefrom, with regard to the above-considered equations, we obtain the wellknown connection of macroscopic characteristics:

= 1 + 4.

(IX.1)

At microscopic description, the value of polarization is determined by


microscopic polarizability 0 = ni i (ni is the concentration and i is the
i

polarizability of structural elements of the substance) and the value of local field Eloc:
P = 0Eloc,
where the local field equals
Eloc = E + P.
The coefficient of depolarization for isotropic dielectric equals to 4/3. A
connection between macroscopic and microscopic 0 polarizabilities may be
deduced from the above-considered correlations:

0
.
1 0

The formula connecting the dielectric constant with microscopic polarizability


(the ClausiusMossotti formula) may be also deduced:

1 4
4
=
0 =
ni i .
+2 3
3 i

(IX.2)

Electron polarizability el in the range of optical frequencies equals:

el =

f 0k
e2
,

2
2
m k 0k D

where D = 3.21015 rad/s is the frequency corresponding to D the sodium bend; 0,k
and f0,k are the frequency and strength of oscillators for the electron spectrum of the
substance associated with the 0 k transition. By the order of magnitude, electron
polarizability el is 1024 cm3.
The appropriate contribution in equation (IX.2) is of the following form:

447

4
ni ( el )i =

Ri ,
3 i
M i
where is density; M is molecular mass; Ri is the molar refraction of the i-th
structural element (in the case of polymers, calculation is performed per a repeat unit).
Orientational polarizability dip in weak fields associated with the rotation of
constant dipoles d, equals:

dip =

d2
.
3k BT

At room temperature, for dipoles d 1D the value of dip also equals


approximately 1024 cm3. The appropriate contribution to the ClausiusMossotti
formula, associated with the presence of polar groups in the structure, will be
presented in the following form:

ni dip i = M P j ,
3 i
j
where the sum by j is summed up by all polar groups of structural elements. Finally,
equation (IX.2) will have the following form:

1 M

= Ri + Pj .
+2
i
j

(IX.3)

In the case of polymers, calculation is performed per a repeat unit. Taking into
account that the density of polymers may be calculated by equation (II.6), we reduce
it to the following expression, suitable for calculation of the dielectric constant:
k avg P
1
=
.
+ 2 N A Vi

(IX.4)

The value of molar polarizability P is additive and is composed of


polarizabilities of atoms, as well as polarizability constants associated with the
presence of various types of chemical bonds (double, triple) and other features of the
structure of molecules. In this case, the situation is the same as with estimation of the
molar refraction. For non-polar dielectrics, the dielectric constant is stipulated by
deformation polarizability only and, according to the Maxwell correlation, is
practically coincident with the square of the refractive index in the area of high
frequencies n2. For these polymers (polyethylene, polytetrafluoroethylene,
polybutadiene, etc.), the molar refraction R is practically coincident with the molar
polarization P.
For polar dielectrics, the situation is more complicated. Under the effect of an
electric field the orientation of constant dipoles proceeds in them. These dipoles occur
at the expense of the presence of polar groups in the polymer, for example, such as

448

2+ &2 &22 &O 1+&2 1+&22 &1 HWF 7KLV FDXVHV

exceeding of refraction R by polarization P for these groups.


Let us analyze the correction, necessary to be introduced into refraction in
order to calculate polarizability of the molecule. For polar groups of various chemical
nature, polarizability is given as
Pi = Ri + Ri,

(IX.5)

where Ri is the molar refraction of this group; Ri is a correction associated with the
orientation of dipoles. Values of these corrections were calculated with the help of
linear regression analysis based on comparison of refractive indices and dielectric
constants of a great number of polar polymers. The calculation itself was performed
using formulae (IX.3) and (IX.3), the combination of which gave the following
expression:

kavg Ri + R j

i
j
1

.
=
N A Vi
+2

(IX.6)

Calculation results are indicated in Table 34, which gives numerical values of
Ri for various polar groups most often met in polymers. If these values are known,
dielectric constants for a wide range of organic polymers of various classes may be
calculated. The results of the calculations performed indicate good coincidence with
the experimental data [133, 214] (see Table 34a).
Table 34

2
Group

Values Ri for calculation of polarizability


Group
Ri, cm3/mol
3.557
C NH

Ri, cm3/mol
21.000

5.371

0.845

8.728

&O

3.900

17.085

&)

1.352

10.300

O
O C
O
O C O

2+
&N

3.500
5.464

N
O

To obtain more accurate values of the dielectric constant for polymers at room
temperature, it is desirable to take into account the temperature dependence of the

449

coefficient of molecular packing. This relates, first of all, to polymers existing in the
rubbery state at room temperature. According to ref. [128], the temperature
dependence k(T) for these polymers is described by the correlation:
Table 34a
Calculated and experimental values of the dielectric constant for a series of polymers
Polymer
Error, %
calc
exper
1
2
3
4
Polytetrafluoroethylene
1.98
2.00; 1.96;
1.0
2.01; 2.10
Poly(4-methyl-1-pentene)
2.27
2.13
6.6
Polypropylene
2.27
2.15; 2.20
3.2
Polyisobutylene
2.23
2.23
0.0
Poly(vinyl cyclohexane)
2.38
2.25
5.8
Poly(1-butene)
2.25
2.27
0.9
Polyethylene
2.23
2.20; 2.30
1.4
2.35
2.1
2.40
Poly(,,,-tetrafluoro-p-xylylene)
Polyisoprene
2.28
2.37
3.8
Poly(o-methyl styrene)
2.54
2.49
2.0
Poly(1,4-butadiene)
2.27
2.51
9.6
2.51
5.6
2.65
Poly(-vinyl naphthalene)
Polystyrene
2.57
2.55; 2.60; 2.50
0.8
2.57
1.2
2.54
Poly(-methyl styrene)
Poly(cyclohexyl methacrylate)
2.70
2.58
4.6
Polychlorotrifluoroethylene
2.70
2.80; 2.60
3.8
2.60
1.9
2.65
Poly(-vinyl naphthalene)
Poly[oxy(2,6-dimethyl-1,4-phenylene)]
2.77
2.75; 2.65; 2.60
0.7
Poly[1,1-cyclohexane bis(4-phenyl)carbonate]
2.97
2.60
14.2
Poly(p-xylylene)
2.58
2.65
2.6
Poly(p-chlorostyrene)
2.88
2.82; 2.63; 2.65
2.1
Polyvinylbutyral
2.70
2.69
0.4
Ethylcellulose
2.71
2.70
0.4
Poly(isobutyl methacrylate)
2.71
2.70
0.4
Poly(dimethyl siloxane)
2.75
2.75
0.0
Poly[oxy(2,6-diphenyl-1,4-phenylene)]
2.78
2.80
0.7
Poly(m-chlorostyrene)
2.88
2.80
2.8
Poly(n-butyl methacrylate)
2.64
2.82
6.4
Poly(vinylidene chloride)
2.87
2.90; 2.92; 2.85
0.7
Bisphenol-A-polycarbonate
3.11
3.00; 3.05; 2.90
2.0
Poly(N-vinyl carbazole)
2.69
2.90
7.2
Poly[1,1-ethane bis(4-phenyl)carbonate]
3.18
2.90
9.6
Poly(3,4-dichlorostyrene)
3.16
2.94
7.5
Poly(chloro-p-xylylene)
2.89
2.95
2.0
Poly(vinyl chloride)
3.14
3.15; 3.05; 2.95
0.3
Poly(1,4-cyclohexylidene dimethylene terephthalate)
2.94
3.00
2.0
Poly(ethyl methacrylate)
2.84
2.80; 3.00; 2.90
1.4
Poly(oxy-2,2-dichloromethyltrimethylene)
3.06
3.00
2.0
Poly(p-methoxy-o-chlorostyrene)
2.94
3.08
4.5
Poly(methyl methacrylate)
2.94
2.94; 3.15; 3.10
0.0
Poly[thio(p-phenylene)]
2.99
3.10
3.5
Polyoxymethylene
2.96
2.95; 2.85; 3.10
0.3
Poly(tetramethylene terephthalate)
3.09
3.10
0.3
3.20; 3.16; 3.10
1.9
3.26
Poly(ethyl -chloroacrylate)
Poly[4,4-isopropylidene diphenoxydi(4-phenylene)sulfone]
2.93
3.18
7.9
Poly(ether etherketone)
3.00
3.20
6.2
Poly(hexamethylene sebacamide)
3.60
3.80; 3.20;
5.3

450

Poly(vinyl acetate)

2
3.10

Poly(ethylene terephthalate)

3.29

Poly(p-hydroxybenzoate)
Poly[2,2-(m-phenylene-5,5-bibenzimidazole)]
Poly(methyl -chloroacrylate)
Poly[4,4-diphenoxydi(4-phenylene)sulfone]
Poly(hexamethylene adipamide)

3.25
2.72
3.47
3.54
4.13

Poly[N,N-(p,p-oxydiphenylene)pyromellitimide]
Poly[4,4-sulfondiphenoxydi(4-phenylene)sulfone]
Polyacrylonitrile

3.39
3.72
3.01

k (T ) =

kg

1 + 3.56 10 4 T Tg

3
3.02; 3.30;
3.20; 3.22; 3.25
3.40; 3.50;
3.10; 3.25
3.28
3.30
3.45; 3.32; 3.40
3.44
4.14; 4.10;
4.00; 3.50
3.50
3.80
3.26; 3.15;
3.10; 4.00

4
2.6
1.2
0.9
-17.6
0.6
2.9
0.2
3.1
2.1
2.9

; Tg < T.

At room temperature we obtain


k (298) =

kg

1 + 3.56 10 4 298 Tg

For copolymers, the equation to calculate the dielectric constant is presented in


the following form:

1
=
+2

k avg (1P1 + 2 P2 + ... + n Pn )



N A 1 Vi + 2 Vi + ... + n Vi


i
1
i
2
i
n

(IX.5)

where 1, 2, , n are the molar parts of the components 1, 2, , n of copolymer;

V , V , , V are the Van-der-Waals volumes of the same


i
i
i

i
1 i
2
i
n
components; P1, P2, , Pn are polarizabilities of the components 1, 2, , n.
In the reduced form, equation (IX.5) looks as follows:
k =n

1
=
+2

kavg
k =n

k Pk

k =1

N A k Vi

k =1 i
k

(IX.6)

451

where k and Vi are the molar part and the Van-der-Waals volume of the k-th

i
k
component, respectively; Pk is its polarizability.
It is desirable to express the dielectric constant of a copolymer via dielectric
constants of the components, then equation (IX.5) may be rewritten in the following
form:

1
=
+2

1
1
Vi + 2 2
Vi + ... +
1 1

1 + 2 i
2 + 2 i
1
2

1 Vi + 2 Vi + ... +

i
1
i
2

1
+ n n
Vi

n + 2 i

+ n Vi

i
n

(IX.7)

n,

where 1, 2, , n are dielectric constants of the components 1, 2, , n.


In the reduced form, equation (IX.7) becomes the following:
k =n

k k + 2 Vi

k
1 k =1
i
k
=
,
k =n
+2

k Vi
k =1 i
k

(IX.8)

where k is the molar part of the k-th element; Vi is the Van-der-Waals

i
k
volume of the k-th element.
To estimate the dielectric constant of organic liquids which are solvents of
polymers, let us now turn to the calculation scheme.
As mentioned above, the coefficient of molecular packing for organic liquids
depends significantly on the chemical structure and is not a constant value. That is
why calculation of the dielectric constant by formula (IX.3) is difficult, because the
density of the liquid can not be calculated with acceptable accuracy. However, this is
not the main reason why the dielectric constant of liquids may not be estimated with
the help of the ClausiusMossotti equation with acceptable accuracy. For example, if
the value is calculated for such a solvent as n-propyl alcohol and the value Ri is
assumed the same as for polymers, we will get the following numerical values: Ri =
3.3 cm3/mol, exper = 0.799 g/cm3, Vi = 70.65 3, Ri = 17.579 cm3/mol, P =
i

21.079 cm3/mol. Substituting these values into equation (IX.3), we obtain = 2.17,
which is by the order of magnitude lower than the experimental value, which equals

452

20.1. The same calculations have also been performed for a series of other liquids
and, therewith, the result is analogous to the former one: in all cases, the calculated
dielectric constant is significantly lower than the experimental one. Hence, even if it
is possible to calculate the density of a liquid with high accuracy, this would not lead
to proper values of the dielectric constant calculated by equation (IX.3).
It might be suggested that the value Ri for the same polar group contained in
polymers and low-molecular liquids must be different. The calculations performed
indicate that this is true not only for comparison of the behavior of organic liquids and
polymers, but also for comparison of liquids themselves related to the same class. For
example, contribution of OH-group to the value of Ri is different in the sequence of
alcohols and depends on the chemical structure of alcohol. In all cases, for liquids
related to the same class, the contribution of a polar group to the value Ri increases
with the Van-der-Waals volume of the liquid. Such analysis has been performed on
the basis of equation (IX.6), into which the average value of the coefficient of
molecular packing kavg for liquids of various classes is substituted, and values Ri
i

are calculated from Tables shown in refs. [28] and [128].


Table 34b
Dependence of Ri on the Van-der-Waals volume of liquids

Ri = f Vi
Class of liquids
i

Chlorinated compounds containing 1 carbon atom R 1 = 0.49 V + 43.8


Cl

Chlorinated compounds containing more than 1


carbon atom
Chlorinated compounds with a double bond at
chlorine atom
Alcohols

2
RCl
= 0.332 Vi + 45.52

Acids

Esters

RCOOH = 3.75 10 3 Vi 0.825 Vi +


i
i

+ 57.0
RCOO = 0.18 Vi + 20.2

Ketones

RCO = 0.525 Vi + 8.25

Ethers
Aldehydes

RO = 29.0

=
RCl
= 0.325 Vi + 33.5
i
2

ROH = 1.708 10 3 Vi + 0.761 Vi +


i
i

+ 0.343
2

i
i
2

Nitriles

RCOH = 1.67 10 3 Vi + 0.751 Vi +


i
i

+ 2.648
RCN = 0.525 Vi + 9.75

Nitrocompounds

R NO2 = 0.667 Vi + 3.64

i
i

Therewith, experimental values of the dielectric constant were used, and


values Ri for each polar group were calculated by equation (IX.4). A total of eleven

453

classes of organic liquids was analyzed, indicated in Table 34b. Therewith, for
chlorinated compounds, the influence of the chlorine atom on polarization, OH-group
for alcohols, COOH-group for acids, etc. was taken into account.
Figure 74 indicates the dependence of ROH on the Van-der-Waals volume of
an alcohol molecule. Clearly, all points fit well the generalized curve that may be
approximated with the help of a correlation
2

ROH = 1.708 10 3 Vi + 0.761 Vi + 0.343 .

i
i

Figure 74. Dependence of correction R to polarizability on the Van-der-Waals volume

Vi

for

OH-group

Correlations for calculation of Ri values for all remaining classes of liquids


were obtained in an analogous manner. These correlations are shown in Table 34b. If
they are known, contribution of each polar group to the value Ri may be calculated
easily. These calculations were performed for a multiplicity of organic liquids (Table
35). The calculations were performed with the help of correlations indicated in Table
34b; Van-der-Waals volumes and molar refractions were determined according to the
common procedure [28, 128]. The calculations performed have indicated quite good
coincidence with experimental values of , which could not be obtained by other
methods. Hence, it is possible to calculate the dielectric constant of polymers and
their solvents; this may be performed on the basis of the chemical structure of the
repeat unit of a polymer or a molecule of an organic liquid.
Table 35
Values of the Van-der-Waals volume, polarizability, calculated and experimental values of
dielectric constant for a series of organic liquids
P,
Vi , 3
Liquids
Error, %
calc
exper
i
cm3/mol
1
2
3
4
5
6
n-Pentane
97.40
25.29
1.932
1.844
4.80
Isopentane
97.40
25.29
1.932
1.843
4.80

454

1
n-Hexane
Cyclohexane
n-Heptane
n-Octane
n-Nonane
n-Decane
Benzene
Toluene
o-Xylene
m-Xylene
p-Xylene
Ethylbenzene
Naphthalene
Isopropylbenzene
Styrene
Chlorobenzene
Bromobenzene
Ethylbromide
1,2-Dibromoethane
Methyliodide
Ethyliodide
Cyclohexanol
Methanol
Ethanol
n-Propanol
n-Butanol
Isobutanol
n-Pentanol
n-Hexanol
n-Octanol
Acetic acid
Butyric acid
Isovaleric acid
n-Valeric acid
Methyl formate
Ethyl formate
Methyl acetate
Ethyl acetate
n-Propyl acetate
Ethyl propyonate
n-Butyl acetate
Acetone
Methyl ethylketone
Diethylketone
Cyclohexanone
Diethyl ether
Dipropyl ether
Diisopropyl ether
Dibutyl ether
Diamyl ether
Diisoamyl ether
Butylethyl ether
Ethanale
Propyonale
Butynale
Nitromethane

2
114.00
102.00
132.00
149.00
166.00
183.00
88.24
105.00
121.00
121.00
121.00
122.00
134.00
139.00
118.00
102.00
109.00
68.02
89.83
64.68
81.75
110.00
36.51
53.58
70.65
87.72
87.72
105.00
122.00
150.00
55.16
89.29
106.00
106.00
57.29
74.36
74.36
91.43
109.00
109.00
126.00
64.84
81.91
98.98
104.00
89.86
124.00
124.00
158.00
192.00
192.00
124.00
47.77
64.84
81.91
46.19

3
29.31
27.71
34.53
39.14
43.76
48.38
26.31
30.93
35.54
35.54
35.54
35.54
41.65
40.16
35.08
35.07
34.07
19.20
26.97
19.62
24.24
32.79
11.90
16.52
21.14
25.75
25.75
30.37
34.99
44.23
17.24
26.47
31.09
31.09
21.82
26.44
26.44
31.05
35.67
35.67
40.29
21.44
26.05
30.67
33.09
25.87

35.11

4
1.945
1.990
1.941
1.947
1.951
1.950
2.261
2.240
2.568
2.374
2.270
2.220
2.339
2.207
2.256
4.160
5.400
9.380
4.780
7.000
7.820
14.790
32.660
26.610
21.440
17.580
17.580
14.630
12.500
10.450
6.240
2.890
2.650
2.650
9.640
7.170
7.170
6.090
5.450
5.450
5.080
22.550
18.920
17.070
17.930
4.400
3.410
4.880
3.030
2.810
2.810
3.410
23.410
18.000
14.720
39.050

5
1.890
2.023
1.924
1.948
1.972
1.991
2.284
2.379 (25)
2.568
2.374
2.270
2.412
2.540 (85)
2.380
2.430 (25)
5.621
5.400
9.390
4.780
7.000
7.820
15.000
32.630
24.300
20.100
17.100
17.700
13.900
13.300
10.340
6.150
2.970
2.640
2.660
8.500
7.160
6.680
6.020
5.690
5.650
5.010
20.700
18.510
17.000
18.300
4.335
3.390 (26)
3.880
3.060 (25)
2.770 (25)
2.820
3.060 (25)
21.100
18.500
13.400
35.870 (30)

6
2.90
1.60
0.90
0.05
1.20
2.10
1.00
5.80
0.00
0.00
0.00
8.00
7.90
7.30
7.20
26.00
0.00
0.10
0.00
0.00
0.00
1.40
0.70
8.80
6.01
2.28
1.10
4.80
6.40
0.80
0.70
2.60
0.20
0.60
13.10
0.10
7.10
1.00
4.30
3.60
1.20
8.90
2.20
0.40
2.00
1.50
0.60
25.80
1.00
1.40
0.30
11.40
10.90
2.70
9.80
8.90

455

1
Nitroethane
1-Nitropropane
2-Nitropropane
Nitrile acetate
Nitrile propyonate
Nitrile butyrate
Nitrile pentanate
Nitrile isoheptanate

2
63.26
80.33
80.33
48.96
66.03
83.09
100.00
117.00

16.72
21.34
25.95
30.57
35.19

4
30.540
27.080
26.760
40.780
25.240
20.450
18.110
16.730

5
28.060 (30)
23.240 (30)
25.520 (30)
37.500
27.200
20.300
17.400
15.500

6
8.80
16.50
4.90
8.70
7.20
0.70
4.10
7.90

Chapter X. Equilibrium rubbery modulus for polymer


networks

X.1 Calculations of the equilibrium modulus


To estimate the equilibrium rubbery modulus E and molecular mass of an
linear fragment Mc in the case of elastomer networks in the case of fairly sparse
cross-links, the equation of the classic rubber elasticity theory is used:
E = 3

RT
,
Mc

(X.1)

where is the density of a cross-linked elastomer; R is the universal gas constant; T is


absolute temperature.
Application of equation (X.1) to high-crosslinked networks, the linear
fragment of which contains an extremely small number of units, down to 1 and even
lower, causes a substantial divergence between the experimental and calculated values
of E.
For equation (X.1) to be true for description of the properties of highcrosslinked networks, the so-called front-factor is introduced into it:
E = 3

RT
.
Mc

(X.2)

However, introduction of an unpredictable front-factor into equation (X.1)


does not improve the situation, because, comparing the calculated and experimental
values of E, we may only estimate this front-factor. In this connection, ref. [31]
indicates an attempt to obtain a generalized correlation for estimation of E and Mc
which is true both for sparse and high-crosslinked networks.
Let us perform a detailed analysis of the influence of a great number of
network cross-linked points on the equilibrium rubbery modulus. Preliminarily, it
should be noted that for sparse networks, the Van-der-Waals volume of cross-linked
points is extremely lower than the Van-der-Waals volume of linear fragments. That is
why it may be neglected when the compressibility of a network system is estimated.
In the case of high-crosslinked networks, it is impossible to do this, because the total
Van-der-Waals volume of cross-linked points is approximately identical to the total
Van-der-Waals volume of linear fragments and may even exceed it.
Considering an elastomer network as a system consisted of two subsystems
elastic and rotational-isomeric, let us analyze first the latter one. Ref. [28] indicates
that determination of the elasticity coefficient of the rotational-isomeric subsystem
requires knowing of the difference between the energies of rotational isomers, which
depends on the size of molecular defects lid in the polymer in the following manner:

()

2
Ei = 4 Di a 2 lid ,

(X.3)

457

where Di is the average energy of intermolecular interaction, typical of the atom of


this type; a is the Morse potential constant:

( R ) = D e a ( R r0 ) 1 1 ,

(X.4)

where D is the depth of the potential trough; r0 is the equilibrium distance between
atoms in harmonic approximation.

Figure 75. Cylinder of interaction of two hydrogen atoms (see text).

To determine the size of the molecular defect lid , let us consider a cylinder of
interaction (Figure 75), the notion of which was introduced in work [91]. Its volume
is composed from volumes of interacting atoms Vi and the volume of the defect
itself, characterized by iVi:
cryst
= 2Vi + icryst Vi ,
Vcyl
amph
= 2Vi + iamph Vi ,
Vcyl

(X.5)
(X.6)

amph
where Vi is the Van-der-Waals volume of the i-th atom; values icryst and i
participating in correlations (X.5) and (X.6), are discussed in detail in work [31];
cryst
is the volume of the cylinder of interaction for a crystalline
therewith, Vcyl
amph
polymer, and Vcyl is the same for an amorphous polymer.

Therefrom, the defect value equals


lid =

amph
cryst
= Vcyl
Vcyl
amph icryst Vi
= i
,
S cyl
2S cyl
2

(X.7)

where Vi is a part of the Van-der-Waals volume of the i-th atom which is overlapped
by the Van-der-Waals volume of the atom, chemically bonded to it; digit 2 in the
denominator appears due to the equality of the amplitude of the i-th atom jump-over
to a half of the defect size; Scyl is the cross-section of the cylinder of interaction.
Let us estimate the value of lid . For this purpose, let us consider the limiting
case when the base radius of the cylinder of interaction equals the Van-der-Waals
radius of the i-th atom. Then

458

1
amph
icryst Ri .
lid = i
3

(X.8)

For a hydrocarbon polymer, Ri = RH; the defect size is the constant value
which equal lid = 0.053RH, where RH is the Van-der-Waals radius of the hydrogen
atom, equal to 1.17 . Substituting the value of lid obtained into expression (X.3), we
obtain the value of hydrogen atom contribution to the difference of the energies of
rotational isomers: EH = 4.56 kJ/mol. The value found correlates by the order of
magnitude with spectroscopic data on the difference of energies of rotational isomers.
For further analysis, let us rewrite expression (X.3) in the following form:
2

ld
Ei = 4 Di a 2 Ri2 i ,
Ri

(X.9)

where Ri is the Van-der-Waals radius of the i-th atom; lid is the size of the defect
formed by this atom.
Ref. [28] indicates estimation of the value aRi; therewith, it was found that
aRi 6.
To estimate the contribution of the rotational-isomeric subsystem to the
temperature dependence of the elasticity modulus of the polymer in the transition
region and in the range of the rubbery state, values of lid for various atoms and types
of the intermolecular interaction should be found.
As the calculations performed have indicated [28], melting points of polymers
Tm and the energy of intermolecular interaction Di, included in expression (X.9), for
the atom of the present type depend on the fact whether it participates in the
composition of a group of atoms performing the hydrogen bond or dipoledipole
interaction. That is why when Ei is calculated by expression (X.9), the influence of
the types of specific effect mentioned will be generally displayed via the energy of
intermolecular interaction Di of the atom of this type. Therewith, as calculation
amph

indicates, the constants i


and icryst are independent of the influence of
hydrogen bonds and dipoledipole interaction.
For the case when the repeat unit consists of a selection of atoms of different
types, let us introduce the notion of efficient difference in the energies Eeff of
rotational isomers, which may be found as follows. Apply a correlation deduced in
ref. [28] to estimate the elasticity modulus of amorphous polymers:

E=

Vi
i

V S
il i
ii
i

(X.10)

where Vi is the Van-der-Waals volume of the i-th atom of a repeat unit; Si is the
Van-der-Waals surface of the i-th atom, through which the intermolecular interaction

459

occurs; i is the elasticity coefficient of the i-th atom bond; li is the characteristic size
of the bond (see Figure 75).
For the rubbery state, the denominator in expression (X.10) is equal to
Vi Si
l , where ri is the elasticity coefficient of the rotationalisomeric subsystem.
ri i
i
According to ref. [28], the value of ri is

ri =

RT

( )

2
qi lid

(X.11)

E
where qi = exp i ; Ei is the difference in the energies of rotational isomers.
RT
With regard to correlation (X.11), the denominator of equation (X.10) obtains the
form:
2
(
lid ) qi Si
Vi .

(X.12)

RTli

Let us introduce a value qeff so that it is determined from the condition:


2
d 2
(
) qeff Seff V .
leff
(
lid ) qi Si
Vi =

li

leff

(X.13)

Since leff, according to correlation (X.3), is calculated from the formula

(leffd )2 = 4DEeffa 2 ,
eff

d
= l, then condition (X.13) may be reduced
where Deff = RTm, and Seff = S and leff
to the following form:
2
(
lid ) qi Si
Vi

li

=
T =Tm

Eeff qeff S
4RTm a 2 l

Vsurf .

(X.14)

The left part of expression (X.14) is known, and S and l in the right part are
known, too. That is why the problem is reduced to obtaining a solution of equation of
the type b = xex, where x = Eeff/RTm, and b includes all the known components of
equation (X.14). On this basis, the influence of atoms of various types on Eeff may
be determined.
If the hydrogen atom is assumed to be the most typical one included in
polymers, and the value EH appropriate to it is taken, appearance in the structure of a
repeat unit from atoms of another type with Ei will cause a change of Eeff.

460

Therewith, this change depends on the weight coefficient of this atom in the repeat
unit. The value Eeff will either increase or decrease: Eeff > EH, if Ei > EH, and
Eeff < EH, if Ei < EH.
Let us now consider the influence of the cross-linking degree on the elasticity
modulus of a polymer network existing in the rubbery state. Representing the polymer
network as a mixture of linear fragments and cross-linked points, let us write down
that
1/Enet = /El.f. + (1 )/Ecrl.p.,
or
E net =

E l.f.

+ (1 )

El.f.
Ecrl.p.

(X.15)

where El.f. and Ecrl.p. are the moduli for linear fragments and cross-linked points,
respectively; is the concentration of linear fragments.
Because equation (X.15) was deduced by summing up compressibilities of
linear fragments of the polymeric chain and cross-linked points, consequently,

= Vi
Vi , where Vi is the Van-der-Waals volume of the

i
l.f. i
i
l.f.
linear fragment;

Vi

is the Van-der-Waals volume of the repeating fragment of

the polymer network. As regards the notion of the cross-linked point from the present
viewpoint, it was formulated above: the cross-linked point of a network is a group of
atoms consisted of the atom, from which branching starts, plus neighboring atoms,
chemically bonded to the former one, and their nearest substituents.

The value Vi
Vi may be reduced to the following form:

l.f. i
0

Vi
n Vi

i
l.f.
i
l.f.
=
,
0
Vi

i
n Vi + Vi

i
l.f. i
crl.p.

(X.16)

where Vi
is the Van-der-Waals volume of the repeat unit of the linear

i
l.f.
fragment; n is the number of repeat units per single cross-linked point.
Introducing designations = El.f./Ecrl.p.

and = Vi

i
crl.p.

Vi ,

i
l.f.

where El.f. is the elasticity modulus of the linear polymer; Ecrl.p. is the elasticity

461

modulus of a polymer network consisted of cross-linked points only; Vi


is

i
crl.p.
the Van-der-Waals volume of the network cross-linked point, we reduce to the
following correlation: = n/(n + ), and expression (X.15) is transformed to the form:
El.f.
n

= + (1 ) =
+
.
E net
n+ n+

(X.17)

Since the polymer composed of cross-linked points only has the modulus
many orders of magnitude higher than the modulus of the linear elastomer, then
<< 1, and for high-crosslinked networks << . That is why the second term in
expression (X.17) may be neglected. Therefrom, we obtain that
El.f.
n
=
.
E net n +

(X.18)

d
Let us turn back to efficient values leff, qeff, leff
and Seff. Taking into account
that the elasticity modulus of the rotational-isomeric subsystem is proportional to the
mean-statistic number of chains n concluded in linear fragments between cross-linked
points, correlation (X.10) may be reduced to the form:

El.f. =

leff RT
,
d
leff qeff S eff n

( )

(X.19)

and for the case of n0 = 1 and temperature T0, the rotational-isomeric modulus of such
a hypothetical polymer equals:

(El.f. )0 =

leff RT0

(leffd )2 qeff0 Seff

(X.20)

Taking the ratio of these moduli, we obtain


0
El.f.
T qeff
=
,
(El.f. )0 T0 nqeff

(X.21)

E
E
0
= exp eff .
where qeff = exp eff and qeff
RT
RT0
Substituting this expression into correlation (X.18), we obtain:
0
E net
n + T qeff
=
.
(El.f. )0 n 2 T0 qeff

(X.22)

462

For practical calculations, it should be taken into account that the repeating
fragment of the network consists of n/2 of repeat units. That is why expression
(X.22) with regard to functionality must be reduced to the form:
E net
2(n + ) T
=
(El.f. )0
n 2 T0

0
qeff

qeff

(X.23)

Let us analyze in more detail equation (X.23) comparing it with equation (X.1)
of the classical rubber elasticity theory. Clearly, these equations are analogous,
equation (X.23) describing the modulus of both high-crosslinked and sparse networks.
Actually, writing down n = Mmix/M0, from equation (X.23) we obtain:
0
2 (M c / M 0 + ) T qeff
,
E net = (El.f. )0
(M c / M 0 )2 T0 qeff

(X.24)

where M0 is the molecular mass of the repeat unit.


For sparse networks, << Mc/M0. Then
0
2 M 0 T qeff
E net = (El.f. )0
.
M c T0 qeff

(X.25)

Comparing expressions (X.1) and (X.25), we obtain that

(El.f. )0 = 3RT0 q0eff

(X.26)

2M 0 qeff

0
, and
If measurements of the modulus are performed at T = T0, then qeff = qeff
equation (X.23) is reduced to the form:

Emix = (El.f. )0

2( n + )
n 2

(X.27)

In the presence of large chain fragments between cross-linked points (sparse


networks) the condition << n is fulfilled, and equation (X.27) is reduced to the
correlation:
E net = (El.f. )0

2
n

(X.28)

0
qeff and M0n = Mc, we obtain:
and, with regard to correlations T0 T = qeff

E net =

3RT
,
Mc

(X.29)

463

i.e. correlation (X.29) is entirely in agreement with equation (X.1). Correlation (X,27)
is applicable to practical calculations of the modulus of high-crosslinked networks,
the value (El.f.)0 being calculated either by equation (X.26) or determined from
experimental data on the equilibrium modulus for sparse networks, when n >> 1.
0
Let us perform a series of transformations. When T = T0 and qeff = qeff
,
expression (X.29) obtains the form

(El.f. )0 = 3RT
2M 0

(X.30)

Then from correlations (X.30) and (X.27) we obtain:


E net =

3RT (n + )

.
(X.31)
M 0n2
Taking into account that Mc = M0n, we obtain from expression (X.31) that
E net =

3RT (n + )
.
M cn

(X.32)

Dividing the numerator and the denominator of expression (X.32) by n, we


reduce it to the form:
E net =

3RT ( n + n)
.
Mc

(X.33)

Mc =

3RT ( n + n)
.
E net

(X.34)

or

For a sparse network, << 1 and n >> 1, that is why /n 0, and generalized
equation (X.33) is transformed into usual equation (X.1).
To estimate the value Mc for the case of high-crosslinked networks, expression
(X.34) may be used. Preliminarily, for convenience, it may be transformed with
regard to n = Mc/M0 to the following form:
M
3 RT 1 + 0
Mc
E net =
.
Mc

(X.35)

Then from equation (X.35) we get:


Mc =

3 RT + 3RT (3RT + 4 E net M 0 )


2 E net

(X.36)

For the sparse networks, 0 and expression (X.36) transforms to equation


(X.1).

464

As an example, let us discuss model networks based on polydimethylsiloxane


used in ref. [188]. The network is of the following structure:
...
O
(H3C
CH3
...

(O

Si)m +1
CH3

(H3C

Si CH3)m +1
O
CH3
Si

(Si

CH3

Si

CH3)m +1

...

O)m +1

O
..
.
The cross-linked point of the network is marked by dotted lines. For this
network,

Vi
= VSi,175 + 4VO,135 = 34.7 3;

i
crl.p.

Vi = m(VSi,172 + 2VC,106 + 4VO,135 + 6VH,124) = 72.1 3;

i
l.f.
= 34.7/72.1 = 0.481.
Calculation of the equilibrium rubber modulus of high cross-linked networks
by equation (X.27) and the glass transition temperature Tg by equation (IV.85)
requires taking into account all details of chemical structure of the network. The point
is that when m = 0, the structure of this network obtains the form
...
O
H3C
CH3
...

Si

CH3

O
O

CH3
H3C

...

Si CH3
Si

O
Si

Si
CH3

CH3

O
O

Si

...

O
...

O
...

CH3
In this case, only a part of the unit

Si
CH3

remains as a linear fragment

465

between neighboring cross-linked points, because the oxygen atom in this unit is a
component of the neighboring cross-linked point. Determining n as
*

Vi

i
l.f.

Vi , where

i
l.f.

Vi is the Van-der-Waals volume of

i
l.f.

CH3
Si

group, we get n = 0.99. If m = 1, n = 1.99, etc. Of prime importance is

CH3
taking into account of this fact in calculation of the glass transition temperature Tg,
which is calculated by equation (IV.85). In this case, equation (IV.85) transformed to
the formula
0

Tg =

Vi
+ m Vi + Vi

i
crl.p.
i
l.f. i
l.f..
0

K i Vi
+ m ai Vi + ai Vi

i
crl.p.
i
l.f. i
l.f.
0

(X.37)

where ai Vi and ai Vi are series of increments for structures

i
l.f.
i
l.f.
CH3
Si

CH3
O

and

CH3
considered,

Si

, respectively (note that in the case of the network

CH3

b j

= 0).

Calculation of the equilibrium rubbery modulus Enet by equation (X.27)


requires the value (El.f.)0 to be estimated first. As mentioned above, this may be
performed by two methods. The first method concludes in application of expression
(X.30), i.e. the value (El.f.)0 is determined empirically. Substituting = 1.169 g/cm3,
= 4, M0 = 74.15, T = 293 K into this equation, we obtain that (El.f.)0 = 230 MPa.
The second method concludes in application of expression (X.27) and
experimental value of Enet for a sparse network. If the experimental value of Enet
determined in ref. [188] is used then, basing on expression (X.27), we obtain that
(El.f.)0 = 197 MPa. Calculations of Enet by expression (X.27) using this value of (El.f.)0
give equilibrium rubbery moduli shown in Table 36. Clearly, Enet increases sharply
with shortening of the distance between neighboring cross-linked points of the
network, and for high-crosslinked network (n = 1) reaches high values. Therewith, the
glass transition temperature remains below room temperature.

466

Table 36
M
1
1
2
3
4
6
9

n
0.99
1.99
2.99
3.99
4.99
6.99
9.99

Values of , Enet and Tg for polydimethylsiloxane networks


Emix, MPa

0.482
148.0
0.241
61.5
0.160
38.2
0.120
27.6
0.096
21.6
0.069
15.0
0.048
10.3

Tg, K
183
166
161
158
156
154
153

X.2 Heteromodular and gradient-modulus polymers


Concluding this Chapter, let us consider application of calculation schemes
which allow estimation of the glass transition temperature Tg and the equilibrium
rubbery modulus E to create polymeric materials with unusual properties. The case
in point is the production of elastic polymeric materials of two types: 1)
heteromodular ones, displaying the elasticity modulus constant for every sample,
which changes in a very wide range when transiting from sample to sample; 2)
gradient-modulus ones displaying the elasticity modulus changing smoothly in the
frames of the same sample along the assigned direction, therewith, the material
possesses no layers or interfaces. In the latter case, it is necessary to obtain a smooth
transition from rubber to plastic or vice versa in the volume of the same material that
enables heteromodular constructions to be created without using traditional methods
of bonding gluing, welding, etc.
Complexity of realization of this idea is associated with two main features of
polymeric material behavior, which must be clarified.
Figure 76 schematically shows the temperature dependence of the elasticity
modulus for a typical polymer: unfortunately, in the glassy state range, the elasticity
modulus weakly depends on the chemical structure of the polymer and varies in the
range (23)103 MPa. This the first difficulty of obtaining heteromodular and
gradient-modulus materials.

Figure 76. Schematic representation of dependence of elasticity modulus E on temperature T.

In the area of transition from the glassy state into the rubbery state, the
elasticity modulus decreases abruptly in a narrow range of temperature (T = 2030)

467

by several orders of magnitude (see Figure 76). In this connection, at first glance, it
may seem that the problem of obtaining polymers with different elasticity moduli
lying, for example, in the range from 3103 MPa to 3 MPa, is quite simple: to
accomplish this, polymers possessing the glass transition temperature Tg close to room
temperature must be produced (if heteromodular materials should work at room
temperature). However, it is common knowledge that materials in the transition zone
display a clearly expressed viscoelastic behavior and, moreover, their mechanical
properties change sharply at extremely small changes, both decrease (transition to
plastic) and increase (transition to rubber), of temperature. This is the second
difficulty of obtaining heteromodular materials, which besides the wide range of
change of the elasticity modulus must possess elastic but not viscoelastic properties.
Moreover, they must retain the assigned gradient of the properties in a wide
temperature range.
Theoretically, the wide range of the elasticity modulus without applying any
plasticizers or fillers may be obtained by creating high-crosslinked network structures
containing bulky cross-linked points bonded by flexible linear chains of controlled
length (Figure 77).

Figure 77. Schematic representation of network consisting of bulky cross-linked points and short
flexible chains as linear fragments.

This yields from generalized equation (X.33) for estimation of the equilibrium
rubbery modulus E for network systems.
According to equation (X.33), a high value of E (or Enet) may be reached
transiting to high-crosslinked networks with bulky cross-linked points when n = 1 and
> 1. Therewith, to keep glass transition temperature Tg low, linear fragments
connecting cross-linked points must be extremely flexible. Polyisocyanurates
networks, the chemical structure of which is displayed in Chapter VIII, were
synthesized [45, 46] as the structures containing rigid bulky cross-linked points
bonded by linear flexible chains (R).
The role of a cross-linked point (marked by dotted lines) is played by
isocyanurate cycle possessing the functionality equal to three.
Short organosilicon chains, the structure of which is indicated in Chapter VIII,
were used as linear fragments.
In another method, polyisocyanurate copolymer networks were produced
[127].
The general principle of synthesis of these materials is that application of
olygomeric diols as one of the original components interacting with a diisocyanate,
2,4-toluene diisocyanate, in particular, by the urethane-forming reaction, gives first
macrodiisocyanates:

468

H3C

CH3

O C N

N C O
O
NH C O

O
R O C

HN

Industrial oligomeric rubbers with hydroxyl end groups may be used as diols.
One of them is PF-OP-15 polyether representing the oligomeric copolymer of
tetrahydrofurane and propylene oxide:
R:{

[O

(CH2)4]m

(O CH2

CH)n }p
CH3

m/n = 19.7/1.7; p = 1.2.

Figure 78. Dependence of calculated glass transition temperature Tg of polyisocyanurate polymers


based on diisocyanates with different structure R as dictated by the mole fraction of
rubber PF-OP-15:

1) R:

3) R:

; 2) R:

CH2

CH2

; 4) R:

(CH2)6

Macrodiisocyanates obtained from oligomeric rubbers form polymer networks


by the reaction of polycyclotrimerization in situ and copolymerize with diisocyanates
of any chemical structure (aromatic, alicyclic or aliphatic), capable of forming
polyisocyanurate networks at an acceptable rate. In reality, mixed copolymeric
structure is formed in this case. The polymer may contain a network composed of
products of homopolycyclotrimerization of oligomer and diisocyanate simultaneously

469

with a network formed by interaction of the oligomer with diisocyanate producing


structures with the arbitrary disposition of linking chains.
As the excess of diisocyanate in the initial reactionary mixture increases
compared with oligomeric macrodiisocyanate, the concentration of rigid cyclic
structures in the network formed increases. Hence, the ratio of flexible (polyether) and
rigid (isocyanurate cycles with joint aromatic cycles) fragments may be changed as
desired that regulates mechanical properties in a wide range. To check the real
possibility of this regulation of polymer networks properties, calculation of Tg value
was performed preliminarily with application of equation (IV.85). Figure 78 indicates
a smooth decrease of Tg value as the part of flexible rubber fragment increases.
Calculations of the equilibrium elasticity modulus performed by equation
(X.27) indicated [127] that at low Tg (below room temperature) the equilibrium
modulus E may obtain high values, intermediate between values of the moduli for
rubber and plastics. Such preliminary calculations were performed for networks with
organosilicon linear fragments [45]. Consequently, an expression was deduced for the
structure considered, which connects Tg of the network with the amount of
dimethylsiloxane units n in the linear fragment:
Tg =

751 + 108.15n
.
2,450 + 721.5n

(X.38)

Calculations performed according to expression (X.38) are shown in Table 37,


which indicates that the glass transition temperature lies below room temperature and
decreases as the value of n increases approaching the glass transition temperature of
polydimethylsiloxane at n = 44.
Table 37 also indicates the values of equilibrium moduli, which obtain
different values in the range from 3 to 870 MPa in dependence on the value of n.
Synthesis of polymer networks with the above-discussed structure performed
confirmed correctness of the suppositions and calculations made [45, 46]. For
example, E of obtained networks with organosilicon linear fragments changes in the
range from 3 to 3103 MPa (Figure 79) in dependence on the length of the linear
fragment.
Table 37
Values of calculated glass transition temperatures Tg and equilibrium modulus of elasticity E of
cured macrodiisocyanate depending on the number of dimethylsiloxane repeat units
n
Eexper, MPa
Ecalc, MPa
Tg, C
0
33

1
2
884
886
2
5
249
283
3
41
123
149
6
67
40
43
9
80
22
24
19
99
8
7
44
112
3
3

Of special importance for these materials is the process of microphase


separation associated with the substantially different surface energy of organosilicon
chains (21 dyn/cm) and isocyanurate cross-linked points (35 dyn/cm). Microphase
separation has been confirmed by X-ray photoelectron spectroscopy, electron
microscopy and dynamic mechanical analysis.

470

Figure 79. Logarithmic dependence of the elasticity modulus E on a number of dimethylsiloxane units
n in linear fragment.

Generally speaking, the microphase separation may proceed both in linear


block-copolymers or mixtures of polymers and in networks.
In these cases, two clearly expressed maxima are observed on the temperature
dependence of mechanical loss factor tg.
Figure 80 represents the temperature dependence of mechanical loss factor tg
for bulky samples of considered polyisocyanurates possessing different lengths of
linear organosilicon fragments. The low-temperature peak is shifted to the side of low
temperatures with increasing length of the linear fragment, approaching Tg of
polydimethylsiloxane. The high-temperature maximum associated with devitrification
of the whole system does not practically depend on the value n starting from n = 2.

Figure 80. Temperature dependence of the mechanical loss-factor tg for bulky specimens of
polyisocyanurate networks with different length of linear organosilicon fragments n: 1
1.2; 2 2.1; 3 3.2; 4 5.4; 5 9.2; 6 22 .

Hence, two transition temperatures, one of which may be related to the


devitrification temperature of the polydimethylsiloxane microphase in the polymeric
network and the second to the devitrification temperature of a microphase

471

containing isocyanurate cross-linked points with attached branchings, are clearly


observed. Thereby, the existence of microphase separation in the system associated
with thermodynamic incompatibility of polydimethylsiloxane fragments with
isocyanurate cross-linked points and branchings attached to them should be admitted.
Taking into account that the transition temperatures of both microphases are
shifted towards each other, the composition of microphases may be calculated basing
on the condition that the transition temperature must coincide with the glass transition
temperature of microphases of the present composition. Because the transition
temperature in the microphase generally composed of organosilicon fragments is
higher than the glass transition temperature of polydimethylsiloxane, it is obvious that
neighboring fragments are contained in the polydimethylsiloxane microphase.
To answer the question about the structure of these fragments, the glass
transition temperatures of the structures shown below have been calculated in ref.
[45]:
CH3
NH C O CH2
O

CH2

O CH2

Si
CH3

CH3
O

Si
n CH3

CH2

O CH2

CH2

O C NH
O

The glass transition temperature was calculated from formula (IV.41), the
step-by-step calculation of Tg for polydimethylsiloxane fragment with different n and
attached parts of chains marked by dotted lines being performed. These parts were
lengthened until the calculated glass transition temperature coincided with the
experimental temperature of the first transition. Consequently, the coincidence of
calculated and experimental values of the transition temperature was observed in the
case when polydimethylsiloxane domains contained parts of chains marked by double
dotted lines. Therewith, the structure of these parts is independent of the value n, i.e.
of the length of the organosilicon chain. This method of estimating the composition of
microdomains may be extended to any object characterized by microphase separation.
If a microphase separation proceeds in the system, then in calculation of the
equilibrium rubbery modulus the notion of the network cross-linked point from the
viewpoint of its chemical structure must be broadened. Actually, the presence of a
rigid microphase (the above-considered one, for example, which includes
isocyanurate cross-linked points with attached branchings) enables the microphase to
play the role of a cross-linked point. Therewith, all one needs to do is to be convinced
that the glass transition temperature of this microphase is above room temperature.
The presence of this macrocross-linked point causes a significant increase of value
in equation (X.27) that promotes an increase of the equilibrium rubbery modulus.
Ref. [45] displays this analysis performed for the system discussed above.
Consequently, it was obtained that the glass transition temperature of rigid domains
was 33C, and their Van-der-Waals volume equaled 751 3.
Taking into account that the Van-der-Waals volume of the repeat unit of
polydimethylsiloxane equals 180.15 3, the value = 751/180.15 4.2.
Experimentally, the equilibrium rubbery moduli were determined [46] with the
help of measurements of stress relaxation curves approximated with the help of a
physically proved relaxation memory function (see below).
The mechanical behavior of polyisocyanurate networks containing
organosilicon linear fragments was demonstrated above (see Figure 71). Figure 79
indicates the dependence of the elasticity modulus of polyisocyanurate networks on
the number of repeat units of polydimethylsiloxane chains linking cross-linked points.

472

The elasticity modulus of such networks overlaps the range of moduli typical of the
transition zone from the glassy state to the rubbery state. However, in spite of this, the
mechanical behavior of present materials is not viscoelastic as for all polymers in the
transition zone, but elastic typical of polymeric glasses. This is proved by the form of
stress relaxation curves displayed in Figure 71.
Let us call attention once again to the mechanical behavior of polyisocyanurate network with linear organosilicon fragments with n = 6.2. At these sizes of
linear chains, the initial stress 0 is approximately coincident with 0 for a
viscoelastic material (7.5 MPa), but mechanical behavior is significantly different
from the former one: stress relaxes fast by a low value at the initial moment of time,
and then stress decrease stops, i.e. the material behaves itself as an elastic glassy
polymer1.
Let us now turn to analysis of the properties of gradient-modulus materials
produced on the basis of oligomeric rubber PF-OP-15 and 2,4-toluylene diisocyanate.
For producing gradient-modulus materials, initial components are dosed
smoothly to ensure directed control of the chemical composition of the network in the
same sample.

Figure 81. Dependence of the elasticity modulus lgE or E on concentration of 2,4-toluylene


diisocyanate G in its mixture with oligomeric macrodiisocyanate; the concentration of G
varies along the specimen length l.

In this manner, the samples were produced in refs. [19, 129] in which the
elasticity modulus changed from 4.5 MPa (typical of rubbers) to 2000 MPa (typical of
1
It should be taken into account that there are no absolute elastic polymeric materials (excluding ideal
crystals). That is why the terms elastic behavior and elastic material in relation to polymers are
conditional: by elastic behavior we mean extremely slow stress relaxation.

473

plastics). Therewith, this change proceeded smoothly in the same material with no
interfaces and intermediate layers.
Figure 81 displays an example of the dependence of the elasticity modulus on
the concentration of 2,4-toluylene diisocyanate in the initial mixture. It is clearly
observed that the elasticity modulus changes linearly along the sample and, hence, as
mentioned above, a smooth transition from rubber to plastic is performed with no
interfaces.
To analyze the mechanical behavior of materials obtained, stress relaxation
curves were measured for microspecimens cut off from the initial macrospecimen in
different points of the gradient (Figure 82). This Figure indicates for comparison the
stress relaxation curve for a viscoelastic material epoxy resin ED-20 cured by
polysebacic acid anhydride in the presence of azelaic acid, for which Tg is located
near room temperature, i.e. for the same material, the relaxation behavior of which is
shown in Figure 71. Stress relaxation curves were plotted in relative stress time
coordinates. Relative stress was calculated as /0, where is the current relaxing
stress, 0 is the initial stress developed at the moment of the end of immediate
setting of deformation.

Figure 82. Curves of relative stress relaxation /0 for polyisocyanurate networks prepared from
oligomeric macrodiisocyanate and 2,4-toluylene diisocyanate when the concentration of
the latter in percent by weight is: 1 6, 2 46, 3 27, respectively; 4 epoxy oligomer
ED-20 cured with methyltetrahydrophthalic anhydride; 5 epoxy viscoelastic polymer (the
composition is indicated in the text).

From this Figure, it is obvious that as for usual viscoelastic material the
relative stress relaxes fast down to zero, for polymer networks obtained in refs. [19,
129] a slower decrease of stress is observed, typical of polymeric glasses or rubbers
with further transition to extremely low stress relaxation.
Hence, in spite of the fact that a definite part of the material in the same
specimen possesses values of the modulus typical of the transition zone, mechanical
behavior is elastic as for glasses of rubbers but not viscoelastic typical of all polymers
in the transition zone.
Dynamic mechanical analysis indicates that tg for the networks obtained is
extremely low which is typical of elastic materials, despite the values of the storage
modulus E typical of the transition zone.
Basing on polyisocyanurate networks, films with a gradient of the elasticity
modulus perpendicular to the film surface (in direction of thin) were obtained. These

474

films formed from high-crosslinked polymer networks, possess good mechanical


properties (strength of 50 MPa and ultimate elongation of 90%). Hence, obtaining of
gradient-modulus materials allows us to avoid the main drawback of high-crosslinked
polymer networks their brittleness.

Chapter XI. Description of relaxation processes in polymers

XI.1 Stress relaxation


Various variants of the memory function in the appropriate Boltzmann
Volterra equations have been suggested to date to describe processes of stress
relaxation and creep. Combined description of these memory functions and their
resolvent is present in monograph [112]. The memory function contains three or four
parameters, moreover they usually possess a fractional exponent of time, because only
in this case experimental data may be described with a good approximation by stress
relaxation and creep.
Analysis of proposed memory functions has indicated that at proper selection
of parameters they give the course of relaxation processes with acceptable accuracy.
However, the physical meaning of these parameters is not always clear, although
attempts to disclose the physical meaning of some memory functions have been made
[74, 104]. Reference [7] suggested an approach to production of relaxation memory
functions based on consideration of thermodynamic functions and their changes
during relaxation.
Suppose that the stress relaxation proceeds as a result of interaction and
diffusion of kinetic units relaxants. Relaxants may be various groups of atoms,
repeat units, larger fragments and the whole fragments of macromolecules. Among
relaxants are also separate elements of the free (empty) volume, i.e. microcavities,
stress concentrators, etc. Interacting with each other, these microcavities may unite,
be rearranged and diffuse in a polymeric material during relaxation, forming a
structure which promotes a decrease of relaxing stress. The polymeric material may
then be considered as the one consisting of relaxants and non-relaxants, the most part
of the sample material after instantaneous setting of deformation consisting of
relaxants interacting with each other forming a non-relaxing material. Appearance of
particles of two types (relaxants and non-relaxants) and their diffusion cause
production of entropy in the system, which increases in the course of stress relaxation.
The production of entropy (or the rate of appearance of entropy) is determined
dS 1
by the expression
, where S is entropy, t is time, V is the volume of the system.
dt V
Relaxation memory functions were obtained in ref. [7], assuming that the driving
force of the process is production of the system (sample) entropy, which increases up
to the maximum value during stress relaxation.
It is common knowledge that if the system consists of two types of particles,
the entropy of mixing S in this system determined with the help of the Boltzmann
equation is calculated from the expression
m*!
,
S = k B ln
m1*! m2*!

(XI.1)

where m* is the total number of particles (in this case, the number of relaxants and
non-relaxants per specific volume); m1* and m*2 are numbers of relaxants and non-

476

relaxants in the specific volume, respectively; kB is the Boltzmann constant. Taking


into account that two types of particles are present, it may be written down that
S = kB ln

m* !

( m ) !(1 ) m !
*

(XI.2)

where is the part of relaxants in the total number of particles. Using an approximate
formula for calculating the factorial at large m*, basing on equation (XI.2), we obtain:

S = kB ln

( )

2 m* m*

2 m* m*

m*

m*

e m

2 (1 )m* (1 ) m*

(1 ) m* (1 ) m*

.,

(XI.3)

After some transformations and neglecting low components of the expression,


we obtain the following form:
S = kBm*[ln + (1 )ln(1 )].

(XI.4)

The value of changes with time t from 1 to 0.5, because at = 0.5 the
entropy of mixing reaches its maximum.
Assume that the memory function in the BoltzmannVolterra equation is
associated with the reverse dependence of the type
1
1
T ( ) = S 0
S S max

*
T ( ) d ,
0

(XI.5)

where T*() is the variable part of the memory function. Then, substituting expression
(XI.4) into equation (XI.5), we obtain
T ( ) =

S0
1
1

k B m ln + (1 ) ln(1 ) ln 0.5

(XI.6)

where m = m* T * ( )d .
0

Now it is necessary to indicate the dependence of on (remember that is


the part of relaxants in the total number of particles in the system). As mentioned
above, the change of with time may be stipulated by two reasons: interaction of
relaxants and their transition into non-relaxants, and diffusion of kinetic units (note
that the mechanism associated with diffusion of microdefects is described in refs.
[104, 119]). Let us discuss these reasons.
Since the interaction of relaxants is complicated, it is reasonable to describe it
by the reaction equation of the n-th order. If in a usual chemical reaction, for example,
the third order is rarely observed (because this requires active collision of three
particles simultaneously), in this case, relaxants are condensed in the sample, and
the elementary act of their interaction may include several relaxants together (for

477

example, fusion of several microcavities into a single one). Therewith, the reaction
order may be fractional. For this case, the following kinetic equation is true:
dc
= kc n ,
d

(XI.7)

where k is the reaction constant; c is the concentration. Integrating (XI.7) from = 0


to t, we obtain that
c=

c0

1 + c0n 1 (n 1) k n1

where c0 is the initial concentration of relaxants of any type (assume for ease that
these concentrations are equal for different types of relaxants). Then

c
=
c0

(1 + k * / )

(XI.8)

1
where k * = kc0n 1 ; =
; n is the reaction order.
n 1
To obtain the relaxation memory function, it is necessary to substitute
expression (XI.8) into equation (XI.6). Therewith, it should be taken into account that
at the end of deformation the fraction of relaxants is not 1, but somewhat smaller,
equal to 1 0. Taking into account all the above-mentioned, we obtain

S0
1

T1 ( ) =

k B m1

1
1

0 ln
0 +

*
*

1+ k /
1+ k /

1
1


ln 0.5
1
1

+ 0 ln1
+0

*
*

1+ k /
1+ k /

(XI.9)

The function T1() is of the physical sense only under the condition that
1
0.5 . Hence, the memory function T1() contains four parameters:

*
1+ k /
S
1
and 0.
A = 0 ; k * = kc0n 1 ; =
k B m1
n 1

478

According to requirements on the relaxation memory function, the value 0


must be extremely low; in ref. [7] 0 was assumed to be equal to 1010 basing on
approximation of stress relaxation curves for a series of polymers.
At low values of k*/, i.e. in initial stages of relaxation, the function T1() is

substantially simplified. Actually, at low k*/, the value 1 + k * /


Then

S0
1

T1 ( ) =

k B m1 1
1

0 ln
0 +

*
1 + k *

+
k

(XI.10)

1
1


ln 0.5
1
1
1
+ 0 ln1
+0

1 + k *

1 + k *

= 1 + k*.

After further transformations, we obtain


T1 ( ) =

S0
1
1

*
.
kB m1 k + 0 ln k * + 0 1 ln 0.5

) (

(XI.11)

Neglecting the second summand in the braces due to its smallness, it is


obtained that
T1 ( ) =

S0
1
.
*
kB m1 k + 0 ln k * + 0 1

) (

(XI.12)

Let us integrate expression (XI.12) after some preliminary transformations:


t

T1 ( )d =

k * +
0 .

d
* *
*

e
k B m1k 0 k + 0 k + 0

ln
e
e
S0

(XI.13)

Then, we obtain
t

T1 ( )d =

S0
k B m1k

ln

ln k * + 0
e

)t =
0

S0
k B m1k

ln

ln k *t + 0 1
.
ln 0 1

(XI.14)

479

Let us substitute the value of T1 ( )d into the Boltzmann equation:


0

= 0 1 T1 ( )d ,
0

(XI.15)

where 0 is the initial stress appeared as a result of instantaneous deformation.


Consequently, we obtain the relation for description of stress relaxation in the
initial part (at low values of t):

=0 +

0 S0
k B m1k

ln

ln k *t + 0 1
.
ln 0 1

If expression (XI.16) is true, the dependence of on ln

0 S0

(XI.16)

ln k *t + 0 1
must
ln 0 1

, and the section cut-off on


k B m1k *
the axis of ordinates 0. Figure 83 displays such dependence for poly(methyl
methacrylate). It is clear that experimental points fit the straight line with high
accuracy.
represent a straight line, tangent of which equals

ln k *t + 0 1
(for explanation see text). Stress relaxation for
ln 0 1
PMMA-material under conditions of uniaxial compression when the temperature is 295 K
and the deformation is 2.2%.

Figure 83. Dependence of on ln

Let us now consider the diffusion mechanism of relaxation. When the


continuous-time random walks of kinetic units take place, the amount of sites
occupied by them at time and, consequently, the part of non-relaxants 1 is
determined from the correlation [146]:
(1 ) = ab/2,

(XI.17)

480

where 0 < b < 1; a is a constant. In the case, if b = 1, correlation (XI.17) corresponds


to the Fick diffusion:
1/ 2

4 D
1 =

,
1/ 2

4 D

l
Substituting correlation (XI.17) into equation (XI.6), we obtain

where l is the sample size; D is the diffusion coefficient; i.e. a =

T2 ( ) =

S0
k B m2

1
1

a ln a + (1 a ) ln(1 a ) ln 0.5

(XI.18)

where = b/2. The function T2() is of the physical meaning only under the condition,
if a 0.5.
S
The memory function (XI.18) contains three parameters: A = 0 , a and
k B m2
. It represents a function with a non-essential singularity at = 0.
At low values of a the function T2() may be simplified. In this case,
correlation (XI.18) will be reduced to the form:
T2 ( ) =

S0
1
1


.
k B m2 a (ln a 1) ln 0.5

(XI.19)

Neglecting the second summand in the square brackets, we obtain

S0
1
.
T2 ( ) =
(XI.20)
k B m2 ln a
a

After a series of preliminary transformations, let us integrate correlation


(XI.20):
t

T2 ( )d =

S0
k B m2 a

a
d
a ln a e

0
e
e
t

Let us make variable substitution:


1/

a
ex
ex
= x ; = ; =
e
a
a
Then

481

1 / 1

xe

a
x ln x

x2

S0

T2 ( )d = k Bm2 a

x1

dx.

(XI.21)

dx.

(XI.22)

In what follows we obtain that


t

T2 ( )d =

S 0 e1 / 1
k B m2 a

Taking into account the x1 = 0 and x2 =


t

T2 ( )d =

1/

x2 1 / 2

x
1

ln x

at
, we get
e

a 1 / 1
li
t1
1/
e

k B m2 a
S 0 e1 / 1

(XI.23)

where li is the integral logarithm.


t

Substituting the value of

T2 ( )d

into the Boltzmann equation (XI.15), we

deduce the correlation describing stress relaxation in the initial section of the
relaxation curve under the condition that at << 1:
1 / 1

S e1 / 1 a
li
t1 ,
=0 + 0 0

k B m2 a1 / e

(XI.24)

a 1 / 1

t1 , we obtain the
If correlation (XI.24) is fulfilled in li
e

straight line, the tangent of which equals

0 S 0 e1 / 1

, and the section cut-off from


k B m2 a1 /
the axis of ordinates equal to 0. If the limiting stage of relaxation is interaction of
relaxants and transition of them into a non-relaxing material, memory function (XI.9)
or reduced expression (XI.16) on the initial part must act. If the limiting stage of the
process is diffusion of non-relaxants formed, memory function (XI.18) or reduced
expression (XI.24) in the initial part must act.
t

As an example, Figure 84 shows dependences of on

T1 ( )d

and on

0
t

T2 ( )d . Clearly, in this case, the memory function T1()represents better the stress
0

relaxation for poly(methyl methacrylate).

482

Figure 84. Dependences of on T1* ( )d (1) and T2* ( )d (2). (Experiment conditions are similar
to those shown in Figure 83).

Let us now turn back to the relaxation memory function, based on the analysis
of kinetics of relaxants interaction and their transition into a non-relaxing material.
Memory function (XI.9) has been deduced from the condition that the interaction of
relaxants represents a unilateral irreversible reaction of the n-th order, i.e. it is
assumed that the relaxants interacted transit into a non-relaxing material in a way that
makes the process irreversible.
Ref. [44] discusses the interaction of relaxants by a reversible reaction of the
n-th order, i.e. it is assumed that a non-relaxing material may produce relaxants during
the process. The kinetic equation of this reaction is of the following form:

d
= k n k (1 ) n .
dt

(XI.25)

Equation (XI.25) is deduced under the condition that rate constants of direct
and reverse reactions are identical and equal to k. This leads to the situation when the
system transits to the equilibrium state, parts of relaxants and non-relaxants become
the same and equal to 0.5. Equation (XI.25) is integrated up to the end only in some
particular cases, for example, at n = 2. In the general case, when n is a fractional
value, only numerical integration may be performed. To find the dependence of the
transformation degree on time t, the RungeKutta numerical method with automatic
selection of the integration step was used in ref. [44]. The resultant values of values ,
calculated at various small steps with respect to t, were used for computer calculations
t

of the values of the integral of the variable part of the memory function T3* ( )d ,
0

where
T3* ( ) =

1
1
+
.
( 0 ) ln( 0 ) + (1 + 0 ) ln(1 + 0 ) ln 0.5

(XI.26)

483

Hence, we get three relaxation memory functions T1(), T2() and T3(),
therewith
S
T1 ( ) = 0 T1* ( ) ;
(XI.27)
k B m1
S
T2 ( ) = 0 T2* ( ) ;
(XI.28)
k B m2
S
T3 ( ) = 0 T3* ( ) ,
(XI.29)
k B m3
where T1* ( ) , T2* ( ) and T3* ( ) are variable parts of memory function T1(),T2() and
T3(), respectively.
These variable parts of the relaxation memory function are described by
correlations:
T1* ( ) =

where =

1
1

,
( 0 ) ln( 0 ) + (1 + 0 ) ln(1 + 0 ) ln 0.5
1

(1 + k / )
*

T2* ( ) =

(XI.30)

a ln a

+ (1 a ) ln(1 a )

1
.
ln 0.5

(XI.31)

The variable part of the memory function T3() is described by correlation


(XI.26). Substituting memory function T1(), T2() and T3() into the Boltzmann
equation, we obtain:
t

S
= 0 0 0 T1* ( ) d ;
k B m1

(XI.32)

S
= 0 0 0 T2* ( )d ;
k B m2

(XI.33)

S
= 0 0 0 T3* ( )d .
k B m3

(XI.34)

0
t

0
t

In equations (XI.32)(XI.34)

m1 = m1* T1* ( )d ;

(XI.35)

m2 = m2* T2* ( ) d ;

(XI.36)

484

m3 = m3* T3* ( )d ,

(XI.37)

where m1* and m3* are the quantities of inhomogeneities (relaxants) in the polymeric
material which interact with each other in the course of relaxation; m*2 is the number
of diffusing inhomogeneities in the material during stress relaxation.
If equations (XI.32)(XI.34) describe correctly the shape of experimental
t

stress relaxation curves, then in coordinates T1* ( )d , T2* ( ) d and


t

T3 ( )d
*

a straight line must be formed, tangents of which are equal to 0S0/kBmi,

and the section cut-off from the axis of ordinates is 0.


To use equations (XI.32), (XI.33) and (XI.34), it is necessary to know values
t

of integrals T1* ( )d , T2* ( ) d and T3* ( ) d . These values depend on two pairs
of parameters: k* and , a and , respectively. These integrals have been determined
by computerized numerical methods [13, 44] at varied parameters k* and , a and .
Tables 3840 show numerical values of variable parts T1* ( ) , T2* ( ) and
T3* ( ) of relaxation memory function T1(), T2() and T3(), taken from refs. [13, 44].
Basing on these values, the stress relaxation curves may be approximated in a wide
range of process time t.
According to equations (XI.32), (XI.33) and (XI.34), to process experimental
data by the root squares fitting, it is advisable to perform the following procedure.
Initially, all values of integrals

*
*
*
T1 ( )d , T2 ( )d and T3 ( )d in the form of

three dimensions are stored in the computer memory. Each dimension contains values
t

of integrals T1* ( )d , T2* ( ) d and T3* ( ) d at different chosen times t for each
pair of system parameters k* and , a and (dimension 1 corresponds to Table 38,
dimension 2 to Table 39, dimension 3 to Table 40). Every experimental
dependence (t) is approximated by equation (XI.32), (XI.33) or (XI.34), and values
of pairs of parameters k* and , a and are automatically selected, for which the sum
of deviation squares of the experimental values from the experimental ones is
minimal, and the correlation coefficient is maximal.
Experimental checking of memory function (XI.9), (XI.18) and (XI.26) and of
the whole procedure was performed in a series of works [11, 12, 14, 38] on the
example of poly(methyl methacrylate), polyoxadiazole, polyimide, polybenzoxazole,
and other polymers.

485

Table 38
t

Values of T1* ( ) d
0

t
1

= 0.2
2

= 0.3
3

= 0.6
6

= 0.7
7

= 0.8
8

17.46
18.80
20.05
20.69
21.06
21.30
21.44
21.62
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66
21.66

17.45
18.78
20.01
20.63
20.98
21.19
21.31
21.45
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47
21.47

17.45
18.76
19.98
20.57
20.91
21.10
21.22
21.33
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34
21.34

133.4
144.5
156.6
164.2
169.7
174.1
177.8
186.0
196.5
210.3
217.5
222.0
227.2
229.7
231.6
231.9
232.0
232.0
232.0
232.0
232.0
232.0
232.0

= 0.4
= 0.5
4
5
k = 0.1
17.49
17.47
18.86
18.82
20.20
20.11
20.92
20.78
21.37
21.19
21.67
21.45
21.88
21.62
22.22
21.86
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
22.39
21.94
k = 0.01
133.4
133.4
144.5
144.4
156.5
156.4
164.0
163.9
169.5
169.3
173.8
173.6
177.4
177.1
185.4
185.0
195.4
194.7
208.1
206.7
214.2
212.1
217.6
215.0
221.2
217.6
222.4
218.2
222.8
218.2
222.8
218.2
222.8
218.2
222.8
218.2
222.8
218.2
222.8
218.2
222.8
218.2
222.8
218.2
222.8
218.2

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000

17.57
19.06
20.62
21.56
22.22
22.73
23.13
23.98
24.88
25.70
25.87
25.88
25.90
25.90
25.90
25.90
25.90
25.90
25.90
25.90
25.90
25.90
25.90
25.90

17.52
18.93
20.34
21.14
21.66
22.03
22.31
22.82
23.21
23.31
23.31
23.31
23.31
23.31
23.31
23.31
23.31
23.31
23.31
23.31
23.31
32.31
23.31
23.31

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080

133.5
144.6
156.8
164.5
170.2
174.8
178.6
187.3
198.6
214.7
223.1
230.3
238.9
244.3
250.7
254.0
255.9
257.1
258.0
258.1
258.1
258.1
258.1

133.4
144.4
156.4
163.8
169.3
173.5
177.0
184.8
194.3
205.8
210.7
213.2
215.2
215.4
215.4
215.4
215.4
215.4
215.4
215.4
215.4
215.4
215.4

133.4
144.4
156.3
163.8
169.2
173.4
176.9
184.6
193.9
205.1
209.7
211.9
213.5
213.5
213.5
213.5
213.5
213.5
213.5
213.5
213.5
213.5
213.5

133.4
144.4
156.3
163.7
169.1
173.3
176.8
184.4
193.7
204.6
208.9
210.9
212.2
212.2
212.2
212.2
212.2
212.2
212.2
212.2
212.2
212.2
212.2

100,000

258.1

232.0

222.8

215.4

213.5

212.2

218.2

486

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000

1,027
1,110
1,201
1,258
1,300
1,334
1,362
1,427
1,514
1,644
1,724
1,784
1,872
1,936
2,027
2,093
2,145
2,188
2,341
2,472
2,556
2,578
2,579
2,579

1,027
1,110
1,200
1,257
1,299
1,333
1,361
1,426
1,513
1,640
1,719
1,776
1,860
1,919
2,001
2,059
2,101
2,136
2,244
2,308
2,319
2,319
2,319
2,319

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000

7,900
8,556
9,256
9,692
10,010
10,270
10480
10970
11,620
12,570
13,170
13,610
14,270
14,760
15,470
16,000
16,430
16,800
18,230
19,760
21,360
22,930
24,090
24,090

7,900
8,556
9,256
9,692
10,010
10,270
10,480
10,970
11,620
12,570
13,160
13,600
14,260
14,750
15,450
15,980
16,390
16,760
18,130
19,560
20,940
22,130
22,820
22,820

0.5
1
2
3
4

59,840
65,230
70,910
74,420
76,970

59,840
65,230
70,910
74,420
76,970

5
k = 0.001
1,027
1,027
1,110
1,110
1,200
1,200
1,257
1,257
1,299
1,299
1,333
1,333
1,361
1,361
1,426
1,425
1,512
1,511
1,639
1,637
1,716
1,714
1,772
1,769
1,853
1,849
1,910
1,905
1,988
1,980
2,041
2,029
2,079
2,065
2,109
2,092
2,194
2,163
2,227
2,182
2,227
2,182
2,227
2,182
2,227
2,182
2,227
2,182
k = 0.0001
7,900
7,900
8,556
8,556
9,256
9,256
9,692
9,692
10,010
10,010
10,270
10,270
10,480
10,480
10,970
10,970
11,620
11,620
12,570
12,570
13,160
13,160
13,600
13,600
14,260
14,260
14,740
14,740
15,440
15,440
15,900
15,950
16,380
16,360
16,730
16,720
18,080
18,050
19,460
19,390
20,720
20,590
21,170
21,450
22,170
21,780
22,170
21,780
k = 0.00001
59,840
59,840
65,230
65,230
70,910
70,910
74,420
74,420
76,970
76,970

1,027
1,110
1,200
1,257
1,299
1,333
1,361
1,425
1,511
1,637
1,713
1,768
1,847
1,901
1,974
2,022
2,055
2,081
2,142
2,154
2,154
2,154
2,154
2,154

1,027
1,110
1,200
1,257
1,299
1,332
1,360
1,425
1,511
1,636
1,712
1,767
1,845
1,899
1,970
2,017
2,049
2,073
2,127
2,135
2,135
2,135
2,135
2,135

1,027
1,110
1,200
1,257
1,299
1,332
1,360
1,425
1,510
1,636
1,711
1,766
1,844
1,897
1,967
2,013
2,043
2,067
2,116
2,121
2,121
2,121
2,121
2,121

7,900
8,556
9,256
9,692
10,010
10,270
10,480
10,970
11,620
12,570
13,160
13,600
14,260
14,740
15,440
15,950
16,360
16,710
18,030
19,350
20,500
21,280
21,520
21,520

7,900
8,556
9,256
9,692
10,010
10,270
10,480
10,970
11,620
12,570
13,160
13,600
14,260
14,740
15,440
15,950
16,350
16,700
18,020
19,320
20,440
21,150
21,340
21,340

7,900
8,556
9,256
9,692
10,010
10,270
10,480
10,970
11,620
12,570
13,160
13,600
14,260
14,740
15,440
15,950
16,350
16,700
18,010
19,300
20,390
21,060
21,210
21,210

59,840
65,230
70,910
74,420
76,970

59,840
65,230
70,910
74,420
76,970

59,840
65,230
70,910
74,420
76,970

487

1
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,400
2,880
5,760
10,080
100,000

2
78,990
80,670
84,500
89,610
96,900
101,400
104,700
109,700
113,300
118,600
122,500
125,700
128,400
139,000
150,800
163,600
177,500
189,700
189,700

3
78,990
80,670
84,500
89,610
96,900
101,400
104,700
109,700
113,300
118,600
122,500
125,600
128,400
139,000
150,600
163,200
176,800
188,300
188,300

4
78,990
80,670
84,500
89,610
96,900
101,400
104,700
109,700
113,300
118,600
122,500
125,600
128,300
138,900
150,500
163,100
176,400
187,500
187,500

5
78,990
80,670
84,500
89,610
96,900
101,400
104,700
109,700
113,300
118,600
122,500
125,600
128,300
138,900
150,500
162,900
176,200
187,100
187,100

6
78,990
80,670
84,500
89,610
96,900
101,400
104,700
109,700
113,300
118,600
122,500
125,600
128,300
138,900
150,500
162,900
176,000
186,800
186,800

7
78,990
80,670
84,500
89,610
96,900
101,400
104,700
109,700
113,300
118,600
122,500
125,600
128,300
138,900
150,400
162,800
175,900
186,600
186,600

8
78,990
80,670
84,500
89,610
96,900
101,400
104,700
109,700
113,300
118,600
122,500
125,600
128,300
138,900
150,400
162,800
175,800
186,500
186,500
Table 39

Values of T2* ( ) d
0

t
1

a = 0.315
2

a = 0.284
3

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000
t

0.144
0.236
0.375
0.494
0.576
0.656
0.727
0.900
1.142
1.464
1.609
1.675
1.707

0.189
0.317
0.523
0.693
0.843
0.977
1.099
1.418
1.915
2.754
3.303
3.693
4.198
4.488
4.745
4.808

a = 0.199

a = 0.169

0.5
1
2

0.521
0.837
1.316

0.660
1.081
1.744

a = 0.252
4
= 0.1
0.247
0.425
0.722
0.977
1.205
1.417
1.614
2.144
3.029
4.711
5.991
7.036
8.684
9.948
11.780
13.030
13.910
14.540
15.850

a = 0.220
5

a = 0.189
6

a = 0.158
7

0.323
0.567
0.984
1.353
1.691
2.007
2.307
3.132
4.567
7.492
9.902
12.000
15.590
18.640
23.670
27.780
31.250
34.240
46.290
56.880
61.700

0.421
0.750
1.326
1.844
2.327
2.784
3.220
4.443
6.628
11.285
15.300
18.930
25.410
31.180
41.320
50.190
58.160
65.440
100.100
146.400
200.800
252.000
277.400

a = 0.139
= 0.2
0.860
1.431
2.353

a = 0.1096

a = 0.0798

0.556
1.003
1.799
2.527
3.213
3.867
4.498
6.283
9.541
16.700
23.070
28.960
39.780
49.680
67.710
84.090
99.290
113.600
187.100
301.100
470.700
707.900
948.100
1694.000
a = 0.050

1.127
1.913
3.221

1.601
2.757
4.721

2.577
4.493
7.811

488

1
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000
t

2
1.694
2.011
2.286
2.531
3.133
3.979
5.124
5.647
5.885
6.003

3
2.285
2.754
3.173
3.555
4.536
6.047
8.530
10.10
11.17
12.47
13.13
13.56

4
3.130
3.819
4.448
5.029
6.574
9.097
13.740
17.140
19.830
23.910
26.800
30.800
33.160
34.570
35.390

5
4.349
5.368
6.310
7.195
9.596
13.680
21.740
28.180
33.650
42.770
50.270
62.200
71.460
78.930
85.090
106.500
117.800

6
6.447
8.031
9.514
10.920
14.800
21.610
35.740
47.660
58.250
76.860
93.160
121.300
145.300
166.800
186.000
274.300
382.800
493.300
567.100

7
10.780
13.530
16.130
18.620
25.590
38.090
65.000
88.580
110.100
149.300
185.000
249.300
307.400
361.000
411.200
667.700
1063.000
1649.000
2471.000
3310.000

a = 0.126

a = 0.104

a = 0.0819
= 0.3
1.913
3.076
4.894
6.381
7.677
8.841
9.908
12.690
17.110
24.900
50.350
34.500
40.450
44.460
49.180
51.430
52.390
52.690

a = 0.0598

a = 0.378

a = 0.0157

2.621
4.265
6.888
9.080
11.020
12.790
14.430
18.800
26.030
39.720
50.220
58.900
72.860
83.880
100.500
112.600
121.700
128.700
147.500

4.064
6.680
10.940
14.560
17.810
20.810
23.620
31.240
44.240
70.200
91.380
109.800
141.400
168.300
213.600
251.300
283.800
312.400
436.300
568.700
669.000

a = 0.0494
= 0.4
3.858
5.928
9.030
11.490
13.590
15.450
17.140

a = 0.0346

a = 0.0198

9.064
15.020
24.870
33.380
41.120
48.330
55.140
73.880
106.600
174.900
233.100
285.400
379.000
462.800
611.800
744.400
865.600
978.300
1545.000
2401.000
3650.000
5383.000
7146.000
12981.00
A = 0.005

5.370
8.310
12.790
16.400
19.540
22.340
24.910

8.928
13.890
21.560
27.850
33.370
38.360
42.970

29.87
46.81
73.33
95.35
114.80
132.70
149.30

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000
t

1.219
1.901
2.914
3.701
4.358
4.926
5.429
6.664
8.396
10.750
11.830
12.320
12.560

1.488
2.362
3.697
4.762
5.673
6.478
7.204
9.049
11.830
16.240
30.350
20.650
22.600
23.440
23.780

a = 0.079

a = 0.0642

0.5
1
2
3
4
5
6

2.456
3.712
5.526
6.909
8.053
9.036
9.902

3.007
4.583
6.904
8.710
10.230
11.560
12.740

489

1
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000
t
0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100.000

2
12.020
14.970
18.990
20.870
21.750
22.210

3
15.720
20.110
26.900
30.870
33.380
36.050
37.060

4
21.450
28.120
39.410
46.970
52.530
60.110
64.860
69.760
71.490
71.860

5
31.620
42.360
61.830
76.130
87.580
105.300
118.700
137.600
150.300
158.900
164.900
175.000

6
55.220
75.440
114.100
114.400
170.100
212.900
248.400
305.900
352.000
390.500
423.400
553.900
664.600
706.500

7
193.90
269.50
420.50
544.70
654.10
845.60
1014.00
1306.00
1562.00
1792.00
2004.00
3047.00
4571.00
6734.00
9665.00
12595.00

a = 0.05

a = 0.0403

a = 0.0209

a = 0.0113

a = 0.00158

4.738
6.850
9.792
11.980
13.760
15.280
16.600
19.810
24.230
30.180
32.950
34.250
34.920

5.767
8.384
12.080
14.870
17.170
19.160
20.910
25.250
31.510
40.870
46.140
49.370
52.580
53.600

a = 0.0306
= 0.5
7.420
10.830
15.710
19.450
22.570
25.290
27.740
33.850
43.030
57.930
67.480
74.220
82.960
87.970
92.330
93.240

10.42
15.28
22.33
27.80
32.42
36.50
40.17
49.57
64.13
89.34
107.00
120.70
141.10
155.60
174.80
186.20
193.00
196.90

17.95
26.42
38.84
28.60
56.93
64.34
71.08
88.57
116.40
167.30
205.50
236.90
287.60
328.10
391.20
439.50
478.00
509.70
620.30
680.20

102.4
150.4
221.1
277.2
325.4
368.6
408.1
511.9
681.0
1003.0
1258.0
1476.0
1850.0
2169.0
2713.0
3178.0
3590.0
3965.0
5757.0
8281.0
11745.0
16307.0
20775.0
34975.0
Table 40

Values of

T3 ( )d
*

t
1

= 0.2
2

= 0.3
3

0.5
1
2
3
4

17.570
19.060
20.740
21.700
22.370

17.52
18.94
20.47
21.29
21.82

= 0.4
= 0.5
4
5
K = 0.1
17.49
17.47
18.87
18.83
20.32
20.23
21.06
20.93
21.52
21.35

= 0.6
6

= 0.7
7

= 0.8
8

17.460
18.800
20.180
20.840
21.230

17.45
18.78
20.13
20.78
21.15

17.45
18.77
20.11
20.73
21.10

490

1
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000

2
22.880
23.290
24.170
25.150
26.140
26.400
26.470
26.516
26.520
26.521
26.521
26.521
26.521
26.521
26.521
26.521
26.521
26.521
26.521

3
22.20
22.48
23.02
23.50
23.79
23.81
23.81
23.81
23.81
23.81
23.81
23.81
23.81
23.81
23.81
23.81
23.81
23.81
23.81

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000

133.50
144.69
157.73
165.29
171.35
175.93
179.75
188.68
200.54
217.70
227.25
233.75
242.53
248.07
254.85
258.50
260.62
261.90
265.01
265.38
265.39
265.39
265.39
265.39

133.40
144.53
157.44
165.17
170.79
175.23
178.91
187.39
198.35
213.32
220.82
225.43
230.88
233.66
236.25
237.11
237.41
237.52
237.69
237.69
237.69
237.69
237.69
237.69

0.5
1
2
3
4
5
6
9
15
30
45

1,027
1111.5
1209.9
1268.4
1311.2
1345.4
1373.9
1441.2
1532.3
1669.9
1752.6

1,027
1111.6
1209.6
1268.0
1310.7
1344.7
1373.2
1440.0
1530.4
1666.1
1746.7

4
5
21.84
21.62
22.06
21.80
22.45
22.12
22.73
22.31
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
22.85
22.38
k = 0.01
133.40
133.40
144.50
144.48
157.36
157.29
165.02
164.91
170.57
170.42
174.94
174.74
178.54
178.30
186.79
186.42
197.29
196.63
211.11
209.76
216.55
215.57
221.22
218.71
225.14
221.83
226.82
223.00
228.08
223.76
228.36
223.88
228.42
223.90
228.43
223.91
228.45
223.91
228.45
223.91
228.45
223.91
228.45
223.91
228.45
223.91
228.45
223.91
k = 0.001
1,027
1,027
1111.5
1111.4
1209.4
1209.3
1267.7
1267.5
1310.3
1310.1
1344.3
1344.0
1372.7
1372.4
1439.3
1438.9
1529.3
1528.6
1664.0
1662.7
1743.5
1741.6

6
21.480
21.640
21.910
22.060
22.104
22.105
22.105
22.105
22.105
22.105
22.105
22.105
22.105
22.105
22.105
22.105
22.105
22.105
22.105

7
21.38
21.53
21.77
21.90
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93
21.93

8
21.32
21.46
21.68
21.79
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82
21.82

133.40
144.47
157.26
164.85
170.33
174.62
178.15
186.17
196.20
208.87
214.29
217.10
219.76
220.67
221.21
221.28
221.29
221.29
221.29
221.29
221.29
221.29
221.29
221.29

133.40
144.46
157.23
164.80
170.26
174.54
178.04
186.00
195.89
208.26
213.41
216.01
218.38
219.14
219.56
219.61
219.61
219.61
219.61
219.61
219.61
219.61
219.61
219.61

133.40
144.45
157.21
164.77
170.21
174.47
177.96
185.87
195.67
207.81
212.78
215.23
217.42
218.09
218.44
218.48
218.48
218.48
218.48
218.48
218.48
218.48
218.48
218.48

1,027
1111.3
1209.1
1267.3
1309.8
1343.8
1372.1
1438.6
1528.2
1661.8
1740.3

1,027
1111.3
1209.0
1267.2
1309.7
1343.6
1371.9
1438.3
1527.8
1661.2
1739.4

1,027
1111.2
1208.2
1267.0
1309.5
1343.4
1371.7
1438.1
1527.5
1660.7
1738.6

491

1
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000

2
1813.0
1902.3
1966.9
2061.3
2128.4
2180.5
2222.8
2390.4
2536.4
2638.6
2685.1
2692.0
2692.5

3
1805.0
1889.8
1949.9
2035.3
2093.6
2137.0
2170.9
2293.4
2373.2
2405.2
2409.8
2409.9
2409.9

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300
360
720
1,440
2,880
5,760
10,080
100,000

7900.0
8543.0
9257.0
9696.1
10020.0
10279.0
10496.0
11003.0
11682.0
12705.0
13318.0
13768.0
14439.0
14932.0
15668.0
16208.0
16639.0
16999.0
18531.0
20172.0
21886.0
23592.0
24827.0
26080.0

7900.0
8543.0
9257.0
9696.1
10020.0
10279.0
10496.0
11003.0
11682.0
12700.0
13310.0
13758.0
14425.0
14914.0
15642.0
16174.0
16597.0
16949.0
18429.0
19963.0
21465.0
22789.0
23555.0
24104.0

0.5
1
2
3
4
5
6
9
15
30
45
60
90
120
180
240
300

59,840
65,248
71,358
74,932
77,514
79,553
81,247
85,191
90,455
98,281
102,922
106,304
111,320
114,985
120,433
124,422
127,605

59,840
65,248
71,358
74,931
77,514
79,553
81,246
85,190
90,454
98,278
102,918
106,300
111,313
114,975
120,417
124,401
127,577

4
5
1800.6
1798.0
1883.2
1879.2
1940.9
1935.5
2021.6
2013.4
2075.3
2064.3
2114.2
2100.4
2143.7
2127.3
2244.0
2215.0
2296.8
2255.0
2310.7
2262.7
2311.4
2262.9
2311.4
2262.9
2311.4
2262.9
k = 0.0001
7900.0
7900.0
8543.0
8543.0
9257.0
9257.0
9696.1
9696.1
10020.0
10020.0
10279.0
10279.0
10496.0
10496.0
11003.0
11003.0
11682.0
11682.0
12699.0
12699.0
13309.0
13308.0
13756.0
13755.0
14421.0
14419.0
14909.0
14905.0
15633.0
15627.0
16161.0
16153.0
16580.0
16569.0
16928.0
16914.0
18381.0
18351.0
19860.0
19796.0
21252.0
21122.0
22382.0
22139.0
22936.0
22583.0
23243.0
22789.0
K = 0.00001
59,840
59,840
65,248
65,248
71,358
71,358
74,931
74,931
77,514
77,514
79,553
79,553
81,246
81,246
85,190
85,190
90,453
90,453
98,277
98,277
102,916
102,915
106,297
106,295
111,309
111,307
114,970
114,967
120,409
120,404
124,390
124,383
127,563
127,554

6
1796.3
1876.6
1932.0
2007.9
2056.9
2091.3
2116.5
2196.5
2229.6
2234.6
2234.7
2234.7
2234.7

7
1795.1
1874.7
1929.4
2004.0
2051.8
2084.9
2109.0
2184.0
2212.9
2216.7
2216.8
2216.8
2216.8

8
1794.1
1873.2
1927.5
2001.1
2048.0
2080.3
2103.5
2175.1
2201.4
2204.5
2204.5
2204.5
2204.5

7900.0
8543.0
9257.0
9696.1
10020.0
10279.0
10496.0
11003.0
11682.0
12699.0
13308.0
13754.0
14418.0
14903.0
15623.0
16148.0
16562.0
16906.0
18332.0
19755.0
21037.0
21983.0
22363.0
22520.0

7900.0
8543.0
9257.0
9696.1
10020.0
10279.0
10496.0
11003.0
11682.0
12699.0
13308.0
13754.0
14417.0
14902.0
15621.0
16145.0
16558.0
16900.0
18319.0
19726.0
20978.0
21877.0
22218.0
22348.0

7900.0
8543.0
9257.0
9696.1
10020.0
10279.0
10496.0
11003.0
11682.0
12699.0
13308.0
13753.0
14416.0
14901.0
15620.0
16142.0
16554.0
16896.0
18303.0
19704.0
20934.0
21801.0
22117.0
22230.0

59,840
65,248
71,358
74,931
77,514
79,553
81,246
85,189
90,453
98,276
102,915
106,294
111,305
114,965
120,401
124,379
127,549

59,840
65,248
71,358
74,931
77,514
79,553
81,246
85,189
90,452
98,276
102,914
105,294
111,304
114,963
120,398
124,376
127,545

59,840
65,248
71,358
74,931
77,514
79,553
81,2476
85,189
90,452
98,276
102,914
106,293
111303
114,962
120,397
124,373
127,542

492

1
360
720
1,440
2,880
5,760
10,080
100,000

2
130,268
141,656
154,141
167,823
182,744
195,357
211,967

3
130,233
141,581
153,980
167,475
181,990
193,966
209,097

4
130,216
141,544
153,899
167,298
181,608
193,256
207,617

5
130,206
141,522
153,850
167,192
181,377
192,826
206,718

6
130,199
141,507
153,818
167,121
181,223
195,538
206,120

7
130,194
141,496
153,795
167,071
181,113
192,335
205,700

8
130,190
141,488
153,778
167,033
181,032
192,184
205,393

Calculation results are shown in Table 41 and in Figure 85. Clearly, the
memory function T1() displays the course of stress relaxation better than T2() does.
Therewith, the correlation coefficient is close to 1. Consequently, in this case, the
limiting stage of the process in the initial stage of the stress relaxation curve is the
interaction of relaxants and their transition into a non-relaxing material. Table 41 also
shows that the value k*, proportional to the rate constant of interaction k, is constant
at different temperatures. The main contribution to the rate of relaxation is made by
the value kBmi/S0, proportional to the number of kinetic units mi which cause
relaxation. These values naturally decrease as temperature increases. Naturally, more
accurate determination of parameters of memory functions T1() and T2() requires
longer experiments.

Figure 85. Dependences of on

*
T1 ( )d (a) and
0

T2 ( )d
*

(b) for bulky specimens of PMMA

(deformation is 2.2%) at T, K: 294 (1), 313 (2), 323 (3) and 333 (4)
Table 41
Values of parameters k*, , Ai = 0S0/(kBmi ), kBmi/S0, a,, memory functions Tl() and T2(), initial
stresses 0, and correlation coefficients r for poly(methyl methacrylate) with deformation equal
to 0 = 2.2%
T2(); a = 0.05 min-1; = 0.5
T1(); k*=0,0001 min-1; = 0.2
4
T, K
k B m1
k B m2
A2,
Al10 ,
r
r
0, MPa
0, MPa
S0
MPa
MPa
S0
294
1.6
49.64
310,250
0.998
2.272
40.39
17.78
0.991
313
2.3
52.29
227,350
0.999
3.197
39.33
12.30
0.995
323
1.8
38.25
212,500
0.999
2.465
28.20
14.44
0.991
333
2.3
39.25
170,650
0.999
3.177
26.31
8.28
0.991

493

In conclusion, one practical problem should be discussed, which is associated


with the reliability of determination of relaxation parameters by approximation of
stress relaxation curves.
t

For this purpose, let us consider dependences of k * T3* ( ) d on lgt for the

0
*
memory function T3 ( ) d , shown in Figure 86. Clearly, for every chosen value the

rate constants of the reaction k* for different values of in initial stages coincide first,
and then diverge. The duration of the relaxation process, at which divergence of these
curves depends on k*: the higher this value is, the shorter is duration of the process of
t

the onset of divergence of k * T3* ( ) d dependences on t. This area of t and k*


0

(cross-hatched in Figure 87) characterizes the duration of the relaxation process,


which must be realized in the experiment in order to obtain reliable values of which
characterize the reaction order n, because = 1/(n 1). The experiment performed at
shorter values of t, allows no reliable determination of the reaction order. Such
experiment enables us to determine the rate constant only. For example, if the rate
constant equals 0.1 min1, the duration of a relaxation process performed in
experiments must exceed 2 min; at k* = 0.01 min1 this duration is already 60 min,
etc.

Figure 86. Dependences of k

T3 ( )d
*

on lgt when k* = 0.1 (1); 0.01 (2); 0.001 (3); 0.0001 (4),

0
and 0.00001 (5).

Hence, the minimal time, during which the stress relaxation experiment must
be conducted for reliable determination of the process parameters, may be determined
with the help of Figure 87. Actually, a reduction of the rate constant of the process
must cause a significant increase of the experiment duration.
The above-discussed procedure of approximation of stress relaxation curves
(t) is true for the case of linear mechanical behavior of polymeric materials, when
the parameters of the process are independent of its duration and deformation value.
We should dwell on the possibility of description of non-linear relaxation processes,
which are most typical of polymeric materials, even at low strains.

494

Figure 87. Logarithmic dependence of the minimum duration of relaxation process t on k* at which
reliable determination of parameter is possible.

At the present time, the most wide-spread method of approximation of stress


relaxation curves in the non-linear area of mechanical behavior is the one based on the
main cubic theory by Iliyushin [73]. According to ref. [73], the relaxation modulus
Er(t) = (t)/0 is first approximated in the linear area of viscoelasticity and then, by
introduction of one more parameter and application of the same relaxation memory
function but with different parameters, relaxation curves in the non-linear area are
approximated. Another method is based on the application of equations containing a
fractional exponent of the time degree of magnitude, this index being assumed to be
dependent on the value of deformation [220] which is maintained constant during the
relaxation process. In both cases, although good coincidence of the experimental and
calculated curves is reached, the physical meaning of the introduced new parameters
is not disclosed.
In this section, let us consider an approach to description of stress relaxation
curves in the non-linear area with the help of physically stated parameters included in
the relaxation memory function (XI.9).
It should be noted that the memory function (XI.9) has been obtained on the
basis of the mixing entropy of only two types of units relaxants and non-relaxants; if
the number of types of kinetic units making a significant contribution to the relaxation
process exceeds two, several rate constants of interaction of relaxants of various types
appear with respective several values of k*. Finally, this will lead to the appearance of
a spectrum of interaction constants, which is analogous to the spectrum of relaxation
times. However, numerous tests have indicated that to describe the stress relaxation
curves, a single rate constant of relaxants interaction k* is enough even at a significant
duration of the process t. This means that although relaxants may be of different
types, only one of them makes a substantial contribution to the relaxation process.
That is why in further discussion, transiting to the description of the approximation
procedure of relaxation dependences in the non-linear area of the mechanical behavior
of polymers, we shall use the memory function T1().
The meaning of such approximation is the following [15]. Let us write down
an expression for the temperature dependence of the rate constant:
k * = k0* exp(U / RT ) ,

(XI.38)

where k0* is the pre-exponential multiplicand; U is the activation energy of the


interaction process; R is the universal gas constant; T is the absolute temperature. It is

495

common knowledge that the free volume of polymers increases during their
deformation (in this case, the free volume means the empty volume, which represents
the difference between the real volume of the polymeric substance and the Van-derWaals volume of atoms occupied by them in the polymeric substance). At significant
deformation of solid (glassy and crystalline) polymers, the free volume increases up
to a very high value, which simplifies significantly the jump-over of kinetic units
from one position into another. This very fact leads to induced elasticity, i.e. to
induced softening of the material. That is why if we assume that the activation energy
of relaxants interaction decreases as mechanical stress increases, at quite high stress
values this may cause appearance of an excessive free volume. Based on this
phenomenon, the expression for the temperature dependence of stress relaxation time
is deduced [1, 65]. Hence, it may be presented
U 0 r
U 0 Er 0
*
k * = k0* exp
,
= k0 exp
RT
RT

(XI.39)

where Er is the relaxation modulus; U0 is the initial interaction energy of relaxants;


r is the relaxing stress; 0 is the constant deformation; is the fluctuation volume in
which the elementary act of relaxants interaction proceeds.
Therewith, for the linear area of mechanical behavior when stress is not so
high yet to form an excessive free volume, let us assume the value = 0, i.e. the rate

constant k * = k0* exp( U 0 RT ) and does not depend on mechanical stress. As the
given deformation 0 increases, a moment appears when a large excessive free volume
forms that greatly simplifies interaction of relaxants and accelerates the relaxation
process. This, from the positions considered, is the transition to the non-linear
behavior. In this case, the value k* is not constant, but becomes dependent on the
relaxation modulus according to expression (XI.39). Taking this into account, we can
approximate the stress relaxation curves in the non-linear area and carry out
simultaneous determination of excessive fluctuation volume , in which the
elementary act of relaxants interaction proceeds.
Before describing the procedure of approximation of relaxation curves with
the help of the suggested approach, let us rewrite the Boltzmann equation in the
following form:
t

E (t ) = E0 E0 T ( ) d ,

(XI.40)

where E0 is the initial modulus appearing after the instantaneous assignment of


deformation; T() is the relaxation memory function. Experiment indicates that the
best approximation of relaxation curves for glassy polymers is reached at the
application of the memory function T1(), which we will use in future discussion.
Substituting the memory function T1() into equation (XI.40), we obtain:
t

E S
E (t ) = E0 0 0 T1* ( ) d ,
k B m1
0

(XI.41)

496

where T1* ( ) is the variable part of the memory function T1(), described by equation
(XI.30).
For the case of stress relaxation in the non-linear area of mechanical behavior,

U 0 Er 0
k 0* exp

RT

1 +

(XI.42)

It follows from comparison of equations (XI.8) and (XI.42) that the


component k 0* exp( U RT ) , independent of the relaxation modulus Er, corresponds
to k* in equation (XI.8) for the linear part of the mechanical behavior. Hence, it may
be indicated that

E
k * exp r 0
RT
1 +

(XI.43)

The procedure of approximation is concluded in determination of the value ,


at which the value of function (), which is the sum of the squares of deviations, is
minimal:
n

( ) = Ei, calc Ei, exper 2 ,


i =1

where n is the number of experimental points; Ei,calc and Ei,exper are the values of the
relaxation modulus calculated by equation (XI.42) and determined experimentally,
respectively.
The calculation algorithm is the following. The values of relaxation moduli for
the experimental stress relaxation curves are stored in the computer sequentially in the
ascending order of deformation values 0. Each introduced curve, except the first one,
is compared with the previously introduced curve. If every value of the modulus of
the newly introduced curve at one and the same relaxation time is lower than that of
the averaged curve, and the mean arithmetic value of relative deviations exceeds 10%,
such a curve is assumed to be related to the non-linear area of the mechanical
behavior. Then, for the averaged curve, the relaxation parameters for the linear area
are calculated by the above-described method and, basing on them, approximation of
the case related to the non-linear area is performed.
Search for the minimum of the function () is conducted by the method of
the reverse variable step, and integration of the relaxation memory function is
performed by the Simpson method with the given accuracy (usually, 0.001%).

497

Function participating in the integrand (XI.41) contains the relaxation


modulus Er which, as the function itself (XI.43), depends on time . However, at
numerical integration of the relaxation memory function (XI.9), the relaxation
modulus is taken in the area between two experimentally determined points ti and ti+1
as a constant value equal to Ei+1,exper, which, as shown below, does not affect the
approximation results.
It must be noted that it follows from equations (XI.9) and (XI.41) that at the
point = 0 = 1, and the difference ( 0) equals 0.9999999999, which is
approximated by the computer as 1 and considered as an error, because the argument
of the natural logarithm becomes 0. To eliminate this difficulty, the Boltzmann
equation may be rewritten in the following form:
t

E (t ) = E1 E0 T ( ) d ,
t1

where t1 and E1 are values of time and relaxation modulus, respectively, for the initial
point of the experimental curve. Simultaneously, such presentation significantly
reduces the working time of the program, because there is no need for integration of
the steepest (starting) part of the relaxation curve.

XI.2 Sorption and swelling processes


Let us apply the above-considered apparatus for describing relaxation
phenomena in polymers to the processes of sorption and swelling. The point is that
these processes proceed not only by filling in of separate pores in a polymeric
substance, but also cause conformation rearrangements of macromolecules, i.e. are
accompanied by relaxation processes. This idea has been put forward long ago in a
series of works [63, 67, 71]. However, the detailed analysis of this process with regard
to the relaxation mechanism of sorption and swelling was performed in refs. [10, 72].
This approach makes it possible to identify mechanism of the sorption of vapors by
polymers. As mentioned above, the mechanism is associated with the fact that the
penetration of vapors of low-molecular liquids into a polymer causes a rearrangement,
sometimes significant, of submolecular organization which, in turn, affects the
kinetics of further sorption and diffusion.
Let us perform analysis on the basis of solving a system of differential
equations [120]:
f

= 0 ; v
= 0,
x
x

(XI.44)

where f is the force affecting the diffusing particle; v is the number of diffusing
particles; is the velocity of movement of diffusing particles; D is the diffusion
coefficient; P is the osmotic pressure equal to (v/N)RT; N is the total number of
particles in the system; R is the universal gas constant; T is the absolute temperature.
From equation (XI.44) with regard to the expression for P, it is obtained that

498

D(t ) =

RT (t )

.
N
f

(XI.45)

Since the polymer is a viscoelastic medium, movement of the diffusing


particle in it, determined by its velocity v, depends on kinetic relaxation processes of
the medium. To put it differently, it is necessary to connect the rate of diffusing
particle, which becomes dependent on time, v(), as the particle penetrates into the
viscoelastic substance, with the force f participating in the system of equations
(XI.44).
Let us perform general analysis of sorption and swelling processes using the
hereditary BoltzmannVolterra theory and choosing the above-mentioned memory
function for describing the creep of polymeric substances.
Because resolvents of memory function (XI.9) and (XI.18) have not as yet
been found, the same memory function (XI.9) and (XI.18) but with the parameters
different from those suitable for description of stress relaxation, may be used in the
description of creep of polymers.
With the help of memory function (XI.9) and (XI.18), creep is described by
the correlations:
t

(t ) = 0 1 + T1 ( )d ;

(t ) = 0 1 + T2 ( )d ,

(XI.46)

(XI.47)

where (t) is the deformation developed up to the moment of time t; 0 is the elastic
(instantaneous) deformation.
If the process is limited by the rate of interaction of relaxants, the creep is
described by equation (XI.46), and if it is limited by diffusion of inhomogeneities in
the material, the creep is described by equation (XI.47).
Let us analyze sorption and swelling with the help of a device applied in ref.
[72] using memory functions T1() and T2().
The value will equal

( ) =

d
[ ( )l0 ],
d

(XI.48)

where l0 is the initial length typical of the sample.


Using expressions (XI.46) and (XI.47), we obtain:

or

() = 0l0T1()

(XI.49)

() = 0l0T2().

(XI.50)

From correlations (XI.49) and (XI.50), we get

499

l f
( ) = 0 T1 ( )
SE

(XI.51)

l f
( ) = 0 T2 ( ) ,
SE

(XI.52)

or

where S is the cross-section of the sample; E is the instantaneous elasticity modulus; f


is the force loading the sample.
Substituting correlations (XI.51) and (XI.52) into expression (XI.46), we
obtain
D( ) =

RT l0
T1 ( )
N SE

(XI.53)

D ( ) =

RT l0
T2 ( ) .
N SE

(XI.54)

or

Solving the equation of diffusion with the variable diffusion coefficient D(),
as it was made in ref. [72], for a plate of thickness l, the kinetic dependence of the
relative amount of sorbed substance may be calculated:

2n + 1 2 RT
2n + 1 2

M (t )
1

=
f * (0) exp
exp

M () n = 0 ( 2n + 1) 2
l N
l

RT *
f (t )
N

2n + 1 2 RT

exp
(
0
)

2
l N

n = 0 ( 2n + 1)

(XI.55)

2n + 1 2 RT

exp
f * () ,

l N


where
t

l0

f (t ) =
T1 ( ) d + B1

SE

(XI.56)

l
f * (t ) = 0 T2 ( ) d + B2 .
SE

(XI.57)

or

Therewith,
l
f * (0) = 0 B1 or
SE
l
f * () = 0 (A1 + B1 ) or
SE

l
f * (0) = 0 B2 ,
SE
l
f * () = 0 (A2 + B2 ) ,
SE

(XI.58)
(XI.59)

500

where

A1 = T1 ( )d
0

Because

or A1 = T2 ( )d .
0

RT *
f (t ) << 1 , then on the basis of equation (XI.55) it may be
N

written that
M (t )
f * (0) f * (t )
=
.
M () f * (0) f * ()

(XI.60)

Substituting expressions (XI.56), (XI.58) or (XI.57), (XI.59) into equation


(XI.60), we obtain, respectively:
t

T1( )d
M (t ) 0
=
M ()
A1

(XI.61)

or
t

T2 ( )d

M (t ) 0
=
M ( )

A2

(XI.62)

In the case of desorption or syneresis, kinetic equations are reduced to the


form:
t

M (t )
= 1 T1 ( ) d
M0

(XI.63)

or
t

M (t )
= 1 T2 ( ) d ,
M0

(XI.64)

where M0 is the initial mass of the swelling sample.


Taking into account that relaxation memory function (XI.9) and (XI.18)
contain constant values, which may be taken out of the integral, we obtain from
expressions ((XI.61) and (XI.62):
t

T1 ( )d
*

M (t ) 0
=
M ( ) *
T1 ( )d
0

or

(XI.65)

501

T2 ( )d
*

M (t ) 0
=
,
M ( ) *
T2 ( )d

(XI.66)

where T1* ( ) and T2* ( ) are variable parts of memory function T1() and T2(),
respectively.
To confirm experimentally the possibility of describing sorption with the help
of expressions (XI.65) and (XI.66), measurement of kinetics on free films and coating
based on cured epoxy resin ED-20 were performed in ref. [10].
Figure 88 indicates the kinetic curves of sorption measured on samples of
cured epoxy resins in the form of both free films and coatings on an aluminum
substrate with various types of surface treatment.
The calculation results are shown in Table 42. Calculations of the kinetic
parameters of sorption were performed by equations (XI.65) and (XI.66) using
t

tabulated values of integrals

*
T1 ( )d and
0

T2 ( )d , shown above. If equations


*

(XI.65) and (XI.66) describe properly the process of sorption, graphs in coordinates
t

M(t) T1* ( ) d or M(t) T2* ( ) d must represent straight lines starting from the
origin of coordinates, and the tangents of them are equal to M () A1* or M () A2* .
This requirement is fulfilled by equation (XI.64), which is clearly seen in Figure 89.
t

Experimental points fit well the calculated straight line in M(t) T2* ( ) d
0

coordinates which yields from the origin of coordinates. The correlation coefficient
varies from 0.998 to 0.996. Parameters of the memory function T2() are shown in
Table 42.
Table 42

Type of sample
Free film
Cover on degreased
aluminum surface
Cover on treated
aluminum surface

Kinetic parameters of sorbing processes


Parameters of memory function T2()
10
D10 ,
(time in min)
2
cm /s
a

4.361
0.0346
0.4
3.451
0.0209
0.5
2.707

0.1040

0.3

Rate of change Mt/M0 at t


1 min
0.0297
0.0427

60 min
0.00395
0.00414

0.0653

0.00433

Description of kinetic curves of sorption using the memory function T1()


indicated [10] the worst coincidence of the calculated and experimental values of
M(t). Remember that the memory function T1() is valid if the course of the relaxation
process is limited by the rate of relaxants interaction. When this process is limited by
their diffusion in the material (i.e. self-diffusion), the memory function T2() is valid.

502

As the calculations performed have shown [10], in the case of sorption of vapors of
low-molecular liquids, the process is described well with the help of the memory
function T2(), i.e. self-diffusion of relaxants is the limiting stage of the process lying
in the basis of sorption. The Table also shows the diffusion coefficients calculated by
the usual Fick equation. Table 42 indicates the rate of change of the relative mass
overweight {d[M(t)/M()]}/dt, which is calculated on the basis of correlation (XI.62):
M (t )
1 *
v = d
dt = * T2 ( ) ,

M
(
)
A2

(XI.67)

where the value A2* is taken from Table 39. If the parameters of the memory function
T2* ( ) are known, it is easy to calculate v at various moments of time t.

Figure 88. Kinetic curves of water vapor sorption at p/ps = 0.08: 1 free film of ED-20; 2 ED-20
coating degreased aluminum foil; 3 ED-20 coating aluminum foil processed with
-aminopropyltriethoxysilane.

Figure 89. Dependence of Mt/M0 on T2* ( ) d .


0

503

Hence, the method of description of the kinetic curves of sorption, which takes
into account the relaxation type of this process, enables kinetic dependences to be
approximated with high accuracy using the new relaxation memory function. If the
Fick equation with the constant diffusion coefficient is used for this purpose, the
adequate description cannot be carried out, and the process may be approximated only
over a short initial period of time.

Chapter XII. Solubility of polymers

XII.1 Specific cohesive energy of organic liquids and polymers


Hildebrand solubility parameter
To predict the solubility of polymers in various organic solvents, as well as for
preliminary estimation of the miscibility of polymers with each other or with
plasticizers, the characteristic as the solubility parameter is often used. This
characteristic has been introduced by Hildebrand for description of solutions of nonelectrolytes. The Hildebrand solubility parameters is determined from the correlation

E0
,
V

(XII.1)

where E0 = H0 RT; H0 is the latent heat of liquid evaporation; R is the universal


gas constant; T is the absolute temperature; V is the molar volume of the liquid.
The square of the solubility parameter represents specific cohesive energy of
the liquid, i.e. the value of cohesive energy divided by the molar volume:

2 = E0/V.

(XII.2)

These notions are also applied to polymers, therewith, estimations are


indicated per repeat unit. In this case, the problem is that the value may be
experimentally determined only for low-molecular liquids evaporating without
decomposition. For polymers which cannot be evaporated without decomposition,
values of are determined by indirect methods or calculated from the constants of
energies for separate atoms and groups of atoms [141, 150, 206].
Taking into account the nature of molecules in liquids and polymers leads to
the following equation for calculating the specific cohesive energy [25]:

2 =

N A Vi
i

Ei*
i

N A Vi

(XII.3)

where E* = kE0 is the cohesive energy of the liquid or repeat unit of the polymer,
reduced by the number of times by which the Van-der-Waals volume of the molecule
(or unit) is smaller than the molar volume; k is the coefficient of molecular packing of
the liquid or polymer.
The value of E* is additive and is represented as E* =

Ei* , where Ei*


i

is the contribution of each atom and type of intermolecular interaction to E*.


Values of Ei* are shown in Table 43, which indicates the appropriate notes.
With the help of them, the solubility parameter may be calculated for many
polymers with different chemical structures.

505

Table 43
*
Values Ei for different atoms and types of intermolecular interaction
Atom and type of intermolecular interaction
Designation
Ei* , cal/mol
1
2
3
Carbon
550.7
*
E C
Hydrogen
47.7
*
E H
Oxygen
142.6
*
EO
Nitrogen
1205.0
*
E N
Fluorine
24.2
E F*
Sulfur
1750.0
ES*
Chlorine
222.7
*
E Cl
Bromine
583.0
*
E Br
Iodine
1700.0
EI*
Double bond
323.0
E *
Aromatic cycle (skeleton)
713.0
*
Dipoledipole interaction
Dipoledipole interaction in non-polar aprotic
solvents of:
amide type
dimethylsulfoxide type
Hydrogen bond
Specific interactions in the presence of =CCl2
group
In stressed three-five-component cycles in the
presence of O atom
Hydrocarbon radical isomerism

E d*
E a*, N
E a*, S
E h*
E =* CCl 2

1623.0

1623.0
2600.0
3929.0
2600.0

*
EO,
c

2430.0

Ei*

412.0

Notes:
*
1) The constant E is introduced in the presence of double bonds, not participating in polar groups.
*
2) The constant E is introduced in the presence of aromatic cycles in the amount appropriate to
these cycles.
*
3) The constant E d is introduced in all the cases in the presence of a polar group of any type; in the
case of chlorinated compounds in the presence of two and more Cl atoms, added to the same carbon
*
atom, two E d should be introduced.
*
4) The constant E a , N is introduced at the calculation of for dipolar aprotic solvents of the amide
type; therewith, the usual dipoledipole interaction at the sacrifice of polar groups is taken into account
by introduction of the appropriate amount of constants.
5) The same as for the previous case, but for solvents of the dimethylsulfoxide type.

506

*
6) The constant E h is introduced in the presence of a hydrogen bond of any type.
*
7) The constant E = CCl takes into account the specific interaction in the presence of the
2

Cl
C

Cl
*
group; therewith, the constant E d is not introduced.
*
8) The constant EO, c takes into account appearance of polarity in stressed 35-component cycles

containing the heteroatom O.


*
9) The constant Ei is introduced at transition from normal hydrocarbons to isomers.

The experimental methods of determination of solubility parameter conclude


in the following. The value of intrinsic viscosity of the polymer is evaluated in a
selection of solvents with different values of the solubility parameter. Further on,
dependences of of the polymer on the parameter of solubility of the same solvent,
in which they were measured, are composed.
Figure 90 indicates the schematic representation of this dependence. The
maximum of this dependence determines the parameter of polymer solubility which,
in this case, equals the parameter of liquid solubility, in which the intrinsic viscosity
of this polymer is maximal (see Figure 90).

Figure 90. Schematic representation of dependence of intrinsic viscosity [] of polymers in different


solvents on solubility parameter of solvent p.

Another experimental method of estimating the value is the measurements of


the equilibrium swelling degree and plotting the dependence of this value on the
solubility parameter of the liquid in vapors of which the swelling degree was
measured. This dependence is analogous to the one depicted in Figure 90. It should be
noted that experimental methods of determination of are difficult and not always
reliable. For preliminary estimation of , the calculation methods are preferable.
To predict solubility of polymers, the calculated value p for the polymer is
sometime compared with experimental values s for solvents. If values for polymers
and solvents differ significantly, then no dissolution takes place. However, the
coincidence of solubility parameters of the polymer and the solvent does not yet
guarantee dissolution of the polymer in the current solvent. In practice, in the case of
coincidence of values, dissolution is observed in 50% of cases only (see below).

507

Table 44

1
2
3
4
5
6
7
8
9
10
11
12
13
14

Parameters of solubility of a series of polymers


Name
Ei* , N A Vi ,
i
i
cm3/mol
cal/mol
Poly(methyl methacrylate)
5043
58.5
Poly(ethyl methacrylate)
5689
69.0
Poly-n-propyl-methacrylate
6335
79.3
Poly-n-butyl-methacrylate
6981
89.6
Poly(methyl acrylate)
4397
48.2
Poly(ethyl acrylate)
5043
58.5
Polystyrene
5500
66.0
Polyisobutylene
2584
41.6
Polyacrylonitrile
4623
32.6
Poly(ethylene terephthalate)
10418
102.4
Nylon 6,6
18210
139.2
Phenolphthalein and isophthalic acid
27189
234.7
polyarylate
Phenolphthalein and terephthalic acid
27189
234.7
polyarylate
Phenolphthalein and 4,431397
279.2
diphenyldicarboxylic acid polyarylate

, cal0.5/cm1.5
Calc.

Exper. [54]

9.3
9.1
8.9
8.8
9.55
9.3
9.1
7.9
11.9
10.1
11.4
10.7

9.1; 9.5; 9.4


8.95

10.1
9.4
9.1; 8.6; 8.7
7.95; 7.8; 8.05

10.8

10.7

10.7

10.6

10.4

Table 44 indicates solubility parameters for polymers of various chemical


structures. The chemical structure significantly affects the value . The presence of
C O
strong polar groups of Cl,
, CN types, etc. leads to increase of the
O
cohesive energy and, consequently, of the value . The solubility parameter is most
strongly increased by hydrogen bonds appearing in the presence of OH and
HNC
groups, and by aromatic cycles, as well.
O
For copolymers, the equation for calculating the solubility parameter is
presented in the form:

1 Ei* + 2 Ei* + ... + n Ei*

i
1
i
2
i
n
2 =
,
(XII.4)


N A 1 Vi + 2 Vi + ... + n Vi


i
1
i
2
i
n

where 1, 2, , n are the molar parts of components 1, 2, , n; Vi ,

i
1

V , , Vi are the Van-der-Waals volumes of components 1, 2, ,


i

i
n
i
2

508

n; Ei* , Ei* , , Ei* are the cohesive energies of components

i
n
i
1 i
2
1, 2, , n; NA is the Avogadro number.
In the reduced form, equation (XII.4) is the following:
k =n

2 =

k Ei*

k =1
i
k
,
k =n

NA
k Vi

k =1
i
k

(XII.5)

where k and Vi are the molar part and the Van-der-Waals volume of the

i
k

k-th component, respectively; Ei* is its cohesive energy.

i
k
If it is desirable to express the solubility parameter via the solubility
parameters of the components of the copolymer, it should be written down:

1 12 Vi + 2 22 Vi + ... + n n2 Vi

i
1
i
2
i
n
,
2 =

1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n

(XII.6)

where 1, 2, , n are solubility parameters of the components 1, 2, , n.


In the reduced form, this equation is presented as

2 =

k =n

k =1
k =n

k k2 Vi

k Vi
k =1
i
k

(XII.7)

where k, Vi , and k are the molar part, the Van-der-Waals volume and the

i
k
solubility parameter for the k-th component, respectively.

509

XII.2 Solubility criterion


The problem of prediction of the solubility of polymers has been urgent for
many years. One of the methods of preliminary estimation of polymer solubility
concludes in comparing values of Hildebrand solubility parameters for the polymer
p and the solvent s. Therewith, it is assumed that if the condition p s is fulfilled,
polymer dissolution in the given solvent may be expected. However, experience
suggests that with the help of such comparison we may only neglect the solvents, in
which no dissolution of the given polymer proceeds. These are systems for which
p >> s or p << s. With the help of this estimation, it is possible to shorten
significantly the list of solvents to be tested, in which the polymer may dissolve.
Estimation and experience [128] indicate that this method makes it possible, for
example, to neglect 120130 organic liquids from the list of 160 solvents as obviously
invalid for dissolution. Therewith, the polymer is dissolved in a half of the rest of
solvents conformed to the condition p s.
Consequently, fulfilling the condition p s may not guarantee solubility of
the polymer. It is desirable to obtain a more accurate method of preliminary
estimation of polymers in relation to the solvents, for which the condition p s is
met. Let us consider in detail the solubility criterion suggested in refs. [32, 95], which
possesses quite high predictive force.
The feature of the problem of solubility lies in the fact that here, in contrast to
determination of some simplest properties, not only the chemical structure, but also
the particular supermolecular structure of the polymer should be taken into account.
Actually, it is common knowledge that a crystalline polymer is dissolved
much less readily than an amorphous polymer of the same chemical structure.
Oriented samples are also less soluble compared with isotropic samples. Probably, in
the case of isotropic samples of amorphous polymers, the supermolecular structure
may also be different; however, this problem is still the subject of discussion.
Presented at an International Conference in 1979 in London were experimental and
theoretical data on the absence of a nodular structure in amorphous polymers, the
data of electron-microscopic investigations of the surface of films and chips being
added to artifacts [142]. However, it can be hardly imagined that if the surface of the
film obtained from the solution and the surface of chip of a block sample obtained
from the melt give one and the same electron-microscopic picture of globules, this
picture is the consequence of artifacts.
It is noted that the neutron scattering method is the only one that may give
direct information on this problem, although interpretation of data is not uniform. This
discussion has been continuing. Basing on results of X-ray scattering, it is stated [189]
that a level of packing regularity does also exist in amorphous polymers.
Analyzing the problem of solubility in ref. [32], the authors used the model of
the supermolecular structure developed in refs. [92, 93], in which the supermolecular
structure of amorphous polymers is modeled as globules. Therewith, an attempt was
made in these works to justify the absence of a long period at small-angle X-ray
scattering. It is also assumed that every globule consists of globulesmacromolecules
[4, 102]. Supposing that both types of globules are connected with each other by bond
belts, let us discuss the most typical elementary event of dissolution, i.e.
decomposition of particles down to separate globular macromolecules, which is
schematically represented in Figure 91.

510

Figure 91. Schematic representation of polymer surface and elementary event of dissolution.

Here, the case of an uneven surface is shown (for example, a powder or a film
with surface roughness), when the load is applied from the side of the solvent to the
globule belt by which it is connected with neighboring globules, and detachment of
the globule from the polymeric substance and its transition into the solvent proceeds
at the sacrifice of bond belt rupture.
It should be noted that to estimate the solubility of polymers, it is necessary to
consider the change of free energy, i.e. to take into account both energetic and
entropic components. In this case, we will consider only amorphous polymers
composed of macromolecular globules, which keep their form in solution. In this
particular case, conformational selection is maintained, and the change of entropy at
transition of macromolecules into the solution will be slightly different from the ideal
case only due to the difference in sizes of solvent molecules and macromolecular
globules. That is why in this case, in the first approximation, the entropy change at
dissolution may be neglected. When analyzing the solubility of crystalline polymers
or amorphous non-globular polymers with flexible macromolecules, the entropic
component of free energy may be significant, and the energetic criterion considered is
changed.
Let us dwell on the physical assumptions used in determination of the
solubility conditions. When a polymer sample is immersed into a solvent, the globules
which exist on the sample surface will be detached first. Let us consider forces
affecting the globules. Figure 92 indicates the globule cross-section and the belt of the
globule bonding it to other globules of the supermolecular structure of the polymer.
At the moment of the polymer dipping into solvent, the solid (globule)liquid
interface is formed. Formation of the unit of this surface includes work WA
determined by the process of adhesive wetting:
WA = s-p (p + s),

(XII.8)

where p and s are surface tensions of the polymer and the solvent, respectively; s-p
is interfacial tension. Therewith, WA represents the work of adhesion, i.e. the work
required for separation of the surfaces (restoration of the initial state).
The work of adhesion induces occurrence of forces affecting the globule of the
supermolecular structure; these forces depending on the size and sign of curvature of
the surfaces which form the globule and the bond belt. These forces lead to
detachment of the globule from the rest of the polymer sample. But as soon as this
detachment happens, and the globule transits into the solvent, a fresh (new) surface of
another globule, which was closed, is formed. It is also wetted by the solvent, and the
same forces appear. When this globule is detached, the situation is repeated: new
surfaces of the previously screened globules are formed and due to the effect of
wetting forces they are subsequently transferred into the solvent.

511

Figure 92. Cross-section of the globule and the belt whereby the globule is linked to the other globules
of supermolecular structure (schematic).

Let us consider in more detail the forces affecting the globule and appearing
during its adhesive wetting by the solvent. According to the image in Figure 92, two
forces affect the initial globule of the supermolecular structure and the bond belt.
They are the forces determined by WA, i.e. by the surface tension of the solvent, and
the interfacial tension applied to the surface of the supermolecular structure globule,
disposing over the bond belt. The latter force strives to detach the globule from the
polymer (this happens due to different signs of curvature of these surfaces).
Determining the second force, it may be assumed that the surface to which the force
detaching the globule is applied is the tore surface (shaded in Figure 92). Therewith,
the wetting force intensity will be applied to the half of the tore surface. Because the
cross size of the linking belt is significantly smaller than the greater radius of the belt,
further on, all calculations are performed on typical sizes of the bond belt, the surface
of which is also assumed to be toroidal. Let us determine the conditions of polymer
dissolution. Figure 93 indicates a triangle, apexes of which locate in centers of
globules of the supermolecular structure. Therewith, [AD] = [DB] = r, [AD] = [DB]
= R, [DO] = [OE] = Rt = R/2, where r is the smaller radius of the linking tore, R is the
radius of globule of the supermolecular structure, Rt is the larger radius of the linking
tore.

Figure 93. Characteristic distances between globules (see text).

Globule 1 will be affected by the adhesive wetting force which compresses


globule to the polymer. According to the Laplace law, this force is determined from
the correlation:

512

f gl = Rt

s + p s-p
s + p s-p
= R
.
R
2

(XII.9)

Moreover, the force ft is applied to the bond belt striving to detach the globule
from the substance:

1 1
s + p s-p .
f t = 2Rr sin
2 r Rt

Since Rt = R/2,

(XII.10)

= 6 , then
2

1 1
f t = Rr s + p s - p .
r R

(XII.11)

For the globule to detach from the supermolecular structure, two conditions
must be fulfilled. The first condition is that the detaching force applied from the side
of the solvent, ft, must be greater than fgl which compresses the globule to the
substance, i.e.

( s + p s - p ) 3 8Rr > 0 .

(XII.12)

For globules, 8r/3R << 1. That is why condition (XII.12) may be presented in the
following form:
s + p > s-p.

(XII.13)

The second condition is formed in the following manner: the work performed
by the surface tension forces of the solvent A must exceed the energy of rupture of
intermolecular bonds in the bond belt, i.e. specific cohesive energy of the polymer p2
would be lower or equal to the work of the solvent on rupture of the linking tore,
2

related to the specific volume of the linking tore p* . The value A may be presented
in the form:
A = (ft fgl)maxr,
and
2

p* =

A
8r

s + p s - p 3 max .
Vt
R 4r

(XII.14)

Because 8r/3R << 1, expression (XII.14) transforms to:


2

s* = s + p s - p

)34max
.
r

(XII.15)

513

So, the second condition may be written down as follows:


2

p2 s* .

(XII.16)

Let us transform expression (XII.16):


2

s* = s2

s + p s-p
s

(XII.17)
2

*
*
where s2 is the cohesive energy density of the solvent; s* = 3 max
4r * , max
is the maximal distance between solvent molecules, which are still affected by forces
of interaction, r* is the characteristic size of the linkage of the Frenkel cluster of the
*
r . Making designation = p2 s2 and taking into
solvent, = max r * max

account that, according to refs. [147 149],

s - p = s + p 2 s p 1 / 2 ,

(XII.18)

the condition of solubility (XII.16) may be presented in the following form:


1/ 2

p
2
s

(XII.19)

where = p2 s2 ; p and s are Hildebrand parameters for the polymer and the
solvent, respectively; p and s are the surface tensions of the polymer and the solvent,
respectively;
=

4 VsVp 1 / 3

(Vs1/3 + Vp1/3 )2

(XII.20)

Here Vs and Vp are the molar volumes of the solvent and the polymer (per one unit),
respectively. Note that values are of the order of unity.
Expression (XII.19) is deduced from the condition that there is no swelling of
the polymer. Let us consider another case, i.e. when polymer swells. This means that
the solvent penetrates into a cavity between globules A, B, and C (see Figure 92). In
this case, condition (XII.15) may be formally presented in the same form, however,
the effective value (p/s)eff will play the role of the relation p/s. This effective value
will be smaller than the true value p/s, because penetration of the solvent into the
cavity between globules causes wedging and decreases force ft necessary for
detachment of globule B.
This question is discussed in detail in ref. [32]. Consequently, the following
expression is deduced for the solubility criterion:

514

< 2 2 1 + a ,

(XII.21)

where a = s-p/s; other parameters are the same as in equation (XII.19).


Recall that = p2 s2 ; is a constant; the value is calculated from
formula (XII.20). It follows from expression (XII.21) that the value amin = 1 2
exists, at which dissolution is possible (for example, at 0.95, amin = 0.1). Because
is always positive, solubility is possible at amin < a < 1. The maximal value max will
equal

max = 22.

(XII.22)

Hence, according to criterion (XII.21), solubility will be observed in the case


when the right part of criterion (XII.21) is greater than the left part, both parts of the
criterion being calculated on the basis of the chemical structure of the repeat unit of
polymer and solvent molecule.
Experimental testing of the solubility criterion (XII. 21) is performed on the
example of approximately 300 systems of the polymersolvent type. For polymeric
objects of the study, both traditional amorphous polymers (poly(methyl methacrylate),
polystyrene, polyvinylacetate, polyisobutylene, butadiene and isoprene rubbers, etc.)
and a series of heat resistant polymers of the cardo-type structure (polyarylates,
aromatic polyamides and polyimides, polyphenylquinoxaline) were chosen in ref.
[32]; about 50 organic liquids were used as solvents.
As mentioned above, solubility must be observed if calculated values

2 2 1 + a appear greater or equal to the value = p2 s2 ; therewith,

p and s values used may be both experimental and theoretical.


The value introduced above is not determined directly. However, if the value
is nearly constant, then introducing designation

= 2 1 + a ,

(XII.23)

< 2,

(XII.24)

we obtain that

i.e. in (, ) coordinates, the dependence = 2 represents a straight line yielding


from the origin of coordinates. Above it there must be a point which indicates the
absence of solubility, and below it the presence of solubility. Calculations
performed for the mentioned number of pairs indicate (Figure 94) that points
corresponding to the case of insolubility (dark points) are usually disposed above the
mentioned line, and the points corresponded to solubility (light points) are located
below it. Hence, the range of values is divided into two fields, one of which
represents the field of insolubility, and another one the field of solubility. However,
each of these fields contains certain numbers of foreign points, the fraction of which

515

is about 15%. Generally, the number of dark and light points is approximately equal,
and the value = p2 s2 varies around unity.

Figure 94. Dependence of on for different pairs of the polymersolvent type (for explanation see
text).

The dotted line in Figure 94, corresponding to the dependence = 2, is


traced in the manner that the number of foreign points above and below the straight
line would be minimal and approximately equal. If these conditions are fulfilled, the
value 2 = 1.374. Figure 94 indicates immediately that if the only condition s p is
fulfilled, dissolution may be observed in 50% of cases only.
Consequently, for systems in which the condition s p is fulfilled, according
to criterion (XII.19) or (XII. 21), solubility may be predicted with higher accuracy
equal to 85%. Taking into account that real non-solvents are preliminarily neglected,
for which s >> p and s << p, the predictive strength of the criterion may be
considered high.
Usually, solubility is observed when the surface tension of the solvent is close
to the surface tension of the polymer; the interfacial tension is then low and the value
a is also low. But if the equality s p is fulfilled, the surface tension of the solvent
is significantly lower than the surface tension of the polymer, the relation of the
interfacial tension to the surface tension of the solvent is high, a becomes high, and
becomes low. Points in the diagram (see Figure 94) appears in the left upper field.
Hence, two factors promote the increase of value: equality of the molar volumes of
the solvent and the repeat unit and the equality of the surface tensions of the polymer
and the solvent. In its turn, according to Figure 94, the increased value promotes the
polymer solubility.
It has been assumed for a long time that the like must dissolve in the like.
However, no acceptable definition of such substances has been presented as yet. From
the point of view of the solubility criterion expressed by expression (XII.19) or (XII.
21), these substances may be considered as the ones possessing similar densities of
the cohesive energy (the Hildebrand solubility parameter ), molar volumes and
surface tension coefficients. In this case, solubility and miscibility are usually
observed.

516

Figure 94 indicates that some exclusions do also exist. For example, in the
case of polyarylate,
C O
C

O
C

O
C

according to criterion (XII. 21), there should be no dissolution in chloroform, and


cyclohexanol must dissolve the polymer. In practice, the situation is opposite:
chloroform dissolves efficiently many heat-resistant polymers, and cyclohexanol
not. These very exclusions form the main part of distortion of criterion (XII. 21).
Apparently, in the case of chloroform, dissolution proceeds practically without
swelling, which confirms the preliminary experiments.
Other possible reasons of deviations from criterion (XII. 21) will be analyzed
below.
According to the above-discussed ideas, estimation of the solubility of the
polymer with the given chemical structure in one or another solvent concludes in the
following facts. For this polymer and solvent, values of the solubility parameter are
calculated from formula (XII.3). Then the value of the surface energy for polymers,
p, is calculated by equation (XIII.18) or equations (XIII.28) and (XIII.29). It is also
possible to calculate p with the help of a parachor using equation (XIII.1). The molar
volume of the repeat unit of the polymer, necessary for this purpose, is determined as

Vp =

N A Vi
i

k avg

(XII.25)

where kavg = 0.681. If polymer density, dp, is known, then Vp = M/dp, where M is the
molecular mass of the repeat unit.
The surface tension for the solvent, s, may be assumed as both experimental
and calculated from formula (XIII.11). The value is calculated from formula
(XII.20). Values of Vs and Vp required for this purpose are determined from
correlations Vp = M/dp and Vs = M/ds, where ds and dp are densities of the solvent and
the polymer, respectively. After that, the value of the interfacial tension, s-p, is
calculated from equation (XII.13), and then the value a = s-p/s. Further on, the
value 1.347 2 1 + a is calculated, and the result is compared with

= p2 s2 . If < 1.347 2 1 + a , then dissolution of the polymer in the

current solvent may be expected with 85% probability.


Let us indicate an example of calculation for the poly(methyl methacrylate)
benzene system.
CH3
CH2

C
C O CH3
O

Polymer

Solvent

517

First, it is necessary to determine solubility parameters for the polymer, p,


and the solvent, s.

*
*
*
+ 8 EH
+ 2 EO
+ Ed* = 5550.7 +
For the polymer, Ei* = 5 EC

i
p
847.7 + 2142.6 + 1623 = 5043.3 cal/mol = 21,081 J/mol (all values Ei* are taken
from Table 43).

Vi = VC,10 + VC,13 + VC,1 + VC,49 + VC,41 + VO,139 + VO,129 + 8VH,124

i
p
= 13.1 +17.2 + 5.0 + 15.9 + 20.3 + 5.8 + 3.4 + 82.0 = 96.7 3 (all numbers of atoms
correspond to the numbers in Table 3).

Substituting calculated values of Ei* and Vi into equation

i
p
i
p
(XII.3), we obtain that

p2 =

21,081
= 362 J/cm3; p = 19.0 (J/cm3)1/2.
0.6023 96.7

For the solvent,

*
Ei* = 6 E * + 6 EH
+ E* = 6550.7 + 647.7 + 713 = 4303 cal/cm3 =
C

i
s
17,988 J/ cm3.

Vi = 6VC,18 + 6VH,124 = 612.7 + 62.0 = 88.2 3.

i
s
According to these parameters, from equation (XII.3) we obtain that

p2 =

17,988
= 338.6 J/cm3; s = 18.4 (J/cm3)1/2.
0.6023 88.2

It is clear that s p, and it may be suggested that poly(methyl methacrylate)


is soluble in benzene.
The surface tensions of the polymer and the solvent should then be calculated.

Substituting values of Ei* and Vi , and m = 15 (the number of atoms

i
p
i
p
in the repeat unit of poly(methyl methacrylate)) into equation (XIII.22), we obtain that
poly(methyl methacrylate) relates to polar groups of the type 1):

p = 0.0751

21,081
(96.7) 2 / 3 151 / 3

= 30.5 dyn/cm.

518

For benzene, according to expression (XIII.11), we obtain

s = 0.0287

17,988
(88.2) 2 / 3

= 26.1 dyn/cm.

The molar volume of benzene Vs = 89 cm3/mol; the molar volume of


poly(methyl methacrylate), according to equation (XII.25), equals
Vp =

0.6023 96.7
= 85.5 cm3/mol.
0.681

Now it is necessary to calculate the value of . According to equation


(XII.20), it indicates
=

4 (89 85.5)1 / 3

(89

1/ 3

1/ 3 2

+ 85.5

= 1.

Further on, interfacial tension is calculated from formula (XII.18):

s-p = 26.1 + 30.5 21(26.130.5)1/2 = 0.17 dyn/cm.


The value of a = s-p/s = 0.17/26.1 = 0.00656.
Substituting all values of parameters of the polymersovlent system obtained
into criterion (XII.21), we obtain that

p2
s2

362
< 1.374 1.0 1.0 1.0 2 1 + 0.00656 ;
338.6

1.069 < 1.263

Because the left part of criterion (XII.21) is smaller than the right part of it, then
poly(methyl methacrylate) must dissolve in benzene, which is observed in reality.
Let us display one more example of calculation for a more complicated
polymersolvent system.
As a polymer, let us choose anilinphthalein polypyromellitimide, and
nitrobenzene as a solvent:
O
O
C

N
C

N
O

C
O
C

Polymer

Solvent

519

For the polymer,

*
*
*
Ei* = 30 EC
+ 14 EH
+ 6 EO
+ 2 E *N + 3 Ed* + 4 E * = 30550.7 +

i
p
1447.7 + 6142.6 + 21205 + 31623 + 4713 = 28,175 cal/mol = 117,733 J/mol;

Vi = 4VC,64 + 8VC,19 + 14VC,18 + 2VC,21 + VC,34 + VC,49 + 14VH,124 +

i
p
5VO,139 + VO,129 + 2VH,144 = 414.1 + 88.4 + 1412.7 + 210.2 + 7.9 + 15.7 +
142.0 + 55.8 + 3.4 + 20.9 = 407.6 3.
In accordance with equation (XII.3), we obtain:
117,773
= 479.7 J/cm3; p = 21.9 (J/cm3)1/2.
p2 =
0.6023 407.6
For the solvent,

*
*
*
Ei* = 6 EC
+ 5 EH
+ E * + Ed* = 6550.7 + 547.7 +
+ E *N + 2 EO

i
p
1205 + 2142.6 + 713 + 1623 = 7,369 cal/mol = 30,802 J/mol;

Vi = 5VC,18 + VC,21 + 5VH,124 + VH,150 + 2VO,140 = 512.7 + 10.2 + 52.0

i
p
+ 7.0 + 27.2 = 105.1 3.
According to equation (XII.3), we obtain:
30,802
= 486.6 J/cm3; s = 22.06 (J/cm3)1/2.
s2 =
0.6023 105.1
It is clear that s p, i.e. the present polyimide, principally, may dissolve in
nitrobenzene.
Then, the surface tension of the polymer and the solvent is calculated. The
number of the atoms in the repeat unit of the present polyimide m = 52. Substituting
all parameters of polyimide into equation (XIII.18), we obtain that

p = 0.0751

117,773
( 407.6) 2 / 3 521 / 3

= 43.1 dyn/cm.

According to expression (XIII.11), for nitrobenzene we get:

s = 0.0287

30,802
(105.1) 2 / 3

= 39.7 dyn/cm.

The molar volume of nitrobenzene Vs = 103 cm3/mol; the molar volume Vp of


polyimide, according to equation (XII.25), equals
Vp =

0.6023 407.6
= 360.5 cm3/mol.
0.681

Let us calculate the value of by equation (XII.20):

520

4 (103 360.5)1 / 3

(1031/ 3 + 183.61/ 3 )2

= 0.9576.

The value of the interfacial tension, calculated from formula (XII.18), equals:

s-p = 43.1 + 39.7 20.9576(43.139.7)1/2 = 3.58 dyn/cm.


The value of a = s-p/s = 3.58/26.1 = 0.0901.
Substituting all values of parameters of the polymersolvent system into
criterion (XII.21), we obtain that

p2
s2

479
< 1.374 0.9576 0.9576 0.9576 2 1 + 0.0901 ;

486.6
0.986 < 1.149.

According to criterion (XII.21), the present polyimide will dissolve in


nitrobenzene, which is observed in practice [6].

XII.3 Influence of molecular mass and degree of macromolecule


orientation on solubility
Solubility criterion (XII.21) is true for the case of isotropic amorphous
polymers with a globular supermolecular structure. Moreover, this criterion does not
take into account the influence of the polymerization degree of the polymer on
solubility, although it is common knowledge that it may be significant at transition to
greater molecular masses. Ref. [95] indicates the attempt to take into account the type
of supermolecular structure and the polymerization degree of polymers on their
solubility, as well as to set a connection between the FloryHuggins theory and the
chemical structure of the polymer and the solvent.
Solubility criterion (XII.21) contains constant , which is described by the
correlation:

p r
= max s ,
s
rp
max

(XII.26)

p
where max
is the maximal relative deformation of intermolecular bonds in the
s
polymer at the moment of their rupture and polymer transition into the solvent; max
is the maximal deformation of the liquid, i.e. the deformation, at which the continuity
is distorted; rs is the typical size of the bond of Frenkel clusters in the solvent; rp is the
lower radius of the linking globule for the polymer.
Imagine that in the area of deformation the polymer and the solvent behave
themselves as elastic bodies [77] characterized by elasticity moduli Ep and Es,
respectively. Then

521

( )

( )

2
2
p
s
p2 = Ep max
2 ; s2 = Es max
2,

(XII.27)

and the value will obtain the following form:

p2
s2

( ).
s
)2
Es ( max

2
p
Ep max

(XII.28)

At the moment of globule detachment, tensions in the globule and the solvent
p
s
= Es max
. That is why expression (XII.24) is reduced to
are equal, i.e. = Ep max
the form:

rp
rs

2 .

(XII.29)

1/2
Because rp = n1/2
p a p and rs = ns as (where ap and as are the sizes of repeat
units of the polymer and the solvent, respectively; np and ns are numbers of units in
the connecting globule in the polymer and of solvent molecules, respectively), then
taking into account the rotary-isomeric theory [58], we obtain

2/3

2/3

Ep
Vi exp RT ;

i
p

ap2 =

E
Vi exp RTs ,
i
s

as2 =

(XII.30)

(XII.31)

where Vi,p and Vi,s are Van-der-Waals volumes of the i-th atoms participating in the
repeat unit of the polymer and the molecule of the solvent, respectively; Ep and Es
are differences in the energies of rotary isomers of the polymer and the solvent. With
this regard, expression (XII.29) may be reduced to the form:


Vi
1/ 2

ns i
s

q2
np


Vi
i
p
Ep Es
.
where q = exp

2 RT

(XII.32)

522

The value ns will be calculated (with accuracy up to a constant) from the


condition that ns equals the number of solvent molecules, which cover the connection
globule by a single layer. Then
Vp

ns = 6 np

Vs

2/3

ns
6 Vp

=
1/3
np ns Vs

2/3

(XII.33)

Here Vp = Vi ; Vs = Vi . Substituting expression (XII.33)

i
p
i
s
into formula (XII.32), we obtain that
1/ 2

q 2
n1/3
p

(XII.34)

1/ 2

6
, and q by its meaning equals to (q = ).
i.e. =
n1/3
p
Taking into account that not the whole connecting globule is washed over by
the solvent (a part of it is a part of the globulemacromolecule), we may present the
expression for in the following form:
1/ 2

6
= 1/3
n
p

C 1 ,

(XII.35)

where C is the part of the connecting globule surface washed over by the solvent. C is
the constant, and its value may be determined from the condition = 0.687 at np = 24
[94] (at the polymerization degree N = 104).
Then C = 2.1, and solubility condition (XII.34) will reduce to the form
1/ 2

q
n1/3
p

The value np is the function of the polymerization degree of the polymer only.
Therefrom, it follows that the higher the molecular mass of the polymer, the worse is
its solubility.
1/ 3
Let np = N, where = 0.24103 at N = N0 = 104 [94]. Then n1/3
p = 0.13N

and condition of solubility (XII.35) may be presented in the form

523

2.1

N1/ 6

(XII.36)

Because calculations of all constants in correlation (XII.36) were performed


for N = 104, and for real polymers N may display different values, in more general
form, expression (XII.36) must be present as follows:
1/ 6

10 4

q 1.374
N

(XII.37)

where 2.1/(104)1/6 = 1.374 is the value of the coefficient at in expression (XII.37) at


N = 104.
The value N0 corresponded to 2 = 1.374 is obtained from the graphic shown
in Figure 94, which is depicted with no regard to molecular masses of the polymers
analyzed. Within the framework of the present approach, this value must correspond
to some average polymerization degree N0, at which the polymer solubility in various
solvents must be estimated experimentally. Because the polymerization degree N of
real polymers may differ from N0, condition (XII.37) will finally change to:
1/ 6

N
q 1.374 0
N

(XII.38)

Hence, even if the polymer has the globular supermolecular structure, the
condition of its solubility depends on the polymerization degree. For the solubility
condition to be fulfilled in the form of expression (XII.20), the whole polymer
solvent system must be reduced by molecular weights, i.e. the solubility criterion
must be presented in the form:

* 1.374,

(XII.39)

where * = (N/N0)1/6 ( is of the same meaning as before, i.e. = p2 s2 ).


To illustrate the influence of the polymerization degree on solubility of
polymers, Figure 95 indicates the curve of dependence of */ on (N/N0)1/6. If the real
polymerization degree N < N0, the solubility increases, and vice versa. In some cases,
taking into account the polymer molecular mass improves the predictive power of the
criterion.

524

Figure 95. Dependence of */ on (N/N0)1/6.

Up to now, analysis of the solubility criterion was performed for amorphous


polymers with the globular supermolecular structure. However, it is common
knowledge that the factor as orientation affects solubility of polymers. For example,
polyvinyl alcohol is well soluble in water, but fibers formed from this polymer are
quite stable in water. That is why the solubility criterion in the form of expression
(XII.21) may not be applied to prediction of the solubility of polymeric substances
with stretched macromolecules. In this connection, let us discuss the problem of
changing of criterion (XII.21) at the transition to the fibril structure. Let us confine
ourselves to the case when fibrils touch by the generatrix. In contrast with connecting
globules, connection of the fibril with other fibrils is performed via connecting
cylinders (oriented polymer), which are considered above in detail. Figure 96
indicates the cross-section of fibrils and the scheme of application of forces in it.

Figure 96. Cross-section of fibrils and schematic representation of surface forces applied.

According to the Laplace law, the force ff clamping down the given fibril in
relation to other ones equals
f f = 2 Rp Lf

s + p s-p
R

(XII.40)

where 2Rp is the distance between connecting cylinders of fibrils (see Figure 96); Lf is
the length of the connecting cylinder generatrix, equal to the fibril length. Taking into
account that Rp = R/2, the expression for ff will be presented in the following form:
ff = Lf(s + p s-p).

(XII.41)

525

Here s-p is the coefficient of the interfacial tension of the polymersolvent type.
The force fc applied to the connecting cylinder from the side of the solvent and
striving to detach the fibril equals

s + p s-p
= Lf s + p s - p ,
f c = 2rf Lf sin
R
2

(XII.42)

where rf is the radius of the connecting cylinder; = 60 (see Figure 96).


The first condition of detachment, according to which fc ff > 0, has the same
form as in the case of the globular structure (see above)

s + p > s-p .
This condition is always fulfilled.
The second condition of detachment is associated with the work of
detachment:
p
A = ( f c f f ) max
r,

(XII.43)

p
performed by forces of the surface tension of the solvent; max
is the deformation at
break of the connecting cylinder. This work must exceed the energy of disruption of
intermolecular bonds.
The work of the solvent consumed for disruption of connective cylinders and
related to the volume of two connecting cylinders should be presented in the
following form:

( s* )2 = A (2Vc ) .

(XII.44)

Here Vc is the volume of the connecting cylinder, equal to Vc = r2Lf.


Using expressions (XII.41) and (XII.42), we obtain

( s* )2 = s2 f s + p s-p ,

(XII.45)

where

s
( s* )2 = 3 s max

4rs ;

r 2( 1)
p
f = max s
.
s
r
3
max f
Consequently, the second condition may be presented in the following form:

( )2

s2 s*
or

f ,

(XII.46)

526

where and are of the same meaning as in the case of the globular supermolecular
structure.
This condition differs from condition (XII.24) by the value of coefficient
only. The relation f/gl, where gl corresponds to the polymer with the globular
supermolecular structure, may be presented in the form:

f 2( 1) rgl
=

,
rf
gl
3
where rgl is the radius of the connecting globule; rf is the radius of the connecting
cylinder.
Because according to estimations made in ref. [94] rgl/rf ~ (1/3)0.5,
consequently,

f
0.82.
gl
As a consequence, the condition of polymer solubility with the fibril structure
is stricter than for the globular structure. The same criterion is valuable, but with
lower coefficient 2f = 1.125:

1.125 = 1.125 2 1 + a .

(XII.47)

Hence, the polymer with the fibrillar supermolecular structure may not
dissolve even under the condition of polymer solubility with the globular structure.

Chapter XIII. Surface properties of organic liquids and


polymers

It is important to know the surface tension value and be able to predict it on


the basis of the chemical structure of the substance is important, because many
individual properties of liquids and solids, as well as miscibility and solubility of them
in each other, depend on the surface tension of liquids and surface energy of solids.
Among existing methods of calculation of surface tension, the most wide-spread is the
additive scheme based on summing up of parachors which characterize the
contribution of separate atoms to surface tension. The calculation is performed by the
formula
Pi

P
= = i
V
V

(XIII.1)

where P is the total parachor of the molecule (or the repeat unit of the polymer); Pi are
parachors typical of every atom.
Values of atomic parachors Pi, as well as constants which characterize
contribution of various types of bonds (double, triple) and cycles (three-component,
four-component, etc.) to the general parachor of the substance, are indicated in Table
45.
Table 45
Values of atomic parachors Pi and parachors for a series of constants Pj
Constants
Pj
Atom
Pi
C
4.8
Double bond
23.2
H

17.1

20.0

O2*

60.0

N
S
F

12.5
48.2
27.5

Cl
Br

54.3
68.0

I
91.0
* For two oxygen atoms participating in the ester

Triple bond

46.4

Three-component cycle

16.7

Four-component cycle

11.6

Five-component cycle

8.5

Six-component cycle

6.1

C O

group.

Formula (XIII.1) enables the value to be calculated with high accuracy for
organic liquids and polymers. However, the physical meaning of values Pi
participating in correlation (XIII.1) is not clear yet. This makes it difficult to analyze
the influence of the weak Van-der-Waals interaction and strong polar groups on
formation of some surface properties of organic liquids and polymers. This problem is
quite important, because analysis of such an influence on the quantitative level allows

528

regulation of the chemical structure of substances to ensure that they have the
necessary surface properties.

XIII.1. Surface tension of organic liquids


Let us consider the approach to analysis of influence of the chemical structure
on surface tension, suggested in ref. [34]. This analysis was based on the following
ideas. It is common knowledge that molecules of a liquid located on the surface
participate in an intermolecular interaction, different from that in the volume. If the
volumetric problem is considered (Figure 97), then at hexagonal packing of spherical
molecules present in the volume the coordination number equals 12, and on the
surface 9, i.e. their ratio is 4/3. As the free energy decreases with the coordination
number increase, molecules strive to transit from the surface layer into the volume.
That is why the formation of a unit of new surface require consumption of energy.

Figure 97. Schematic representation of arrangement and interaction of molecules in the vertical (a)
and horizontal (b) planes.

Let us calculate first the number of molecules n located on the surface unit.
It may be easily displayed [34] that at hexagonal packing (see Figure 97) the
value n per 1 cm2 of surface will be
n=

1016 0.2887
r2

(XIII.2)

where r is the molecule radius, expressed in .


Therewith,
Vi

r = i
3
4

1/ 3
1/ 3

= 0.6204 Vi

(XIII.3)

where Vi are Van-der-Waals volumes of atoms participating in the molecule


composition.
With regard to correlations (XIII.2) and (XIII.3), we obtain that

529

n=

1016

Vi

2/3

0.75 .

(XIII.3)

Let us calculate the cohesive energy e* accounted for one molecule.


According to the approach for estimating the solubility parameter (see above), the

Ei*

cohesive energy accounted for one mole of the substance equals

k , where

Ei* are molar constants displaying the meaning of the energy of intermolecular
interaction for each type of atoms and separate polar groups participating in the
molecule or the repeat unit of the polymer (values of Ei* are shown in Table 43); k
is the coefficient of molecular packing in the volume of the substance considered.
Then

Ei* Ei*

e* = i
kN A

= i
0.166 .
k 1016

In formula (XIII.5), the value e* is expressed in ergs, and

(XIII.5)

Ei*

in Joules.

The cohesive energy *, which would be possessed by all molecules in the


volume located in the surface layer of 1 cm2 square equals ne* and, according to
correlations (XIII.4) and (XIII.5), equals

* = 0.125

Ei*
i

k Vi*

2/3

(XIII.6)

The value of * significantly exceeds the value of surface tension, because the
formation of a specific new surface requires overcoming of not the whole cohesive
energy, but just a part of it, because in the surface layer the molecule possesses the
coordination number different from zero. Figure 97b indicates that approximately 1/4
part of the cohesive energy is consumed for the molecule yield from the volume to the
surface, because in this case the coordination number changes from 12 to 9.
Moreover, it should be taken into account that packing of molecules in the surface
layer is not perfect, as it is observed from Figure 97b. Let us designate the true
coefficient of molecules packing in the surface layer as ksrf. For future calculations, let
us determine the coefficient of packing ksrf,p in the surface layer at the ideal hexagonal
packing of spheres. To determine the maximum possible occupied volume Vocc, let us
multiply the number of molecules n in the specific surface layer by the volume of a
single molecule:

530

Vocc =

1016
r

4
0.2887 r 3 = 1.20861016 3
3

(r is expressed in ngstrms). The total volume of the surface layer Vtot = 21016 3,
and then ksrp,p = Vocc/Vtot = 0.604.
Hence, the coefficient of packing ksrf,p at the ideal packing of spheres in the
surface layer equals 0.604, whereas the appropriate coefficient kvol,p in the volume
equals 0.740.
Consequently, the part of molecules present on the surface equal ksrf/0.604,
where ksrf is the true coefficient of molecular packing in the surface layer. Then for
the value *, we obtain

* = 0.125

ksrf Ei*
i

0.604 Vi

2/3

(XIII.7)

Let us now take into account the part of the cohesive energy which must be overcome
for the molecule to transit to the surface from the volume. This part stipulated in
this transition by the coordination number change, will be determined from the
correlation:

12

k
k
9 srf
0.740
0.604 = 1 3 ksrf 1.225 .
k
4 k
12
0.740

(XIII.8)

Then, for the surface tension of liquids, we obtain:

s = * =

0.207ksrf Ei*
i

k 0.74 Vi*

0.919ksrf
1
.
k

2/3

(XIII.9)

Basing on correlation (XIII.9), ref. [34] indicates the calculation of values


ksrf/k for a great number of organic liquids of various nature. Initial data and results of
this calculation for representatives of various classes of liquids are indicated in Table
46. Clearly, for all organic liquids the value ksrf/k < 1. Therewith, it was found that for
many organic liquids, especially for unitypical ones, for example, for hydrocarbons,
alcohols, etc., the value ksrf/k is approximately constant. The calculations performed
indicate [34] that for hydrocarbons, perfluorocompounds, halogen-containing
compounds, aldehydes, ketones and nitrocompounds the average value (ksrf/k)avg =
0.851. For alcohols, amides, and acids, (ksrf/k)avg = 0.954, and for nitriles (ksrf/k)avg =
0.910. The values of the coefficient of molecular packing in the surface layer are also
displayed in Table 46.

531

Because for some series of organic liquids the value ksrf/k plays the role of a
constant, then correlation (XIII.9) may be reduced to the form:

s = A

Ei*
i

Vi*

2/3

(XIII.10)

where
A = 0.207(ksrf/k)(1 0.919ksrf/k).
Taking into account the above-displayed values of ksrf/k, the following
expression for calculation of the surface tension of organic liquids may be presented:

s = Aj

Ei*
i

Vi*

2/3

(XIII.11)

For hydrocarbons, perfluorocompounds, halogen-containing compounds,


esters, aldehydes, ketones and nitrocompounds (group I), A1 = 0.0287; for alcohols,
amides, and acids (group II), A2 = 0.0181; for nitriles (group III), A3 = 0.0229. Recall
that Ei* is expressed in Joules, Vi 3, and then, with indicated coefficients Aj, the
value s is expressed in dyn/cm.
Table 46
Initial data and results of calculation of the surface tension of a series of organic liquids
2/3
, dyn/cm
Ei*, Vi* ,
k
ksrf
Compound
i

Exper.
Calc.
i

J/mol
3

1
2
3
4
5
6
7
Hydrocarbons
n-Pentane
21.30
0.510
0.453
16.03
17.960
13,903
Isopentane
12,180
21.40
0.510
0.447
15.00
17.330
n-Hexane
23.70
0.527
0.462
18.41
19.895
16,603
Cyclohexane
16,206
21.90
0.567
0.443
25.12
23.760
n-Heptane
19,303
25.90
0.541
0.469
20.21
21.604
n-Octane
22,003
28.20
0.549
0.472
21.75
22.700
Isooctane
20,281
28.20
0.545
0.479
18.85
21.710
n-Nonane
30.30
0.558
0.479
22.91
23.940
24,704
n-Decane
27,404
32.30
0.571
0.489
23.92
25.900
n-Tetradecane
38,209
39.90
0.589
0.506
26.96
28.770
Benzene
19,186
19.80
0.597
0.501
28.78
29.760
Toluene
20,687
22.30
0.592
0.489
28.53
28.680
n-Xylene
23,391
24.70
0.610
0.960
30.03
30.650
m-Xylene
23,391
24.70
0.600
0.515
26.63
29.740
o-Xylene
23,391
24.70
0.595
0.499
28.31
29.250
Ethylbenzene
23,391
24.60
0.598
0.514
29.04
29.599
Isopropylbenzene
24,369
26.80
0.596
0.490
28.20
28.790

532

1
n-Perfluoropentane
n-Perfluoroheptane
n-Perfluorooctane
Dichloromethane
Chloroform
Tetrachlorocarbon
Ethyl chloride
1,1-Dichloroethane
1,2-Dichloroethane
1,1,1-Trichloroethane
Tetrachloroethane
Isobutyl chloride
cis-Dichloroethylene
trans-Dichloroethylene
Trichloroethylene
Tetrachloroethylene
Chlorobenzene
Bromobenzene
Iodobenzene
Bromoform
Ethyl bromide
1,2-Dibromoethane
Ethyliodide
Glycol
Glycerol
Cyclohexanol
Methanol
Ethanol
n-Propanol
n-Butanol
Isobutyl alcohol
n-Amyl alcohol
n-Hexanol
n-Octanol
Acetic
Butyric
Isovaleric
n-Valeric
Methyl formiate
Ethyl formiate
Methyl acetate
Ethyl acetate
n-Propyl acetate
Isopropyl acetate
Ethyl propyonate
n-Butyl acetate
Amyl acetate
Ethyl isovaleate
Diethyl carbonate
Benzaldehyde
Acetaldehyde

3
4
Perfluorocompounds
12,724
29.50
0.574
15,228
32.80
0.551
17,732
35.90
0.571
Halogen-containing compounds
14,408
15.14
0.554
13,275
17.60
0.555
12,147
19.87
0.553
11,453
15.52
0.527
17,109
17.94
0.544
17,109
17.94
0.579
15,976
20.24
0.555
28,416
22.40
0.607
15,132
20.96
0.550
15,357
17.20
0.571
15,357
17.20
0.563
18,313
19.57
0.582
21,268
21.70
0.605
23,642
21.80
0.605
27,011
23.00
0.634
31,688
24.60
0.661
23,379
20.73
0.650
14,822
16.70
0.542
23,843
20.05
0.628
19,491
18.41
0.590
Alcohols
39,840
15.50
0.659
59,557
19.41
0.705
33,443
22.96
0.658
20,118
11.02
0.544
22,819
14.10
0.565
25,511
17.20
0.563
28,219
19.77
0.578
26,497
19.77
0.573
30,919
22.30
0.587
33,624
24.70
0.590
39,024
21.90
0.597
Acids
23,015
14.60
0.589
28,420
20.00
0.587
29,398
22.80
0.597
31,120
22.50
0.519
Esters
13,376
14.90
0.567
16,076
17.80
0.565
16,076
17.80
0.565
18,781
20.40
0.565
21,481
22.87
0.573
19,758
22.90
0.567
21,481
22.87
0.573
24,181
25.20
0.577
26,882
27.40
0.584
25,159
27.00
0.570
22,075
24.20
0.593
Aldehydes and ketones
27,672
22.50
0.636
12,782
13.20
0.509

0.523
0.478
0.493

9.87
12.60
13.60

16.904
17.330
18.680

0.458
0.424
0.350
0.423
0.475
0.455
0.440
0.518
0.458
0.468
0.484
0.472
0.481
0.479
0.522
0.558
0.609
0.470
0.497
0.535

28.12
27.16
26.75
22.18
24.75
32.23
25.77
36.04
21.99
28.00
25.00
29.50
32.26
33.19
36.34
37.65
31.68
24.15
38.91
28.10

24.800
23.600
20.900
20.500
23.800
28.010
23.700
34.230
22.160
26.090
25.195
28.600
30.720
32.600
36.790
41.753
38.000
22.520
36.800
30.400

0.625
0.660
0.576
0.539
0.558
0.547
0.562
0.564
0.559
0.564
0.557

46.49
62.29
34.37
22.55
22.32
23.70
24.57
22.98
25.68
24.48
27.53

51.900
71.100
44.600
33.100
23.600
23.930
29.200
28.300
30.820
31.460
32.840

0.566
0.556
0.563
0.555

27.42
26.96
25.31
27.35

30.450
30.700
32.640
31.680

0.490
0.495
0.490
0.499
0.505
0.498
0.505
0.506
0.511
0.504
0.502

24.62
23.84
24.49
23.75
24.28
22.35
24.27
25.20
25.88
23.77
26.44

25.696
25.510
25.510
25.650
26.770
25.350
26.770
27.430
28.470
26.350
28.550

0.513
0.468

39.70
23.32

37.890
24.900

533

1
Acetone
Methylethylketone
Diethylketone
Cyclohexanon

2
15,483
18,183
20,883
23,186

Diethyl ether
1,4-Dioxane

11,796
25,564

Aniline
Pyridine
Formamide

30,008
17,514
31,739

Acetonitrile
Propyonitrile
n-Valeonitrile
Capronitrile
Methacrylonitrile

17,021
19,725
25,126
27,826
20,674

Nitromethane
Nitroethane
2-Nitropropane
Nitrobenzene

15,913
18,613
19,592
31,199

3
16.30
18.98
21.50
22.10
Ethers
20.20
19.70
Amines, amides
21.50
18.50
12.30
Nitriles
13.50
16.40
21.70
24.00
18.50
Nitrocompounds
12.86
15.88
18.70
22.30

4
0.534
0.556
0.570
0.575

5
0.478
0.491
0.503
0.450

6
24.60
25.18
25.18
35.62

7
21.299
24.690
26.660
27.850

0.521
0.614

0.411
0.540

16.49
33.70

18.270
35.590

0.658
0.595
0.640

0.541
0.403
0.585

43.31
37.25
58.35

43.230
29.700
43.700

0.565
0.565
0.589
0.592
0.574

0.514
0.546
0.532
0.533
0.529

29.10
27.25
27.44
27.87
24.40

26.210
26.390
30.600
31.080
27.990

0.518
0.534
0.547
0.615

0.428
0.433
0.471
0.506

36.98
32.06
29.12
43.35

35.500
33.600
32.900
39.700

Calculations performed by formula (XIII.11) indicate quite good agreement of


the calculated and experimental values of , deviations being approximately the same
as in calculations with the help of parachors. Correlation (XIII.11) enables the
contribution of separate groups and specific intermolecular interaction to the surface
tension to be estimated. For example, usually, of interest is the contribution of
hydrogen bonds made in formation of the surface properties of organic liquids. Let us
estimate such contribution on the example of alcohols and acids. For this purpose, let
us transform correlation (XIII.11) to the following form:

s =

where

Eih*

*
*
Eih
+ Eid
,
2/3

i
i

Vi*

0.0181

is the contribution of hydrogen bonds to cohesive energy;

Eid* = Ei* Eih* is the contribution of weak dispersion interaction to


i

cohesive energy.
For ethanol,

Ei*

*
*
*
= 2 EC
+ 6 EH
+
+ EO

5,460 cal/mol = 22,820 J/mol, and

Eih*

= 2550.7 + 647.7 + 142.6 + 3929 =

Eih*

= 16,423 J/mol. Then a part of the surface

tension stipulated by hydrogen bonding will be determined as

534

sh =

0.0181

Eih*

= 20.9 dyn/cm.

i
Vi*

The contribution associated with weak dispersion interaction gives


2/3

sq =

0.0181

Vi*

2/3

Eid*

= 8.1 dyn/cm.

Total surface tension s = sh + sd = 29 dyn/cm, and the contribution of each


component equal 72.0 and 28.0%, respectively.
Analogously, for acetic acid sh = 20.5, and sd = 8.2 dyn/cm; the contribution
of each component equals 71.0 and 29.0%, respectively.
This analysis may be performed not only for hydrogen bonds, but also for
specific interactions of other types. It should be noted that according to the Fowkes
method [143] significantly different contributions of hydrogen bonding to the total
value of surface tension are obtained. For example, for formamide sh = 19 and sd =
58.2 dyn/cm [143]. The part of the surface tension stipulated by hydrogen bonding is
33%. According to calculations performed with the help of data from Table 46, for

Eih*

formamide

= 16,423 and

Eih* Ei*
i

Ei*

= 31,739 J/mol. Then this part is

0.52 or 52%, which is significantly higher than that obtained by

Fowkes.
Besides estimation of the surface tension of organic liquids, of interest is the
relationship of the value with the specific cohesive energy of the liquid 2 or
solubility parameter . This problem has been discussed many times in the literature;
there are empirical correlations, which allow setting of this connection. For example,
the correlation is known [122]

= 4.1(/V1/3)0.43,

(XIII.12)

where V is the molar volume of the liquid. The physical meaning of this correlation is
not clear and, moreover, it does not describe the properties of all organic liquids. For
example, calculation of for alcohols and organic acids by this correlation leads to
significantly reduced values.
With the help of correlation (XIII.11), the dependence connecting the
solubility parameter value with the surface tension may be deduced. For this purpose,
let us use correlation (XII.3), according to which solubility parameter is calculated.
First, correlation (XIII.11) is transformed by multiplying the numerator and
1/ 3

denominator by Vi

. Then, the following expression is obtained

535

s = A i

1/ 3

Vi
i

V
i

Ei*

(XIII.13)

Substituting correlation (XII.3) into formula (XIII.13), we obtain


1/ 3

s = A0.6023 Vi

2 .

(XIII.14)

In correlation (XIII.14), values of Vi are expressed in ngstrms, 2 in


J/cm . Taking into account that A displays several different values for various
different groups of organic liquids, we obtain the correlation:
3

1/ 3

s = Bj Vi

2 ,

(XIII.15)

where Bj = 0.6023Aj. For liquids from the group I, B1 = 0.0172, from the group II
B2 = 0.0109, from the group III B3 = 0.0138. To determine the value by correlation
(XIII.15), both the calculated and experimental values of may be used. The results
of calculation of performed with application of experimental values of are shown
in Table 46. Clearly, the experimental values exper and calculated values calc coincide
quite well.
The opposite task (which is more important) may also be solved, i.e. values of
the solubility parameters may be calculated from the experimental value of surface
tension. For this purpose, let us transform correlation (XIII.15) to the following form
(with regard to the fact that N A Vi = Vk ):
i

1/ 2

(0.6023)1 / 6 s

B1j / 2 k 1 / 6 V 1 / 3

(XIII.16)

Clearly, correlation (XIII.16) by its form is similar to empirical formula


(XIII.12). However, the multiplicand before the value (s/V1/3) is not constant but
depends upon the chemical structure of organic liquids, because the coefficient of
packing k depends on it. In the first approximation, it may be assumed that for groups
of organic compounds considered above the values of k vary in a narrow range, and
their average value may be applied. The calculations performed indicate that for the
group I of organic liquids kavg,1 = 0.580, for the group II kavg,2 = 0.601, and for the
group III kavg,3 = 0.586. Then basing on formula (XIII.16), we obtain:

1/ 2
* s
,
= C j
1/ 3

(XIII.17)

536

where C1* = 7.67; C2* = 9.58; C3* = 8.56.

XIII.2. Surface tension of polymers


Let us now turn to calculations of the surface tension of solid polymers. At
first glance, it seems that the surface tension of polymers may be calculated directly
from formula (XIII.11), where

Ei*

and

Vi

are determined per repeat unit of

the polymer. However, some difficulties appear in this case associated with both chain
structure of the polymer and somewhat ambiguity of interpretation of the repeat unit.
Because the value Vi participates in this correlation in the degree 2/3, and the
value

i
*
Ei in the first degree, then the more the number of chains n is considered

as a kinetic unit (the length of the whole chain, the segment length), the higher value
of will be obtained by formula (XIII.11).
The second reason is clear from the example of polyethylene and other
polymers. If for polyethylene the repeat unit is assumed in the form (CH2CH2),
then from correlation (XIII.11) we obtain = 14.7 dyn/cm. But if the repeat unit is
assumed in the form (CH2), then = 11.65 dyn/cm.
The same result is also obtained in the case of polyamides. For example, for
polyamide 6
(CH2)5

C NH
O

and polyamide 6,6


(CH2)4

C NH

(CH2)6

NH C

O
O
n
when formula (XIII.11) is applied, we obtain = 45.9 dyn/cm (polyamide 6) and =
58.0 dyn/cm (polyamide 6,6). From the point of view of the gross formula, the
chemical structure of these polyamides is identical (polyamide 6,6 contains twice the
number of the same elements, which are contained in polyamide 6). Experimental
values of for these polyamides are approximately equal and vary in the range from
40 to 47 dyn/cm.
The above-discussed problems are analyzed in detail in ref. [33].
Consequently, the following correlation is obtained:

p =Cj

Ei*
i

Vi

2/3

(XIII.18)

m1 / 3

where m is the number of atoms in the repeat unit of polymer. For non-polar polymers
of the group I (hydrocarbons, perfluoropolymers, polyethers), C1n-p = 0.1277; for

537

polar polymers of the group II (polyesters, polymers containing a nitrogroup, etc.),


C1p = 0.0751; for polymers containing alcohol, acidic and amide groups (all of them
are polar and indicate hydrogen bonding), C2 = 0.0476; for polymers containing
nitrile groups (all of them are polar), C3 = 0.0600. If a polymer contains aromatic
cycles (for example, polystyrene, polyphenylene, etc.), the value C4 = 0.1014.
Calculations performed due to formula (XIII.18) display (Table 47) that in the
most of cases, good coincidence of experimental and calculated values of the surface
tension p is observed. The greatest deviations are observed for polymers which
crystallize easily; to them polyethylene oxide and polyformaldehyde, for which the
coefficient of molecular packing exceeds significantly the average value.
Table 47
Initial data and results of calculation of the surface tension of polymers
p, by eqs.
p, by eq.
p, exper.
Ei*

(XIII.28)
and

V
i [133, 214] (XIII.18)
Polymer
i
(XIII.29)
i
dyn/cm
Polyethylene
5,401
34.2
31.0-35.7
36.0
35.6
Polypropylene
14,885
51.3
29.0-29.6
38.8
39.6
Polyisobutylene
10,801
68.6
27.0-33.6
28.8
34.8
Rubber SCB
9,054
64.3
32.0
33.5
32.9
Rubber SCI
10,030
81.5
31.0
29.0
28.6
Polyvinyl fluoride
12,087
39.3
28.0-36.7
43.2
44.3
Polyvinyl chloride
11,055
49.0
39.0-41.5
34.2
36.3
Polyvinyliden fluoride
5,205
44.6
23.0-32.7
29.1
29.0
Polytrifluoroethylene
11,891
49.7
22.0-23.9
36.3
38.9
Polychlorotrifluoroethylene
10,761
65.7
31.0
27.4
30.7
Polytetrafluoroethylene
5,008
55.0
18.5-19.0
24.5
18.8
Polystyrene
22,990
109.7
33.0-43.0
40.5
38.0
Polyvinyl alcohol
23,675
41.6
37.0
46.5
61.0
Polyacrylonitrile
19,324
54.0
44.0
42.6
50.0
Polymethylacrylate
18,379
79.5
41.0
32.6
33.8
Polyamide II
51,562
201.2
31.0-33.0
34.9
31.4
Polyamide 6
38,059
116.0
40.0-47.0
45.0
39.5
Polyamide 6,6
76.118
231.5
39.3-46.0
45.2
35.1
Polyamide 10,10
97,723
368.3
28.5-32.0
36.1
29.1
Polyethyleneterephthalate
43,547
169.9
41.0-47.0
38.1
35.2
Polyethylacrylate
21,080
96.7
35.0
30.4
31.6
Poly(methyl methacrylate)
21,080
96.7
39.0-40.2
30.4
31.6
Poly(ethyl methacrylate)
23,780
113.8
33.0
29.1
30.0
Poly(vinyl acetate)
18,379
79.6
36.0
32.6
33.8
Polyformaldehyde
3,297
27.2
36.0
29.1
36.0
Poly(ethylene oxide)
5,997
43.8
43.0
32.1
43.0
Poly(propylene oxide)
8,698
60.9
32.0
33.2
31.9

For polymers, it is desirable to calculate the surface energy p with the help of
solubility parameter . Clearly, from equation (XII.3) we obtain:

Ei* = 2 N A Vi .
i

Substituting formula (XIII.19) into equation (XIII.18), we obtain

(XIII.19)

538

1/ 2

Vi 2 N
A

p =Cj
m1 / 2

(XIII.20)

or

p = D j

Vi

(XIII.21)

where Dj = CjNA (Dj is the parameter which, similar to Cj, depends on affiliation of
the polymer to the given group). With regard to values of Cj shown above, it may be
stated that for non-polar polymers (hydrocarbons, polyethers, etc.), D1n-p = 0.0769.
For polar polymers containing ester and other polar groups, D1p = 0.0452. For
polymers containing hydroxyl, acidic and amide groups (strong hydrogen bonding),
D2 = 0.0287; for polymers containing nitrile groups, D3 = 0.0361. For polymers
containing aromatic cycles only (for example, polystyrene, polyphenylene, etc.), D4 =
0.061. Therewith, if the Van-der-Waals volume is expressed in 3, and the specific
cohesive energy 2 in J/cm3, then the surface tension calculated by formula (XIII.21)
is determined in dyn/cm.
The drawback of the calculation scheme described above concludes in the fact
that transition from one sequence of liquids or from one group of polymers to another
requires application of the own coefficient Aj or Cj in equations (XIII.11) and
(XIII.18). Difficulties in this case are associated with the fact that molecules of the
liquid or repeat units of polymers will possess larger volumes, and the relative
fraction of specific groups, by which they are separated according to the current
feature, will be small. For example, nitrile groups may be contained not only in
polyacrylonitrile, where their relative contribution to the Van-der-Waals volume and
the cohesive energy will be high, but also in other polymeric systems, where their
contribution may be rather low. Moreover, difficulties appear in the calculation of the
surface energy of copolymers, units of which belong to different classes of
homopolymers.
In this connection, ref. [37] indicates development of the calculation scheme
for estimating the surface tension of organic liquids and polymers, parameters of
which depend on the chemical structure of the organic liquid or polymer only, but not
on their belonging to any class.
In the work cited, the value of kp is expressed via the correlation:

* =

Ei*
i

Vi

2/3

(XIII.22)

Here * is the specific surface cohesive energy, i.e. the cohesive energy per
specific surface. Note that this value differs from the usual density of the cohesive
energy determined by correlation (XII.3).
The following dependence connects the coefficient of molecular packing of
the liquid on its surface kp and the specific surface cohesive energy * [37]:

539

Ksrf = 0.4112 + 0.3012 lg*.

(XIII.23)

After substituting expression (XIII.9) into formula (XIII.23), we obtain

s =*

0.0461 lg * 0.063 0.2768 lg * 0.378


1
.
k
k

(XIII.24)

Substituting expression (II.5) into formula (XIII.24), finally we get:

(0.2768 lg * 0.378) M
1
.

N
V

i
A

s =

*
* (0.0461 lg 0.063) M

N A Vi
i

(XIII.25)

Correlation (XIII.25) enables the surface tension of organic liquids to be


estimated with accuracy valuable for practical calculations independently of the class,
to which the current liquid belongs. Results of calculation indicate [37] that the
calculated and experimental data are in satisfactory agreement in most cases (see
Table 46).
Let us now turn to estimation of the surface energy of polymers. The initial
correlation for deducing equation (XIII.18) has the form [37]:
k
k
*
p = 0.1532 srf
1 0.919 srf ,
k
k m1 / 3

(XIII.26)

where is the constant dependent on the polymer belonging to the class of polar or
non-polar polymers only.
Let us assume that the dependence of the coefficient of molar packing in the
surface layer ksrf on * is the same as for liquids. As often mentioned above, the
coefficient of molecular packing of the polymer in the volume is almost independent
of the chemical structure of polymers, and its average value kavg = 0.681 for solid
polymeric substances.
Using the dependence ksrf on * in the form of correlation (XIII.23) and kavg =
0.681, basing on correlation (XIII.23) we obtain:

p =

*
1/ 3

(0.067761 lg * 0.0925)(1.5549 0.40671 lg * ).

(XIII.27)

The analysis performed indicates [37] that values of are indeed


approximately equal; for polar polymers, avg = 2.097, and for non-polar polymers
avg = 3.055.
Then, the formula for calculation of the surface energy of polar polymers
changes to:

pp =

*
1/ 3

(0.1421 lg * 0.194)(1.5549 0.4067 lg * ),

(XIII.28)

540

and for non-polar polymers,

pn =

*
1/ 3

(0.2070 lg * 0.2826)(1.5549 0.4067 lg * ).

(XIII.29)

The surface energy values calculated from formulae (XIII.28) and (XIII.29)
are displayed in Table 47. Clearly, in most cases, good correspondence between the
calculated and experimental surface energy values is observed. It should be noted that
experimental determination of the surface energy for solids, including polymers, is
rather difficult, although there are various modifications of the method for estimating
p. That is why greatly different data on the surface energy for the same polymer may
be observed in the literature.
Correlations (XIII.28) and (XIII.29) deduced in ref. [37] enable the surface
energy of polymers of any chemical structure to be estimated regardless of their
belonging to any class of chemical compounds. This removes the difficulties typical
of calculation of the surface tension of organic liquids, mentioned above.
In conclusion, let us discuss one more connection between the surface energy
of organic liquids and polymers and the solubility parameter , the value of which is
determined from relation (XII.3).
Taking into account that * =

Ei*
i

Vi

2/3

and substituting this expression

into equation (XII.3), we obtain that


1/ 3

= N A Vi

(XIII.30)

Furthermore, substituting expression (XIII.30) into formula (XIII.29) and


performing necessary transformations, we get:

V
0
.
073
i

1 / 3 0.09221lg 0.01537 lg

s = N A 2 Vi

. (XIII.31)

0.554 lg 0.0923 lg Vi 0.439

Correlation (XIII.31) links value of the surface tension of liquids s with the
solubility parameter . According to the correlation mentioned, the value may be
obtained if experimental values of the surface tension of liquids are known

541

(estimation is performed by the method of inspection). This is of importance because


the surface tension of liquids can be measured quite easily, and the solubility
parameter associated with the latent heat of evaporation of the liquid is more difficult
to calculate. It is especially difficult for polymers, because their solubility parameter
may be determined by indirect methods only by measurements of swelling in
different solvents, viscosity of solutions, etc. It should be noted that the possibility of
estimating the surface energy of polymers experimentally is also important, because
their surface energy is related by simple correlations with cohesive energy and
sublimation energy.
Various situations may appear when calculating the surface energy of
copolymers or homogeneous mixtures. If components of a copolymer relate to the
same group of substances (according to the above-mentioned classification), then, as
the value Cj is identical for them, correlation (XIII.18) applied to copolymers obtains
the form:

p =Cj

1 Ei* + 2 Ei* + ... +

i
1
i
2


1 Vi + 2 Vi + ... + n Vi


i
1
i
2
i
n

+ n Ei*

i
n

,
(1m1 + 2 m2 + ... + n mn )1 / 3

2/3

(XIII.32)

where 1, 2, , n are molar parts of the components 1, 2, , n; Ei* ,

i
1

Ei* , , Ei* are values of the cohesive energy for the components 1,

i
n
i
2

2, , n; Vi , Vi , , Vi are their Van-der-Waals volumes;

i
n
i
1 i
2
m1, m2, , mn are numbers of atoms in repeat units of the components 1, 2, , n. In
the reduced form, correlation (XIII.32) gives

p = Cj

k =n

k =1

k Ei*

.
(XIII.33)
1/ 3
k = n
k =n

k Vi k mk

k =1 i
k k =1

If it is desirable to express the surface energy of a copolymer via the surface


energies of homopolymers p,1, p,2, , p,n obtained from components 1, 2, , n, then
equation (XIII.32) is reduced to the following form:
2/3

542

2/3

p = Cj

1 p,1 Vi

i
1

2/3

Vi m12/ 3 + ... +
i
2

m11 / 3 + 2 p,2


1 Vi + 2 Vi + ... + n Vi

i
1
i
2
i
n

2/3

, (XIII.34)

2/3

+ n p, n Vi m1n/ 3

i
n

(1m1 + 2 m2 + ... + n mn )1 / 3

In the reduced form, correlation (XIII.34) looks as follows:


2/3

k =n

p = Cj

k p, k Vi
k =1
i
k

k = n

k Vi

k =1 i
k

2/3

m1k / 3
1/ 3

(XIII.35)

k =n

k mk

k =1

If all components of the copolymer relate to different groups of substances


(they possess different values of Cj), then correlation (XIII.32) for this copolymer (on
the example of a double copolymer) is presented in the form:

p = 1C j , I + 2C j , II

1 Ei* + 2 Ei*

i
1
i
2


1 Vi + 2 Vi

i
1
i
2

(XIII.36)

2/3

(1m1 + 2 m2 )1 / 3

or
2/3

p = 1C j , I + 2C j , II

1 p,1 Vi

i
1

m11 / 3 C j , I +


1 Vi + 2 Vi

i
1
i
2

2/3

(XIII.37)

2/3

+ 2 p,2 Vi

i
2

m12/ 3 C j , II

.
(1m1 + 2 m2 )1 / 3
In the case of a multicomponent copolymer, various situations may appear: a
part of components may belong to the same group of substances, and the rest belongs
to another group. In the general case, we obtain:

543

Cj = 1Cj,1 + 2Cj,2 + + nCj,n,

(XIII.38)

where Cj,1 is the value of Cj for the group of substances, to which component 1
belongs (it may be Cj,I, Cj,II, Cj,III, or Cj,IV); Cj,2, , Cj,n is the same for components
2, , n. Then, with regard to equation (XIII.38), correlation (XIII.33) obtains the
form:

p =

k =1

k =1

k Ei*

k =n

k C j,k

k =n

k = n

k Vi

k =1 i
k

2/3

1/ 3

k =n

(XIII.39)

k mk

k =1

and correlation (XIII.35) is reduced to


2/3

k =n

p =

k =n

k C j,k

k =1

k p, k Vi
k =1
i
k

k = n

k Vi

k =1 i
k

2/3

m1k / 3 C j , k
1/ 3

(XIII.40)

k =n

k mk

k =1

The above-described correlations for calculation of the surface energy of


copolymers are not quite suitable, because affiliation of the polymer to some group
must be taken into account every time. That is why, for computer realization of the
method, it is desirable to express values Cj via the surface energies of copolymer
components. From equation (XIII.18) we obtain that
2/3

p Vi m1 / 3

Cj = i
.
Ei*
i

Then for the copolymer, we obtain:


2/3
2/3

V m1 / 3

m12/ 3
p2 Vi
i
p1
1

i
1
i
2
+ 2
+ ... +
p = 1

*
*
Ei
Ei

i
1
i
2

544

2/3

m1n/ 3 1 Ei* + 2 Ei* +


p, n Vi

i
n
1
i
2
i

Ei*
1 Vi + 2 Vi +

i
i
n
1
i
2

+ ... + n Ei*

i
n


+ ... + n Vi

i
n

(XIII.41)

2/3

(1m1 + 2 m2 + ... + n mn )1 / 3

where all designations are identical to the above-considered formulae. In the reduced
form, correlation (XIII.41) will have the form:
2/3

k =n

pk Vi m1k / 3
k Ei*

k =n
k =1
i
k
i
k

p = k
.
2/3
1/ 3

k = n
k =1
k =n

Ei*
k Vi k mk


i
k
k =1 i
k k =1

(XIII.42)

Let us show an example of calculation of the surface energy for polyvinyl


alcohol copolymer with poly(vinyl acetate):
CH2

CH

CH2

OH

CH
O C CH3
O

Polyvinyl alcohol (PVA)

Poly(vinyl acetate) (PVAc)

For PVA we obtain

Ei*

*
*
*
= 2 EC
+ 4 EH
+ EO
+ Eh* = 2550.7 + 447.7 + 142.6 + 3929 =

5,363.8 cal/mol = 22,420.6 J/mol.


Vi = VC,10 + VC,39 + VH,125 + VH,124 + VO,132 = 13.1 + 12.2 + 32.0 + 4.7 +
i

5.6 = 41.6 3.
The number of atoms m in the repeat unit of PVA equals 7.
Substituting values

Ei* , Vi , and m into equation (XIII.18) and taking into


i

account that the present polymer belongs to group II of polar polymers (C2 = 0.0476),
we obtain that

545

p = 0.0476

22,420.6
(41.6) 2 / 3 71 / 3

= 46.5 dyn/cm.

For PVAc we obtain:

Ei*

*
*
*
= 4 EC
+ 6 EH
+ 2 EO
+ Ed* = 4550.7 + 647.7 + 2142.6 + 1623 =

4,397 cal/mol = 18,380 J/mol.


Vi = VC,10 + VC,39 + VC,48 + VC,13 + 6VH,124 + VO,129 + VO,139 = 13.1 +
i

12.2 + 15.9 + 17.2 + 62.0 + 3.4 + 5.8 = 79.6 3.


The number of atoms m in the repeat unit of PVAc equals 12.
Using equation (XIII.18) and taking into account that PVAc belongs to polar
polymers from group I (Cp1 = 0.0751), we obtain that

p = 0.0751

18,380
(79.6) 2 / 3 121 / 3

= 32.6 dyn/cm.

To calculate p of the copolymer, let us first use correlation (XIII.36); assume


that the molar part of PVA 1 = 0.4, and for PVAc 2 = 0.6. Substituting all
parameters of the copolymer into correlation (XIII.36), we obtain:
p = (0.4 0.0476 + 0.6 0.0751)

0.4 22,420.6 + 0.6 18,380


(0.4 41.6 + 0.6 79.6) 2 / 3 (0.4 7 + 0.6 12)1 / 3

= 37.0 dyn/cm.

Let us now use correlation (XIII.37):

p = (0.4 0.0476 + 0.6 0.0751)

0.4 46.5 41.6 2 / 3 71 / 3 / 0.0476 +


(0.4 41.6 + 0.6 79.6) 2 / 3

0.6 32.6 79.6 2 / 3 121 / 3 / 0.0751


(0.4 7 + 0.6 12)1 / 3

= 37.0 dyn/cm.

Let us now consider the surface energy of a triple copolymer based on acrylonitrile,
butadiene and styrene (ABSplastic):
CH2

CH

CH2CH=CHCH2

CH2

CH

C N

Polyacrylonitrile (PAN)

Polybutadiene (PB)

Polystyrene (PS)

For PAN,

*
*
Ei* = 3 EC
+ 3 EH
+

4623.2 cal/mol = 19,325 J/mol.

E *N + Ed* = 3550.7 + 347.7 + 1205 + 1623 =

546

Vi

= VC,10 + VC,6 + VC,71 + 3VH,124 + VN,154 = 13.1 + 9.0 + 15.9 + 32.0 +

10 = 54 3.
PAN belongs to polymers of group III, because it contains a nitrile group
(C3 = 0.060). The number of atoms, m, in the repeat unit of PAN equals 7. For it,
according to formula (XIII.18), we obtain that

p = 0.060

19,325
54 2 / 3 71 / 3

= 42.4 dyn/cm.

For PB,

*
*
Ei* = 4 EC
+ 6 EH
+

E* = 4550.7 + 647.7 323 = 2,166 cal/mol = 9,054

J/mol.
Vi = 2VC,10 + 2VC,16 + 6VH,124 = 213.1 + 213.1 + 62.0 = 64.4 3.
i

PB belongs to non-polar polymers of the group I (C1n = 0.1277); for it, m = 10.
Application of formula (XIII.18) leads to the following result:

p = 0.1277

9,054
(64.4) 2 / 3 101 / 3

= 33.4 dyn/cm.

For PS,

*
*
Ei* = 8 EC
+ 8 EH
+

E * = 8550.7 + 847.7 + 713 = 5,500 cal/mol = 22,991

J/mol.
Vi = VC,10 + VC,7 + VC,19 + 5VC,18 + 8VH,124 = 13.1 + 8.7 + 8.4 + 512.7 +
i

82.0 = 109.7 3.
PS belongs to group IV of polymers (C4 = 0.1014); for it, m = 16. Substitution
of all values of PS parameters into formula (XIII.18) gives

p = 0.1014

22,991
(109.7) 2 / 3 161 / 3

= 40.4 dyn/cm.

Calculating the surface energy of the copolymer, we assume that the part of
PAN units 1 = 0.3, the part of PB 2 = 0.2, and the part of PS 3 = 0.5. Then it is
obtained from equation (XIII.37) that
p = (0.3 0.06 + 0.2 0.1277 + 0.5 0.1014)

0.3 42.4 54 2 / 3 71 / 3 / 0.06 +


(0.3 54 + 0.2 64.4 + 0.5 109.7) 2 / 3

0.2 33.4 64.4 2 / 3 101 / 3 / 0.1277 + 0.5 40.4 109.7 2 / 3 161 / 3 / 0.1014
(0.3 7 + 0.2 10 + 0.5 16)1 / 3

= 40.9 dyn/cm.

Chapter XIV. Miscibility of polymers

The problem of miscibility of polymers is one of the most important problems


at present. The point is is that new polymeric materials are now created, as a rule, not
by synthesizing new polymers, but by composing mixtures of the known polymers.
Therewith, it is essential to consider not only mixed composites, into which
components of the mixture are injected in comparable amounts, but also the
introduction of microadditions of polymers, their surface modification, etc. There are
also problems appearing in this case, such as microphase separation, facilities to
control the composition and size of microphases, etc. Having no way of discussing
numerous publications which have appeared recently in this field, let us analyze one
of possible ways of predicting the miscibility of polymers and estimating the
composition of microphases.
One of possible ways of looking at the problem of predicting the miscibility of
polymers is the application of criterion (XII.21), which is intended for analyzing the
solubility of polymers. Therewith, if one polymer is introduced in small amounts into
another polymer, the former of them is considered as the polymer and the latter as
the solvent when criterion (XII.21) is applied. In principle, the following variants are
probable.
1. When criterion (XII.21) is used, it turns out that, in all cases, the left part of
the criterion is greater than the right part, i.e. total incompatibility takes place. Under
all cases we mean those when the first polymer is introduced in small quantities into
the second one and, vice versa, the second polymer is introduced in small quantities
into the first one. Then the criterion of the miscibility of polymers obtains the form:
at introduction of the first polymer into the second one:

1 =

p2,1
p2,2

> 1.374 2 1 + a1 = 2 1 ;

(XIV.1)

at introduction of the second polymer into the first one:

2 =

p2,2
p2,1

> 1.374 2 1 + a2 = 2 2 ;

(XIV.2)

where p,1 and p,2 are the solubility parameters of polymers 1 and 2, respectively;
=

4 Vp,1 Vp,2 1 / 3

(Vp,11/ 3 + Vp,21/ 3 )2

(XIV.3)

where Vp,1 and Vp,2 are the molar volumes of polymers 1 and 2, respectively;
a1 = p,1;p,2/p,2;
a2 = p,1;p,2/p,1,
where

(XIV.4)

548

p,1;p,2 = p,1 + p,2 2(p,1p,2)1/2,

(XIV.5)

where p,1 and p,2 are values of the surface energy for polymers 1 and 2, respectively.
Criteria (XIV.1) and (XIV.2) mean that the polymers display total
incompatibility.
2. When small quantities of the first polymer are introduced into the second
one, criterion (XII.21) indicates that they are miscible, i.e.

1 =

p2,1
p2,2

< 1.374 2 1 + a1 = 2 1 .

(XIV.6)

However, when the second polymer is introduced into the first one, it may appear that
miscibility is not observed, i.e.

p2,2

> 1.374 2 1 + a2 = 2 2 .
(XIV.7)

p2,1
This, at first glance, a paradoxic conclusion is indicated, indeed, which will be shown
below. Here, it should also be noted that, usually, the miscibility of two polymers is
judged by the glass transition temperature of their mixture. A single glass transition
temperature is typical of totally miscible polymers and lies between the glass
transition temperatures of original components. A mixture of totally incompatible
polymers display two glass transition temperatures, each of which is identical to the
glass transition temperature of the original component. For partial miscibility, when
both components are present in different quantities in the microphase, two glass
transition temperatures are also indicated but, compared with the glass transition
temperatures of original components, they are shifted towards each other.
For the case under consideration, when the first polymer is combined with the
second one, the dependence of the glass transition temperature on the composition is
of the shape schematically represented in Figure 98, curve 1. This form of the curve
will be explained below on the example of a polymer 1polymer 2 specific system.

2 =

Figure 98. Schematic representation of dependence of the glass transition temperature Tg on the
composition of the mixture : 1 partially miscible polymers; 2 totally miscible
polymers.

549

3. This case is appropriate to the total miscibility of polymers, i.e. the first
polymer is dissolved in the second one, and the second polymer is dissolved in the
first one. The miscibility criterion obtains the form:
at introduction of the first polymer into the second one:

1 =

p2,1
p2,2

< 1.374 2 1 + a1 = 2 1 ;

(XIV.8)

at introduction of the second polymer into the first one:

2 =

p2,2
p2,1

< 1.374 2 1 + a2 = 2 2 .

(XIV.9)

In the case as the one above-mentioned, a single glass transition temperature


Tg is displayed, and the dependence of Tg on the composition of the mixture has the
shape schematically represented in Figure 98, curve 2.
Let us discuss now the behavior of a series of particular well-studied
polymeric mixtures. The first of them is the mixture of polystyrene with
poly(vinylmethyl ether):
CH2

CH

Polymer 1

CH2

CH

Polymer 2.

O
CH3
The behavior of this mixture of various compositions is studied in detail in a
series of works [131, 153, 168, 198200]. The dependence of the glass transition
temperature Tg of this mixture on the composition is displayed in Figure 99.

Figure 99. Dependence of the glass transition temperature Tg on mole fraction of polystyrene for the
mixture of poly(vinylmethyl ether) with polystyrene (2). Dotted curve (1) represents
dependence of Tg on on the condition, as components are completely miscible.

Clearly, as polystyrene concentration increases from 0 to 40%, the change of


Tg is extremely weak and is almost independent of the composition of the mixture.

550

This dependence is not described by equation (IV.55) valid for statistic copolymers
and homogenous mixtures of polymers (curve 1).
Let us analyze in detail the miscibility of polystyrene (PS) and
poly(vinylmethyl ether) (PVME) with the help of criterion (XII.21). Suppose first that
PS is the solvent for PVME. The initial characteristics necessary for the application
of criterion (XII.21) are shown in Table 48. Substituting values of these
characteristics into equation (XIV.20), we obtain that

2= 0.684 < 1.218 = 22.


Table 48
Initial data for estimating compatibility of a series of polmers: polystyrene (PS), poly(vinylmethyl
ether) (PVME) and poly(2,6-dimethyl-1,4-phenylene oxide) (PPO)
Physical characteristics
PS
PVME
PPO
9.12
7.54
8.93
Solubility parameter , (cal/cm3)1/2
41.6
33.4
44.4
Surface energy , dyn/cm
Molar volume V, cm3/mol
97.08
60.7
103.75
22988
8696
23587
Ei* , J/mol
Cohesive energy
i
109.8
60.7
117.4
Vi , 3
Van-der-Waals volume
i
Glass transition temperature Tg , K
373
245

Hence, because the left part of the miscibility criterion is smaller than the right
part, then PVME is dissolved well in PS.
Suggest now that PVME is the solvent for PS. Then, with the help of
miscibility criterion (XIV.1) we obtain that

1= 1.462 > 1.175 = 21.


Because the left part of the criterion is greater than the right part, this means
that poly(vinylmethyl ether) is not dissolved in polystyrene. That is why when PS is
injected into PVME, a microphase separation should be expected. However, because
PS is the solvent for PVME, a part of PVME will be miscible with PS.
Consequently, two microphases will be formed, one of which contains PVME, and the
second a mixture of PVME with PS. When the PVME concentration in the second
microphase increases, miscibility of this microphase with PVME is improved and at a
definite concentration of PVME, the second microphase will be miscible with PVME.
Let us determine this critical concentration of PVME. For this purpose, let us write
down the correlation for calculating the solubility parameter of the current mixture of
polymers. Basing on equation (XII.3), we obtain that

Ei* + (1 ) Ei*

i
2
i
1
2
=
,
mixture


N A Vi + (1 ) Vi

i
2
i
1

(XIV.10)

551

where is the molar part of PVME in the mixture; Ei* and Ei* are

i
1
i
2

molar cohesive energies for PS and PVME, respectively; Vi and Vi

i
1
i
2
are Van-der-Waals volumes of repeat units of PS and PVME, respectively.
Substituting characteristics of polymers shown in Table 48 into correlation
(XIV.10), we obtain that
2
=
mixture

22,988 14,292
cal/cm3.
4.18(66.13 29.57 )

(XIV.11)

Let us now estimate the surface energy of the mixture according to equation
(XIII.39):

mixture = [C1n + (1 )C4 ]

Ei* +

i
2


N A Vi + (1 ) Vi

i
2
i
1

+ (1 ) Ei*

i
1

[m2 + (1 )m1 ]1 / 3

2/3

(XIV.12)

where C1n and C4 are coefficients from equation (XIII.39) for PVME (non-polar
polymer) and PS, respectively; m1 and m2 are the numbers of atoms in repeat units of
PS and PVME, respectively.
Substituting all the values from Table 48 into correlation (XIII.39), we obtain
that

mixture = (0.0231 0.1046)

22,988 14,292
(109.8 49.1 ) 2 / 3 (16 6 )1 / 3

(XIV.13)

Future analysis requires estimation of the molar volume of the mixture:


Vmixture = 53.68 + (1 )97.088.

(XIV.14)

To calculate left and right parts of criterion (XII.21), let us use correlations
(XIV.11), (XIV.13) and (XIV.14). Therewith, let us consider a two-component
mixture, one of the components of which represents PVME, and the second one is a
mixture (microphase) of PVME/PS type with different molar part of PVME.
Calculation results are displayed in Figure 100 in the shape of two dependences of
both parts of the criterion (XII.21) on the molar part of PVME. The cross-point of
these two dependences corresponds to the PVME concentration in the microphase at

552

which the miscibility of PVME with this microphase appears. This threshold
concentration cr = 0.62. The Van-der-Waals volume of the mixture with the
threshold concentration of PVME will equal

V = 60.70.62 + 109.80.38 = 79.36 3.


i

i
th

Figure 100. Dependences of (2) and 2 (1) on the molar fraction of poly(vinylmethyl ether)
within microphase (for explanation see text).

Now the glass transition temperature of the mixture with the threshold
concentration of PVME may be calculated. For this purpose, let us use equation
(IV.55):

th Vi Vi + Vi

i
2 i
1 i
1

=
Tg, th
. (XIV.15)

V
V
Vi

i i
i
2 i
1
i
1

+ 2 th (1 th ) 0.03
th
+
Tg,1
Tg,1
Tg,2

Substituting all parameters of the system from Table 48 into correlation


(XIV.15) and taking into account that th = 0.62, we obtain that Tg,th = 284 K. Now
the dependence of glass transition temperature Tg on the composition of the mixture
consisting of PVME and critical mixture, i.e. a mixture consisting of PS and PVME
with the critical concentration th of PVME equal to 0.62, may be determined. Using
equation (IV.55) for this purpose, we obtain that

553

Vi Vi + Vi

i
2 i
th i
th

Tg, th =
,

Vi Vi
Vi

i
th
i
th
2 i
+ 2 (1 ) 0.03

T
T
T
g,2
g, th
g, th

(XIV.16)

where is the molar part of PVME.


Correlation (XIV.16) is true only in the range of from 1 down to 0.62. The
curve composed in accordance with correlation (XIV.16) is shown in Figure 99 (curve
2). Clearly, experimental points fit well this dependence.
To describe the second part of the dependence of Tg on the composition of
PVME/PS mixture (inside the range of from 0.62 down to 0), it is necessary first to
determine the Van-der-Waals volume of the mixture, for which Tg = 284 K:


+ 0.380.62 Vi + 0.38 Vi ;

2
i
2
i
1

Vi = 0.62 Vi
i

Vi

= 67.8 3.

Now, the dependence of Tg on the composition of the mixture, one of the


components of which represents PS, and the second the microphase with Tg = 284 K
(see above), may be described. For this system, we obtain that
Tg =

(109.8 67.8) + 67.8


,
109.8 67.8 67.8

+
+

2
(
1
)
0
.
03

373 248 248

(XIV.17)

where is the molar part of PS on the scale from 1 down to 0.38. To determine the
true molar part of PS in the mixture, it should be presented:
(1 ) = 0.38 + 0.62,
where is the molar part of PVME, and (1 ) is the molar part of PS in the total
mixture.
(1 ) 0.38
into correlation (XIV.17), we find that the
Substituting =
0.62
correlation obtained describes well the second part of dependence of Tg in the range of
from 0 to 0.62 on the composition of the mixture at high concentration of PS (see
Figure 99).
Hence, solubility criterion (XII.21) may be successfully applied to description
of the dependence of glass transition temperature Tg of the polymer system, in which
one of polymers dissolves well the second one, but the first polymer is badly
dissolved in the second one.

554

Let us analyze now a simpler case, when two polymers are totally miscible
with each other.
As an example of such mixture, let us consider the polystyrene (PS)poly(2,6dimethyl-1,4-phenylenoxide) (PPO) system, studied well in a series of works [139,
166, 169, 197, 203, 204, 209].
Suggest that PS is the solvent for PPO. Applying solubility criterion (XII.21)
with regard to all physical characteristics of the mixture components (see Table 48), it
is found that 2 = 22 12 = 0.959; = 1.0; 1,2 = 0.0456; a = 0.001096; 22 = 1.328
(1 and 2 are solubility parameters of PS and PPO, respectively). Because 2 < 22,
then according to criterion (XII.17) PS is a good solvent for PPO, i.e. miscibility
must be observed. It should be assumed now that PPO is the solvent for PS. Then 1
= 12 22 = 1.043; = 1.0; 1,2 = 0.0456; a = 0.00103; 21 = 1.33. Clearly, in this
case also, the left part of criterion (XII.21) is smaller than the right part of it, i.e.
miscibility is also possible. Experiments indicate that, indeed, polymers in the present
pair are totally miscible. This leads to a description of the glass transition temperature
of the PS/PVME mixture by the same dependence on the composition as for statistic
copolymers, i.e. by correlation (IV.55). This dependence is shown in Figure 101,
which indicates that experimental points fit well the calculated curve. It should be
noted that the approach considered possesses the advantage before some other
approaches that it requires no introduction of fitted parameters. To put it differently,
this analysis requires understanding of the chemical structure of components, based
on which all physical characteristics of polymers are calculated.

Figure 101. Dependence of the glass transition temperature Tg on the molar fraction of poly(2,6dimethyl-1,4-phenyleneoxide) for the polymer blend with polystyrene.

Chapter XV. Influence of the end groups on the properties


of polymers

The problem of influence of the molecular mass of the polymer and the
problem of the end groups role and their influence on the properties associated with it
has been discussed many times in the literature. For example, there are data on the
dependence of glass transition temperature Tg and melting point Tm on the molecular
mass in the shape of curves with a plateau, as well as information on other properties
dependent on the molecular mass.
As an example, Figure 21 displays the dependence of Tg on the mean
molecular mass of polystyrene. There is a series of correlations describing these
dependences, for example [144],
Tg = Tg() k/M,
where Tg() is the glass transition temperature of the polymer at the molecular mass
M ; k is the polymer constant.
Dependences of this kind do not describe Tg(M) curves along their length.
More complicated dependence Tg(M) is of the form [145]:
Tg = Tg() K*/(M + M*),
where K* and M* are fitted parameters. Obviously, not only glass transition
temperature Tg, but also all other properties depend on the molecular mass. Searching
for these dependences is of special interest.
This problem is of prime importance for analysis of the molecular mass
influence on the properties of propagating chains at a very low number of units n. The
circumstance that the chemical structure of end groups may significantly (and even
principally sometimes) differ from the structure of the repeat unit of the polymer is
not always taken into account in this analysis. This causes that the composition of the
dependence of a parameter of any physical property on n indicates every new value of
n nonequivalent to the former one, because the chemical structure of the averaged unit
changes with transition from n to (n + 1), etc. Only when n 10 20, the influence of
end groups becomes low and dependences of the properties on n approach the values
of parameters typical of the repeat unit of the polymer.
Ref. [8] indicates quantitative description of the dependences of three physical
parameters Van-der-Waals volume, molar refraction and molar cohesive energy
on the number of units in a polymeric chain starting from n = 1. Therewith, the
influence of the type of initiator applied to polymerization on the chemical structure
of end groups and properties of dimers, trimers, etc. followed from it, is taken into
account. Calculations were performed on the example of four polymers poly(methyl
methacrylate) (PMMA), polystyrene (PS), poly(ethylene terephthalate) (PET) and
polycarbonate (PC) based on bisphenol A.
The following types of systems were considered:

556

CH3
H3C

(CH2

C OCH3

CH3

CH3

C)n- 2 CH2
C OCH3

CH
C OCH3

O
CH3

O
CH3
C

C OCH3

C)n- 2 CH
C OCH3

O
CH3
C

CH

(CH2

CH3
H3C

PMMA-2

C OCH3

CH3

CH3
C

C OCH3

C)n- 2 CH
C OCH3

C OCH3

(CH2

O
C

(CH2

O
C

PMMA-1

PMMA-3

CH3

CH3

C)n- 1 CH2
C OCH3

CH
C OCH3

PMMA-4

CH3 O
(CH2

PMMA-5

C)n C
C OCH3
O

H3C

CH

(CH2

CH)n - 2

CH2

H3C

CH

(CH2

CH)n - 2

CH

CH2

CH

PS-1

PS-2

O
H3C

CH

(CH2

CH)n- 1

PS-3

557

HO

HO

(CH2

CH)n C

[C

C O

[C

C O

(CH2)2

PS-4

(CH2)2

O]n H

PET-1

C OH PET-2

O]n C
O

CH3
H

[O

PC-1

O C] n OH
O

CH3
CH3
H

[O

O C] n O
O

CH3

CH3

CH3
H

[O

PC-2

O C] n O

CH3

PC-3

CH3

For all structures, Van-der-Waals volumes

OH

Vi , molar refractions Ri
i

and molar cohesive energies Ei* have been calculated. As an example, let us
i

consider in detail the calculation of these characteristics for PMMA-1 structure.

Vi
= 3VC,13 + VC,1 + 2VC,48 + 2VC,41 + VC,10 + VC,6 + 18VH,124

i
PMMA -1
+ 2VO,129 + 2VO,139 + (VC,10 + VC,1 + VC,13 + VC,48 + VC,41 + 8VH,124 +
VO,129 + VO,139)(n 2);

Vi
= 317.2 + 5.0 + 215.9 + 220.3 + 13.1 + 9.0 + 182.0 + 23.4 +

PMMA -1
25.8 + (13.1 + 5.0 + 17.2 + 15.9 + 20.3 + 82.0 + 3.4 + 5.8)(n 2) = 205.5 +
(n 2)96.7 3.

558

*
*
*
*
*
*
Ei*
= 10 EC
+ 18 EH
+ 2 Ed* (5 EC
+ 8 EH
+ 4 EO
+ 2 EO

i
PMMA -1
+ Ed* )(n 2);

Ei*
= 10550.7 + 1847.7 + 4142.6 + 21623 + (5550.7 + 847.7 +

i
PMMA -1
2142.6 + 1623)(n 2) = 10,182 + (n 2)5,043 cal/mol = 42,561 + (n 2)21,080
J/mol.

Ri
= 10RC + 18RH + 2RO< + 2RO= + (5RC + 8RH + RO< + RO=)(n 2);

i
PMMA -1

Ri
= 102.418 + 181.100 + 21.643 + 22.211 + (52.418 + 81.100 +

i
PMMA -1
1.643 + 2.211)(n 2) = 51.688 + (n 2)24.744 cm3/mol.
Besides these characteristics, let us also calculate heat capacity in the liquid
state Cpl . For PMMA-1, we obtain:

(Cpl )PMMA-1 =

l
Cp,
C (3VC,13 + VC,1 + 2VC,48 + 2VC,41 + VC,10 + VC,6) +

l
l
l
Cp,
H 18VH,124 + C p, O (2VO,129 + 2VO,139) + [ C p, C (VC,10 + VC,1 + VC,13 +
l
l
VC,48 + VC,41) + Cp,
H 8VH,124 + C p, O (VO,129 + VO,139)](n 2);

(Cpl )PMMA-1 = 0.34565(317.2 + 5.0 + 215.9 + 220.3 + 13.1 + 9.0) + 0.62289182.0

+ 0.92998(23.4 + 25.8) + [0.34565(13.1 + 5.0 + 17.2 + 15.9 + 20.3) +


0.6228982.0 + 0.92998(3.4 + 5.8)](n 2) = 91.8 + 43.2(n 2) cal/(moldeg).

Values

Vi ,
i

Ei* ,

Table 49
l
Ri and Cp for PMMA-5 with different values of the

polymerization degree n
n
1
2
3
4
5
7
10
15
20

Vi , 3

i
0

Ei* , cal/mol

i
0

Ri , cm3/mol

i
0

297.0
197.0
163.0
146.5
136.6
115.6
107.0
100.3
97.0

17541
11292
9209
8168
7543
6363
5828
5411
5203

84.42
54.50
44.70
39.80
36.60
31.60
29.07
27.08
26.08

l
,
Cp,0
cal/(moldeg)
125.0
84.5
70.7
63.8
59.6
50.3
46.8
44.0
42.7

559

Analogous calculations were also performed for all other structures. Then
values obtained for

Vi , Ei* , Ri
i

and Cpl were recalculated for a single

unit of polymer by simple division by n, and dependences Vi n = Vi ;

i
0

l
Ei* n = Ei* , Ri n = Ri and Cpl n = Cp,0
on the number

i
0 i

i
0
of units n have been plotted. These dependences for PMMA with various end groups


are presented in Figure 102. Table 49 showing values of Vi , Ei* ,

i
0 i
0

l
Ri and Cp,0
depended on n for PMMA-5, is also displayed as an example.

i
0

Figure 102. Dependences of

'
L

(a),


L

(b),

#
L

l
(c) and C p,0 (d) as a

function of degree of polymerization n for poly(methyl methacrylate). Numbers of


curves correspond to the numbers of PMMA specimens (see text).

Curves in Figure 102 and Table 49 clearly display that dependences of the
Van-der-Waals volume, the cohesive energy, the molar refraction and heat capacity

represent by shape the curves asymptotically approaching values Vi ,

i
0

l
Ei* , Ri and Cp,0
typical of the repeat unit of the polymer. Therewith,

i
0 i
0
the shape of curves is different for different end groups of the polymer. In some cases,
when end groups make no significant contribution into the characteristics considered,

l
even at low n values of Vi , Ei* , Ri and Cp,0
differ

i
0 i
0 i
0
insignificantly from asymptotic one, typical of the current chain. In the case, when the

560

end group possesses a significant Van-der-Waals volume, it contributes substantially


to the physical characteristics, and values Vi , Ei* , Ri and

i
0 i
0 i
0
l
Cp,0
at low n appear sharply different from asymptotic ones. The asymptote is

approached (depending on type of the end group) at n = 1020, which is assumed


sometimes for the value of the chain segment. This situation is typical, for example,
of PMMA-4, when benzoyl peroxide is used as an initiator in polymerization of
methyl-methacrylate.
Analyzing the data obtained in ref. [8], some other regularities of the influence
of the chemical structure of end groups on the physical characteristics of the polymer
may be observed. For example, the Van-der-Waals volume of the averaged repeat unit
of the polymer is generally affected by the volume of end groups. The same may be
indicated for molar refraction. Consideration of Figure 102 confirms this conclusion.
For systems PMMA-1, PMMA-2 and PMMA-3, the Van-der-Waals volume and the
molar refraction weakly depend on n. This is clearly indicated by consideration of
curves 13 in Figure 102. The same may be also said about the cohesive energy of
these systems. This is associated with the fact that in the case of PMMA-1, PMMA-2
and PMMA-3, end groups do not enter any specific intermolecular interaction. That is
why they cause no significant change of the cohesive energy. In turn, the presence of
bulky end groups significantly different from the repeat unit of the polymer by the
chemical structure leads to a sharp change of all the characteristics discussed (see
Figure 102, curves 4, 5). All the above-said also relates fully to systems based on
polystyrene.
Systems containing end groups, although of small volume but possessing a
strong specific intermolecular interaction, behave themselves differently. For
example, they may be end OH-groups in the case of poly(ethylene terephthalate) and
polycarbonate. This specificity of the structure of end groups leads to different
regularities: the characteristics associated with volume and molecular refraction
change insignificantly with n; in turn, cohesive energy, in which specific
intermolecular interaction plays a highly significant role, changes greatly as n
increases. All this must be taken into account when the influence of the
polymerization degree (especially, on initial stages of the process) on the measured
physical characteristics is analyzed.
Above, we analyzed the physical characteristics of a series of polymers
dependent on the chain length. These characteristics are initial for the estimation of
the physical parameters of polymeric substances. For example, due to the Van-der

Waals volume Vi and cohesive energy Ei* important parameters of

i
0
i
0
the substance such as the density of cohesive energy or Hildebrand solubility
parameter may be calculated. Of interest is also the dependence of this parameter on
n at transition from monomer to dimer, trimer, etc.
These calculations were performed in ref. [8] for systems PMMA-4, PS-4,
PETP-1 and PC-1. Calculation results in the form of dependence of on n are
displayed in Figure 103. One can recognize that for some systems (PMMA-4, PS-4)
regardless of whether they contain bulky end groups or not, the solubility parameter
depends significantly on n. This is associated with the fact that the presence of end
OH-groups in these polymers causes the appearance of hydrogen bonds that requires

561

introduction of constants Eh* which take into account the contribution of the energy
of hydrogen bonds to total cohesive energy. As n increases, the contribution of these
bonds becomes weaker and at n = 10 20 is almost negligible.

Figure 103. Dependence of solubility parameter on the degree of polymerization n for PS-4 (1),
PMMA-4 (2), PC-1 (3) and PET-1 (4).

Besides the solubility parameter, of special interest is also consideration of the


influence of the chemical structure of end groups on surface tension.
Figure 104 indicates the dependence of the surface tension on polymerization
degree n for PMMA-5. Obviously, the surface tension decreases as n increases,
striving to the border value typical for PMMA.

Figure 104. Dependence of the surface tension p as a function of the degree of polymerization n for
PMMA-5.

Hence, consideration of the influence of the chemical structure of end groups


of polymers on their molecular characteristics and macroscopic properties indicates
that this influence may be quite different. The presence of the effect of bulky end
groups affects mainly the Van-der-Waals volume of the averaged unit, cohesive
energy and molar refraction. The presence of groups (of the hydroxyl type) with small
volume but possessing specific intermolecular interaction has a strong effect on the
cohesive energy and solubility parameter, as well as on surface tension, with a small
influence on the Van-der-Waals volume and molar refraction. Considering the
dependence of any physical properties on the chain length n, these factors must be
taken into account.

Chapter XVI. Thermophysical properties of polymers

The thermophysical properties are heat capacity, thermal diffusivity and heat
conductivity.

XVI.1 Heat capacity


Heat capacity is the amount of heat which must be consumed for heating a
body up by 1C. Molar heat capacity, if a mole of substance is considered, and
specific heat capacity in the case of 1 g of substance, differ. Heat capacity at constant
pressure Cp equals the rate of enthalpy change with temperature increase, and heat
capacity at constant volume Cv the rate of internal energy change with temperature
increase.
In a quite wide temperature range, heat capacity increases linearly with
temperature, the temperature coefficient of heat capacity increase for solid polymers
possessing the average value of 3103. At phase or physical transition of the polymer,
heat capacity changes in a jump-like manner.
For example, transition from the glassy to rubbery state indicates quite a sharp
jump-like increase of heat capacity. When the physical transition is passed, heat
capacity returns to weak growth with temperature.
The heat capacity of polymers depends on their chemical structure (Table 50).
Among polymeric hydrocarbons, the lowest heat capacity is displayed by
polyethylene and polyoxymethylene.
When hydrogen atoms are substituted by polar groups, heat capacity increases.
Also, heat capacity rapidly increases in transition from aliphatic to aromatic polymers.
Many attempts have been made to calculate the heat capacity of polymers on
the basis of the chemical structure of the repeat unit. Let us consider the method of
calculation discussed in ref. [22]. In the work cited, it is supposed that the molar heat
capacity of a polymeric substance is proportional to the Van-der-Waals volume of
atoms of the repeat unit of the polymer. To put it differently,
C sp = C sp, i Vi + As

(XVI.1)

C lp = C lp,i Vi + Al ,

(XVI.2)

and
i

where C sp and C lp are the molar heat capacities of the polymer in the glassy state and
the rubbery state, respectively; C sp,i and C lp,i are constants for each atom having the
meaning of heat capacities reduced to the unit Van-der-Waals volume and applicable
for the glassy state and the rubbery state, respectively; As and Al are parameters equal
to: As = 0.77 cal/(moldeg), Al = 0.69 cal/(moldeg).

563

Table 50
Calculated and experimental values of molar heat capacities for a series of polymers
C sp , cal/(moldeg)
C lp , cal/(moldeg)
Polymer
Calc.
Experimental
Calc.
Experimental
Polyethylene
11.02 10.4; 11.1; 11.8; 10.15
14.7
15.1
Polypropylene
16.9
15.7; 16.5
21.7
21.6
Polyisobutylene
22.8
22.4
28.8
26.4
Poly-4-methylpentene-1
34.6
33.6

Polybutadiene
20.0
21.0
26.3
24.3
Polyisoprene
25.8
25.7
33.3
31.3
Polystyrene
32.4
30.5; 29.3
43.0
42.6
Polyoxymethylene
8.8
10.2; 8.9
13.2
15.0
Polyoxytetramethylene
26.4
28.2
34.1
35.7
Polyoxypropylene
19.8
19.7
26.0
26.5
Poly(vinyl chloride)
16.2
14.4; 13.8; 16.2;
18.1
18.1
Poly(chlorotrifluoroethylene)
25.4
25.0; 23.3

Poly(tetrafluoroethylene)
23.4
23.0; 21.1
23.0
23.0
Poly(vinyl alcohol)
16.3
13.6; 15.4

Poly(vinyl acetate)
27.2
30.2; 23.6
37.0
39.5
Poly(methyl methacrylate)
33.1
33.0; 32.0
44.0
43.5
Poly(acrylonitrile)
15.5
15.9

Polyamide-6
40.2
39.1; 39.2
59.4
57.8
Polyamide-6,6
81.1
79.0; 76.6

Polyamide-6,10
104.7
107
146.2
147.0
Poly(ethylene terephthalate)
52.9
52.0; 52.2
73.3
71.0
Polycarbonate based on bisphenol A
72.7
72.4; 71.0
97.6
97.8
Poly(propylene sulfone)
29.4
29.4
38.8
38.8
Poly-2,6-dimethyl-1,4-phenyleneoxide
35.1
36.4; 34.9
47.0
50.5
Table 51

Atom
Carbon
Hydrogen
Oxygen
Nitrogen
Fluorine
Chlorine
Sulfur

s
l
Values C p and C p for different atoms
Numerical value,
Designation
Designation
cal/(moldeg)
3
0.232030
C Cs
C Cl
s
l
0.714129
CH
CH
l
0.634726
C Os
CO
s
l
0.314997
CN
CN
0.543367
C Fs
C Fl
l
s
0.368819
C Cl
C Cl
0.273109
CSs
CSl

Numerical value,
cal/(moldeg)
3
0.345646
0.622889
0.929977
2.099874
0.444909
0.284693
0.303031

Values of C sp,i and C lp,i were obtained by regression analysis with the help of
solving the excessive system of equations deduced on the basis of correlation (XVI.1)
or (XVI.2) using the experimental data on heat capacity for well-studied polymers
(so-called polymeric standardspolyethylene, polystyrene, poly(methyl methacrylate),
etc.).

564

Values of C sp,i and C lp,i obtained for each atom are shown in Table 51. With
the help of these values and Van-der-Waals volumes displayed in Table 3, molar heat
capacities C sp and C lp may be calculated for many polymers. Satisfactory accuracy
of calculation is clear from Table 50.
For copolymers, equations (XVI.1) and (XVI.2) are reduced to the following
form:

C sp = 1 C sp,i Vi + 2 C sp,i Vi + ... + n C sp,i Vi + As ,

i
1
i
2
i
n

(XVI.3)

C lp = 1 C lp,i Vi + 2 C lp,i Vi + ... + n C lp,i Vi + Al ,

i
1
i
2
i
n

(XVI.4)

where 1, 2, , n are the molar parts of the components 1, 2, , n; C sp, i Vi ,

i
1

C lp, i Vi , C sp, i Vi , C lp, i Vi , , C sp, i Vi , C lp, i Vi

i
n i
n
i
1 i
2 i
2
are the sets of constants for the components 1, 2, , n, respectively.
If it is required to express the heat capacity of a copolymer via the heat
capacity of homopolymers based on the components 1, 2, , n, it should be written
down that
C sp = 1C sp,1 + 2C sp,2 + ... + n C sp, n ,

(XVI.5)

C lp = 1C lp,1 + 2C lp,2 + ... + nC lp, n ,

(XVI.6)

where C sp,1 , C sp,2 , , C sp, n are the molar heat capacities of polymers based on the
components 1, 2, , n in the glassy state; C lp,1 , C lp,2 , , C lp, n are the molar heat
capacities of polymers based on the components 1, 2, , n in the rubbery state.

XVI.2 Thermal diffusivity and heat conductivity


Thermal diffusivity is a characteristic that describes the rate of temperature
propagation under the effect of heat flow under non-stationary temperature
conditions. This characteristic is determined from the correlation:

Cp

(XVI.7)

where is heat conductivity; Cp is the specific heat capacity under constant pressure;
is polymer density.

565

For solid (glassy and crystalline) polymers, thermal diffusivity decreases


slowly with temperature increase. However, when a polymer transits from the glassy
to the rubbery state, thermal diffusivity decreases abruptly.
Thermal diffusivity depends on the chemical structure of the polymer, as well
as on the crystallinity degree, molecular mass and pressure. Table 52 shows the values
of thermal diffusivity for a series of polymers with different chemical structures.
Table 52
Polymer

Thermal diffusivity for a series of polymers at 293 K


Structural formula

High pressure polyethylene


Polyisoprene

CH2CH2
CH2

CH C CH2

107,
m2/sec
1.40
0.90

CH3
Poly(ethylene oxide)
Poly(dimethyl siloxane)

CH2CH2O

1.08

CH3
Si

0.90

CH3
Polystyrene

CH2

Poly(vinyl chloride)

CH2

CH

CH

0.99

1.21

Cl
Poly(methyl methacrylate)

CH3
CH2

1.19

C
C O CH3
O

Heat conductivity represents the capability of polymeric substances for heat


transfer from more heated substances to less heated ones. The coefficient of heat
conductivity represents the proportionality coefficient between the heat flow and the
temperature gradient. Heat conductivity is associated with the propagation and
dissipation of elastic waves caused by thermal oscillations of particles in the
substance. At temperatures approaching the absolute zero, heat conductivity also
strives to zero. The heat conductivity of polymers depends on their chemical structure
in the frame of one and the same physical state. For amorphous glassy polymers, heat
conductivity is somewhat lower than for crystalline polymers. The temperature
dependence of heat conductivity is affected by the chemical structure of the polymer.
For one series of polymers, the slope of this dependence is positive, and for another
one negative. In the area of physical transition from the glassy to rubbery state, a
low maximum of heat conductivity is observed. Table 53 shows heat conductivity
values for a series of polymers with different chemical structures, from which the type
of the effect of the chemical structure on heat conductivity is clear. Of special
meaning is the anisotropy of the heat conductivity of polymers. This property is
typical of oriented polymeric systems, in which heat conductivity is different along

566

and transverse to the orientation axis. The thermal properties of polymers are
discussed in more detail in reviews [59, 61, 62].
Polymer
Polyethylene of:
High pressure
Low pressure
Polypropylene:
Atactic
Isotactic
Polystyrene

Heat conductivity for a series of polymers at 293 K


Structural formula

Poly(vinyl chloride)

CH2CH2
CH3

Table 53

, W/(mK)
0.380
0.470

CH2

CH

CH2

CH

0.175
0.230
0.130

CH2

CH

0.170

Cl
CH3

0.190

Poly(methyl methacrylate)

CH2

C
C O CH3

Polytetrafluoroethylene
Poly(dimethyl siloxane)

O
CF2CF2
CH3

Si
CH3

0.250
0.167

Chapter XVII. Molecular design and computer synthesis of


polymers with predetermined properties

In previous sections of the book, the possibility of calculating the physical


properties of polymers on the basis on the chemical structure of the repeat unit of
linear polymers or the repeating fragment of polymer networks was demonstrated. At
the present time, the above-mentioned calculations are performed in a computer.
Therewith, main direct and reverse tasks and secondary tasks may be solved.
The direct task concludes in calculating the polymer properties from the data
on the chemical structure of the repeat unit of the polymer or a fragment of network.
The reverse task (which is more complicated) consists of predicting the
structural formula of the repeat unit of the polymer which would have a single or
several required characteristics of the physical properties.
A solution of both tasks is described in refs. [9, 17, 26, 126].
When setting and solving problems, two border approaches and a series of
intermediate approaches are possible. The first border approach concludes in
recording in computer memory a large number of polymers with different chemical
structure for which the physical characteristics mentioned above were previously
calculated. Then, computers role will conclude only in the search for necessary
structural formulas in the database and layout of the required information, when both
direct and reverse tasks are solved. In the case of such approach, the accuracy of
layout characteristics is close to 100%, but the predictive power for structures absent
in the database equals zero. In fact, such approach enables just to create a database.
The second border variant concludes in the following. A table of Van-derWaals volumes of atoms is recorded in the computer memory. A fragment of this
Table is shown as an example below.
H
O
H
C
1.54
1.54

1.54

1.08
1.54

1.28

1.54

C C

1.54

1.08
1.54

C C

1.54

1.08

1.54

1.54

H
C

1.28

1.08

1.54

1.54

1.54

1.54

1.50
1.54

1.54

In principle, each of the atoms squared by dotted lines may be valently bonded
to other atoms; for the above-depicted structures, this possibility is demonstrated by
the matrix (1 joining is possible; 0 prohibition):
Atom, No.
1
2
3
4
5
6
7
8

1
1
1
1
1
1
0
0
0

2
1
1
1
1
1
1
0
0

3
1
1
1
1
1
0
1
0

4
1
1
1
1
1
1
0
0

5
1
1
1
1
1
0
0
1

6
0
1
0
1
0
0
0
0

7
0
0
1
0
0
0
0
0

8
0
0
0
0
1
0
0
0

568

The table representing Van-der-Waals volumes of atoms, shown in the present


monograph, contains over 200 values of volumes for atoms valently bonded with
other various atoms. The task of building a repeat unit of a polymer from all this
multiplicity of atoms, even at their limited amount in the repeat unit, is rather difficult
to be computerized, and its solution by a computer requires too much computer time
due to the appearance of a tremendous amount of variants. In the present case, we are
talking about solving the reverse problem, when polymers with given properties are to
be synthesized using a computer.
The above-discussed situation led to the need to develop an intermediate
approach (between the border ones described) to computerized prognosis of the
physical characteristics of polymers and their computerized synthesis. This approach
concludes in preliminary summation of all constants participating in the initial
correlations for calculation of properties for separate fragments (basic fragment),
from which then a repeat unit of a polymer is constructed on a computer.
The variant of the software described in ref. [17] provides constructing of
polymers, repeat units of which contain two fragments only possessing a possibility of
chemical bonding with each other. According to the method of mutual joining, all the
fragments discussed are subdivided into 8 classes:
Class
I

Structure
C

II

III

IV

1.50

1.37

where R1: CH2;

O
O

1.54

1.48

R1
R2

R1,2
R1,2

Class
V

1.54

C
C

VI

1.48

Structure

1.50

VII

1.37

VIII

C
C

O
N
C

; (CH2)4, etc.; R2:

CH

1.50

1.37

1.37

1.37

1.50

R1,2

1.37

R2

1.37

R1,2
R3

1.37

OH
, etc.;

O
O
C
R3:,

C
N

O
O

C
N

, etc.

The matrix shown below reflects possibilities of mutual joinings of the


fragments related to different classes (1 joining is possible; 0 prohibition):

569

Fragment, No.
I
II
III
IV
V
VI
VII
VIII

I
1
0
0
0
0
0
0
0

II
0
1
0
0
0
0
0
0

III
0
0
0
0
1
0
0
0

IV
0
0
0
0
0
1
0
0

V
0
0
1
0
0
0
0
0

VI
0
0
0
1
0
0
0
0

VII
0
0
0
0
0
0
0
1

VIII
0
0
0
0
0
0
1
0

Generally, ref. [17] presents series of constants calculated for 194 fragments
(basic fragments) of various chemical structures. As an example, some of them are
shown in Table 54. Dotted squares mark the chemical structure of the fragment itself;
also displayed are atoms, joined to the fragment, and lengths of chemical bonds
expressed in .
The possibility of joining of one of 18 fragments to another, taking into
account their belonging to different classes, is demonstrated in the following matrix
(1 joining is possible; 0 prohibition):
Fragment, No.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18

1
1
1
1
0
1
1
1
1
1
0
0
0
0
0
0
0
0
0

2
1
1
1
0
1
1
1
1
1
0
0
0
0
0
0
0
0
0

3
1
1
1
0
1
1
1
1
1
0
0
0
0
0
0
0
0
0

4
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1
0

5
1
1
1
0
1
1
1
1
1
0
0
0
0
0
0
0
0
0

6
1
1
1
0
1
1
1
1
1
0
0
0
0
0
0
0
0
0

7
1
1
1
0
1
1
1
1
1
0
0
0
0
0
0
0
0
0

8
1
1
1
0
1
1
1
1
1
0
0
0
0
0
0
0
0
0

9 10
1 0
1 0
1 1
0 0
1 0
1 0
1 0
1 0
1 0
0 1
0 1
0 1
0 1
0 0
0 0
0 0
0 0
0 0

11
0
0
0
0
0
0
0
0
0
1
1
1
1
0
0
0
0
0

12
0
0
0
0
0
0
0
0
0
1
1
0
0
0
0
0
0
0

13
0
0
0
0
0
0
0
0
0
1
1
0
0
0
0
0
0
0

14
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1

15
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1

16
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1

17
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0

18
0
0
0
0
0
0
0
0
0
0
0
0
0
1
1
1
0
0

The program contemplates both the solutions of the direct task concluded in
the determination of polymer properties basing on the chemical structure of the repeat
unit and the reverse task consisted in searching for the units such that the polymer
composed from them would possess the given physico-chemical properties. Solving
the reverse task, the computer calculates a parameter of one of the required properties
using all probable variants of combinations from basic fragments recorded by its
memory. For polymers, the property of which appears in the required interval, the
parameter of another property is calculated, after which the required polymers are also
selected, etc.
The drawback of this software is that composition of a repeat unit of a
polymer from two large basic fragment does not embrace the tremendous amount of
probable structures. Obviously, the smaller the size of basic fragments used for
composing the repeat unit, the more variants of polymers may be obtained due to the
great amount and multiplicity of basic fragments in the unit and for calculating their
properties.

570

That is why A.A. Askadskii, E.G. Galpern, A.L. Chistyakov, and I.V.
Stankevich [126] have created a software for computer synthesis of polymers from
the scarcest basic fragments, i.e. those which, in principle, cannot be sectioned.
These basic fragments are shown in Table 55.
To denote the possibility of joining these small fragments, recorded in the
computer memory, to each other, definite labels are attributed to every end of basic
fragments. Note that the fragments used for composing the repeat unit possess only
20 different end groups and, consequently, 20 different labels (see Table 55), which
may interact with each other in accordance with the matrix recorded by the computer
memory:
LABELS
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

1
1
1
1
0
0
0
0
0
0
0
0
0
0
1
0
0
1
0
0
1

2
1
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

3
1
1
1
0
0
0
0
0
0
0
0
0
0
1
0
0
0
0
1
0

4
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0

5
0
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0

6
0
0
0
0
0
0
0
0
0
0
1
0
0
0
1
0
0
1
0
0

7
0
0
0
0
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0

8
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

9 10
0 0
0 0
0 0
0 0
1 0
0 0
0 0
0 0
0 0
0 0
0 0
0 0
0 1
0 0
0 0
0 0
0 0
0 0
0 0
0 0

11
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
1
0
0
0
0

12
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0

13
0
0
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0

14
1
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

15
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0

16
0
0
0
0
0
0
0
0
0
0
1
0
0
0
0
0
0
1
0
0

17
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

18
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
1
0
0
0
0

19
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

20
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

In this matrix, no difference is made for the given CC bond of 1.54 (in
aliphatic fragments) and 1.48 (in aromatic fragments), because transition from one
bond to another causes only insignificant changes in the Van-der-Waals volume,
which display no substantial influence on the accuracy of calculation of physical
properties.

Figure 105. Schematic representation of construction of the combinatorial tree when constructing a
repeat unit of polymer (see text).

571

Table 54
Fragment
No.

Class

1
1

2
I

Chemical structure of fragment

Chemical structure and sets of constants for various fragments


3
Vi , M (ai Vi + bi ) 10 , K i Vi 103, ( i Vi + i ) ,
i

1.54

1.54

CH2

1.54

1.54

3
4
17.1

5
14

3K1
6
80.2

43.9

44

79.6

Ri ,

Ei* ,

Ci 103,

cal/mol
10
646

(cm3/mol)MPa1
11
3.090

3K1
7
24.3

8
1.434

cm3/mol
9
4.618

129.0

62.3

3.122

10.879

3058

9.458

86

195.2

129.5

6.778

15.290

6020

17.871

82.1

76

135.8

100.7

7.028

24.107

4208

12.686

24.5

30

33.9

54.2

2.076

6.143

4718

10.061

124.9

144

449

194.4

10.440

31.016

7502

16.618

CH
O CH3

CH3
C

1.54

1.54

C
C O CH3
O

VI

O
5

1.37

1.37

1.54

CH

1.54

OH
6

1.54

C O
O

(CH2)4

O C

1.54

O
571

572

2
I

1.54

C O

O C

O
8

1.54

6
231.8

7
191.6

8
11.090

9
36.651

10
9126

11
16.926

128.9

142

216.2

226.9

9.911

34.934

14334

17.106

136.5

162

208.0

226.4

10.655

40.570

15958

19.113

75.6

76

135.7

96.2

6.463

24.107

4208

12.686

159.8

168

319.6

200.0

14.521

49.864

8559

22.174

130.6

164

273.4

190.6

11.047

36.651

9126

18.626

1.54

C HN (CH2)4 NH C

1.54

I
1.54

5
164

O
9

4
131.0

C HN
O
NH C

1.54

O
10

II

C
11

1.48

II

C
12

1.48

1.48

1.48

II
1.54

C O
O
O

C
O

1.54

572

1
7

573

1
13

2
II

1.48

C HN (CH2)6 NH C
O

14

16

1.50

III

1.50

1.50

1.37

7
274.4

8
12.736

9
44.170

10
15626

11
23.295

163.1

184

395.4

199.0

14.786

51.500

8701

18.976

3.4

16

78

0.20

0.211

1.643

143

+3.198

87.5

108

259.7

103

7.363

27.393

4494

6.290

86.3

108

273

103

7.288

27.393

4494

7.990

74.6

56

321

104.9

6.318

18.472

2584

12.378

O
O

18

6
376.6

1.50

IV

5
170

1.50

III
1.50

4
162.7

C
17

III

C
15

1.48

1.54

(CH2)4

1.54

573

1.37

574

Fragment
No.
1
1
2

CH

Molar
refraction Ri,
cm3/mol
8
4.618

Cohesive energy
*, cal/mol
9
646

1.48; 1.54

1.48; 1.54

34.2

9.236

2915

1.54

1.54

51.3

13.854

1938

1.48; 1.54

1.48; 1.54

51.3

13.854

3561

1.48; 1.54

1.48; 1.54

43.9

10.879

3058

1.48; 1.54

1.48; 1.54

31.9

9.485

1999

1.48; 1.54

1.48; 1.54

46.8

12.752

2705

1.48; 1.54

1.48; 1.54

92.6

28.725

4854

CH3
3

CH3
C

CH3
4

CH

C2H5
5

CH

O CH3
6

7
8

CH

Cl
CCCl2C
C
CH
C

574

Table 55
Smallest basic fragments for the computer synthesis of polymers
Length of bonds with
Labels
Chemical structure of fragment
Van-derneighboring atoms,
Waals volume,
3
Left
Right
Left
Right
2
3
4
5
6
7
1.48; 1.54
1.48; 1.54
I
I
17.1
CCH2C

575

1
9

CH3

10

11

O
C

3
1.48; 1.54

4
1.48; 1.54

5
2

6
2

7
109.8

8
33.343

9
7123

1.48; 1.54

1.48; 1.54

18.6

4.629

693

1.48; 1.54

1.48; 1.54

108.9

32.797

7218

1.48; 1.54

1.48; 1.54

155.2

40.234

8966

1.48; 1.54

1.48; 1.54

81.1

14.694

1797

1.48

1.48

75.6

24.107

4208

O
C
12

13

CF3
C

CF3
14

C
575

576

3
1.48

4
1.48

5
3

6
3

7
75.6

8
24.107

9
4208

1.54

1.54

54.6

13.399

6084

1.48

1.48

100.3

33.127

8791

1.54

1.50

20.2

4.618

646

1.54

1.50

37.8

9.236

2915

1.54

1.50

54.9

13.854

3561

1.54

1.50

96.2

24.725

4854

1.48

1.37

78.8

24.107

4208

16

N
C

O
17

NH

N
18
19

CCH2O
C
CH
O
CH3
CH3

20

CH3
21

CH

22

576

1
15

577

1
23

25

4
1.37

5
3

6
5

7
78.8

8
24.107

9
4208

1.50

1.50

23.3

4.618

646

1.37

1.37

82.1

24.107

4208

1.37

1.37

82.1

24.107

4208

1.37

1.37

115.5

33.343

8746

1.50

1.50

3.4

1.643

143

1.37

1.50

2.7

1.643

143

1.37

1.37

2.1

1.643

143

1.54

1.50

25.1

6.272

2460

1.54

1.37

24.4

6.272

2460

1.54

1.37

18.6

4.618

646

1.54

1.37

20.1

4.629

2316

OCH2O

24

3
1.48

26

27

CH3
O

O
CH3

28

31

COC
COC
COC
C
C O
C

32

29
30

33
34

O
C O

O
CCH2N
C
C
N

577

578

4
1.37

5
3

6
6

7
77.4

8
24.107

9
4208

1.48

1.37

77.4

24.107

4208

1.37

1.48; 1.54

11

14

28.4

8.231

5875

1.37

1.37

80.6

24.107

4208

1.37

1.37

80.6

24.107

4208

1.37

1.37

79.2

24.107

4208

1.37

1.37

79.2

24.107

4208

1.37

1.37

11

11

8.5

3.602

1253

1.76

1.48

76.8

24.107

4208

36

C
37

3
1.48

NHC

O
38

39

O
40

N
N

41

N
42

CNHC

43

578

1
35

579

1
44

S
45

S
S

S
S

52

6
3

7
76.8

8
24.107

9
4208

1.76

1.76

78.0

24.107

4208

1.76

1.76

78.0

24.107

4208

1.76

1.37

78.6

24.107

4208

1.76

1.37

78.6

24.107

4208

1.76

1.37

80.0

24.107

4208

1.76

1.37

80.0

24.107

4208

1.76

1.76

12

12

16.5

8.00

1750

1.76

1.76

12

12

26.1

8.870

3656

50

51

5
7

48

49

4
1.48

46

47

3
1.76

O
CSC
O
C

579

580

54

N
C

6
11

7
144.6

8
46.103

9
12542

1.37

1.37

11

11

190.2

61.441

15553

1.37

1.37

11

113.4

35.105

8375

1.48

1.37

11

110.1

35.105

8375

1.37

1.48

103.4

33.126

8791

O
O

5
11

O
55

4
1.37

3
1.37

C
N

C
O
O

56

C
N

C
57

O
NH

C
N

580

1
53

581

1
58

3
1.76

4
1.37

5
7

6
11

7
111.3

8
35.105

9
8375

1.76

1.48

101.3

33.127

8791

1.64

1.64

13

13

71.6

1.64

1.64

10

10

0.5

1.50

1.50

24.8

4.629

2316

1.37

1.37

11

11

25.6

8.558

1899

1.37

1.37

124.8

39.445

7219

1.48

1.48

121.2

39.445

7219

1.48; 1.54

1.48; 1.54

27.1

4.898

599

N
C
O
NH

59

N
CH3

60

Si

CH3
SiOSi
O
C
O

61
62

63

O
N

CH3
64

65

CCF2C

581

66

582

As an example illustrating building of the repeat unit, Figure 105 shows a


combinatorial tree, which is composed from three fragments (basic fragments) with
various end labels. The rule, according to which the formation of a bond between end
groups is permitted, or prohibited, is reflected in the matrix of their interactions.
Clearly, the quantity of probable variants of the repeat unit of the polymer increases
with every step of joining a consequent fragment.
It should be taken into account that if all basic fragments recorded by the
computer memory participate in the composition and therewith each of them may join
to the propagating unit by both one or another ends, then 2n of variants of the repeat
unit (where n is the number of basic fragments) appear on every stage of joining of
the next fragment. Hence, this procedure represents a geometric progression with
denominator 2.
Besides direct and reverse tasks, software for prognosis of the physical
properties of polymers enables other tasks to be solved. For example, the upper and
lower ranges of the physical characteristics may be determined which, in principle,
organic polymers are able to display; for example, the lower value of the onset
temperature of intense thermal degradation of organic polymers, Td, is 150C, and the
upper limit is 575C. Hence, an organic polymer possessing the onset temperature of
intense thermal degradation above 575C is difficult to produce. Combination of
various properties in the same polymer may also be estimated, etc. (see below).
Software based on using the finest basic fragments enables calculations and
synthesis of polymers of the following classes:
1. Polyolefins, vinyl polymers, etc.
2. Polyethers and polyetherketones
3. Polyesters
4. Polyamides
5. Polysulfones, polyethersulfones
6. Polyimides
7. Polyoxadiazoles
8. Polybenzimidazoles
9. Polysulfides
10. Organosilicon polymers
11. Polycarbonates
The number of polymer classes may be increased, if necessary. This requires
introduction of a group determining affiliation of polymers to the chosen class into the
database; for example, introduction of urethane group into the database enables to
calculate properties of polyurethanes and to synthesize polyurethanes with the given
properties in a computer. The drawback of this software is that the above-mentioned
tasks may be solved only for polymers containing the basic fragments present in the
database. Although the number of polymers is great in this case, some limits do exist.
In this connection, A.A. Askadskii and A.F. Klinskikh [22] created software,
according to which the chemical structure of a polymer is gathered not from basic
fragments, but from separate atoms. In this case, the chemical structure of a repeat
unit is displayed on the screen in the form of a structural formula of an organic
compound, as a chemist writes it down on paper. After that, the physical properties of
the polymer of the given structure are calculated and laid out immediately to the user.
Therewith, the molecular design of a polymer may be performed by changing its
chemical structure, introducing various groups, etc., and its physical characteristics
are obtained immediately. This program enables us to solve the first (direct) task only,
because when polymers are synthesized from separate atoms (but not basic

583

fragments), the number of variants is so high that they cannot be realized on a


computer of any power.
This program enables to calculate the properties of polymers and copolymers,
and polymer networks, listed below:
1. Molecular mass of the repeat unit;
2. Molar volume at room temperature;
3. Temperature dependence of the molar volume;
4. Dependence of molar volume on crystallinity degree;
5. The Van-der-Waals volume;
6. Density at room temperature;
7. Dependence of density on crystallinity degree;
8. Temperature dependence of density;
9. Thermal expansion value in the given temperature range;
10. The Hildebrand parameter of solubility;
11. Surface energy;
12. Refractive index at room temperature;
13. Temperature dependence of refractive index;
14. Dielectric constant at room temperature;
15. Birefringence;
16. Dipole moment;
17. Stress-optical coefficient in the glassy state;
18. Glass transition temperature;
19. Dependence of glass transition temperature on molecular mass;
20. Temperature of the onset of intense thermal degradation;
21. Dependence of temperature of polymer transition to the viscous-flow state on
molecular mass;
22. Coefficient of volumetric expansion in the glassy state;
23. Coefficient of volumetric expansion in the rubbery state;
24. Molar heat capacity under constant pressure in the glassy state at room
temperature;
25. Molar heat capacity under constant pressure in the rubbery state at room
temperature;
26. Temperature dependence of molar heat capacity in the glassy state and the rubbery
state;
27. Specific heat capacity at room temperature under constant pressure in the glassy
state;
28. Specific heat capacity at room temperature under constant pressure in the rubbery
state;
29. Jump of molar heat capacity at transition from the glassy state to the rubbery state;
30. Jump of specific heat capacity at transition from the glassy state to the rubbery
state;
31. Total energy of intermolecular interaction;
32. Energy of dispersion interaction;
33. Energy of strong dipoledipole interaction and hydrogen bonds;
34. The part of energy of dispersion interaction in total energy of intermolecular
interaction;
35. The part of energy of dipoledipole interaction and hydrogen bonds in total
energy of intermolecular interaction;
36. Molar refraction;
37. Molar polarizability;

584

38. Total cohesive energy;


39. The part of cohesive energy stipulated by hydrogen bonds;
40. The part of cohesive energy stipulated by dipoledipole interaction;
41. The part of cohesive energy stipulated by dispersion interaction;
42. Permeability by oxygen;
43. Permeability by carbon dioxide;
44. Permeability by nitrogen;
45. Interface tension on the polymersolvent interface;
46. Interface tension on the polymerpolymer interface;
47. Elasticity modulus in the area of rubber-like plateau;
48. Value of mechanical segment (molecular mass or polymerization degree, at which
the rubbery state is displayed);
49. Activation energy of low-temperature -transition;
50. Newtonian viscosity of polymeric melts;
51. Prediction of solubility of polymers in organic solvents;
52. Influence of molecular mass on solubility;
53. Influence of orientation degree on solubility;
54. Prediction of miscibility of polymers.
For low-molecular liquids, calculation of the properties listed below is
involved in the program:
1. Molecular mass;
2. Molar volume;
3. The Van-der-Waals volume;
4. Density;
5. The Hildebrand parameter of solubility;
6. Surface tension;
7. Refractive index;
8. Dielectric constant;
9. Molar heat capacity in the solid state;
10. Molar heat capacity in the liquid state;
11. Molar refraction;
12. Molar polarizability;
13. Total cohesive energy;
14. The part of cohesive energy stipulated by hydrogen bonds;
15. The part of cohesive energy stipulated by dipoledipole interaction;
16. The part of cohesive energy stipulated by dispersion interaction.
Let us now turn to the principles of computer realization of the method of
atomic constants. Preliminarily, it should be noted that practical validity of this
method for solving the problem of searching for structures of the repeat unit providing
required physicochemical properties of polymer is obvious. For example, in the
frames of the method of constants, the values of typical temperatures (glass transition,
degradation and melting) of polymers may be calculated and structures fulfilling the
requirements on the total complex of the above-mentioned properties may be found.
The structures found may form the base for applying software of computer planning
of organic synthesis (CPOS).
Let us consider in greater detail the calculation scheme of the method of
constants on the example of calculation of the glass transition temperature of polymer
Tg by the chemical formula of its repeat unit. Tg is calculated from equation (IV.39).
The algorithm involves: 1) analysis of the structure by atoms and bonds; 2)
separation of groups of atoms responsible for dipoledipole interaction and hydrogen

585

bonds (including various cases of hydrogen bonding typical of polyamides); 3)


determination of ortho-, metha- and para-substitutions for aromatic cycles in the main
chain; 4) determination of double carboncarbon bonds in the main chain. As these
rules are not in contradiction with each other, they assume a formalized description
providing computer realization of the algorithm. Analogous analysis is performed in
calculation of all other properties of polymers, networks and organic liquids.
Computerization of the method of atomic constants is based on modularity,
developed, compatible Windows-interface and conformity with CPOS programs. First
two principles are well known and common. That is why let us dwell on the third
principle. Being the original object for CPOS application, the structure of the target
compound such be such as to ensure the required physico-chemical properties.
Because searching for such structure is one of the tasks in the method of constants,
coordinated description of structures in the method of constants and CPOS must be
provided. Main blocks for CPOS programs are tables of atoms and the table of
connectivity, which sets a selection of atoms and types of bonds in structure of the
compound. For realization of the method of constants, it has been assumed desirable
to form such tables for each calculated formula of the repeat unit. Of special attention
was solution of the problem of user communication with the program. To depict the
structure of the repeat unit of polymer, the planar (two-dimensional) image of
disposition of atoms linked by definite types of bonds was used. If necessary,
correspondence to linear Wiswesser formulae or to the writing form by Morgan is set
with the help of tables of atoms and connectivity. The problem of the unambiguous
interpretation of the structure of the repeat unit displayed on the screen is also
important. For this purpose, control of the existence of data on atoms, groups of atoms
and bonds in the database is used in the program.
The program represents an assembly of modules providing: 1) the regime of
editing the structure analyzed (entering, deleting and transposition of symbols of
atoms, groups of atoms, data files recording on disk and reading from disk with the
structure, etc.); 2) Analysis of the structure, the calculation of parameters and
treatment of structural errors. Separation of modules of the second group enables to
make the program easily transmitted between various systems. In practice, the
program is used since 1994, first applied to DOS, and at present to Windows98.

Figure 106. Schematic representation of the region of possible existence for two properties of
polymers (see text).

586

The program was tested on all compounds for which, on the one hand,
experimental data were known and, on the other hand, calculations by the method of
constants had been performed before. For example, the properties of polyolefins,
vinyl, acrylic and dienic polymers, aliphatic and aromatic polymers (including
polyamides), polyesters, polyimides, polysulfones, polyetherketones, etc. were
calculated. It has been shown that the program successfully interprets the structure of
the repeat unit in accordance with the calculation scheme of the method of constants,
and provides for required accuracy of the calculation method of 35%. Examples of
solving the direct, reverse and combined tasks of computerized synthesis of polymers
based on the above-described programs are indicated in Appendices 1 3.
Let us consider one more task, which may be solved by the above-described
software.
At present, in connection with the application of polymers to various spheres,
requirements on their physical properties have become much stricter. Problems
appear, solutions of which require polymers with a combination of a series of physical
properties, simultaneously falling within the given ranges of the values of their
characteristics. A question is raised about the combination of properties of polymers.
Described in ref. [23] is the method of its solving with the help of miscibility
diagrams, which consists of the following: assume that for a large group of polymers
values of parameters of any two physical properties (from experiment or calculation)
are known. Let us compose a diagram on which a specific point (Figure 106)
corresponds to every polymer, where the abscissa axis (X) represents the values of the
parameters of the first property, and on the ordinate axis (Y) of the second one. All
these points are located inside some region, which may be contoured. This region is
called the miscibility region of the mentioned physical properties for the given
group of polymers. Attention should be paid to the fact that the density of points
distribution in the region may be rather irregular. Such diagrams enable to estimate
possibility of existence of polymers with given properties without additional
consumption of labor. Actually, let the ranges (X2 X1) and (Y2 Y1) of values of
properties, which must be possessed by polymers, be given. If the area determined by
these ranges, falls within the miscibility region in the diagram, and some points exist
there, i.e. the polymer may exist (case I). In the opposite case, obtaining the required
polymer among polymers of these classes is of low probability (case II).
To compose miscibility diagrams of this kind, either experimental data on
various physical properties of polymers, or calculated data may be used.
Combinations of these data may also be used.
To calculate various physical properties of polymers, it is desirable to apply
the computer program suggested in ref. [17], because with the help of it the properties
of polymers either already synthesized, or the ones which may be, in fact, synthesized
are considered. This very program was used in ref. [23] for composing combination
regions of various properties of polymers. Among these properties the following were
considered: density, ; refractive index, n; glass transition temperature, Tg; the
relation of glass transition temperature to the melting point, Tg/Tm; the onset
temperature of intense thermal degradation, Td; the Hildebrand solubility parameter,
(density of cohesive energy); stress-optical coefficient C.
Ref. [23] indicates 27 diagrams built on a computer, which reflect
combination of any two of the above-mentioned physical properties. Formulae, with
the help of which the connection between two given parameters of properties may be
searched, are shown in monograph [6].

587

Figure 107. Combination diagrams of Td and Tg (a), and Tg (b), and n (c) (see text).

As an example, Figure 107 shows miscibility diagrams of the following


properties of polymers: glass transition temperature, Tg, and the onset temperature of
intense thermal degradation, Td; Tg and solubility parameter, ; refractive index, n,
and . Let us consider several features of these and other similar diagrams.
All regions of miscibility display different density. This means that there are
sub-regions of miscibility exist containing the main bulk of polymers, and from these
very sub-regions the chemical structure of polymers may be most easily chosen,
which would fit the two given properties.
All the diagrams display the presence of clearly expressed dependences of one
property on another. Each of these dependencies characterizes definite series of
polymers, for example, polyesters, polyamides, polyimides, etc.

588

Every diagram contains a point from which these dependences start. This is
associated with the program feature [17], in which several homologues (CH2)n up
to n = 20 are set as original fragments from the total 194 ones, and the point, from
which these dependences start, characterizes polyethylene.
Other features inherent to each diagram are clearly seen in corresponding
figures. Having these diagrams, the possibility of obtaining polymers, which would
possess the necessary combination of one of several properties may be predicted. For
example, if polymers with the solubility parameter = 10 (cal/cm3)1/2 and Tg 300C
should be obtained, it may be made easily, because the point corresponded to these
coordinates falls within the densest part of the diagram in Figure 107b. It would be
quite difficult to produce a polymer with the glass transition temperature Tg ~ 500C
at the same solubility parameter, and at Tg = 600C this becomes practically
impossible, because the point corresponding to these coordinates is out of the range of
the miscibility region. This analysis may be easily performed for any diagram
presented in ref. [23], as well as for a set of these diagrams, which enables us to
predict the possibility of obtaining polymers with a a set of the required properties.
Clearly, that if such diagrams are composed with the help of a computer program,
according to which the polymer is composed from the finest basic fragments, the
regions of miscibility of polymers will be significantly expanded by points reflecting
properties of the very large number of polymers.

Appendix 1. Examples of solution of direct problems of


polymer synthesis

Let us consider solution of the direct problem of estimation of the properties


of polymers based on their chemical structure on an example of natural polymers. The
specific features of these polymers are the following:
1) These polymers contain cyclic non-aromatic structures with a large number of
OH-groups able to form hydrogen bonds.
2) Hydrogen bonds in these polymers may be formed both between neighboring
chains and in the range of a single chain, and even in a single repeat unit. The
scheme representing these interactions is shown in Figure 41. The presence of
interchain and intrachain (including intra-unit) hydrogen bonds depends on the
steric factor dislocation of OH-groups in relation to the plane of circles.
3) Many high-molecular compounds, including cellulose, degrade thermally at
heating before the process of devitrification is onset. Consequently, the glass
transition temperature (the most important characteristic for polymers) may not be
determined experimentally.
4) Many natural high-molecular compounds possess different units, i.e. contain
fragments of different chemical nature in their chains.
In this Section, we will discuss application of the approach developed for
estimating the physical properties of polymers based on their chemical structure to
natural high-molecular compounds, which possess quite complicated but well studied
chemical structure.
Table A-1-1 indicates chemical structures of 16 fragments of hemicelluloses
structures, most often met in the structure of natural polymers. Of special attention is
the fact that in some cases, despite the identical gross-formula, these fragments
included in polymeric chains involve different physical properties of natural
polymers. For example, for structures Fs1 and Fs2, gross-formulae are identical, but in
the structure of Fs2 fragment OH-groups disposed at one side of the circle plane may
form hydrogen bonds between each other. Hence, they participate in the formation of
bonds between neighboring chains. This circumstance influences the glass transition
temperature and some other properties of polymers containing these fragments.
Let consider in greater detail the influence of the chemical structure of the
above-mentioned natural polymers on their glass transition temperature. The glass
transition temperature is calculated from equation (IV.39). Calculating the glass
transition temperature from this equation, all fine features of the structure of the
polymeric unit should be taken into account. If hydrogen bonds are formed between
chains of the polymer (as a result, a physical network is formed at the sacrifice of
intermolecular bonds), then parameter bh = 140103 3K1 should be introduced into
equation (IV.39). In the case if OH-groups in the repeat unit are disposed in the
manner that hydrogen bonds are formed between them in the range of the same unit,
then the network of physical intermolecular bonds between chains is not formed (see
Figure A-3-1,b), and for other groups, constants bd instead of bh must be introduced
into equation (IV.39), which are introduced for every branching from the main chain
of the polymer.
The use of this approximation leads to values of the glass transition
temperature and all other properties, which are displayed in Table A-1-2. It is clearly

590

observed that at the identical gross-formula the glass transition temperature of the
structure Fs1 (541 K) is much higher than for the structure Fs2 (319 K). Therewith,
such important characteristic of polymers as Mc (molecular mass of the mechanical
segment of macromolecule, at which the rubbery state appears) is also changed.
Saturation of the structure of hemicelluloses by OH- and COOH-groups leads to a
sharp increase of the glass transition temperature, if these groups are capable of
forming hydrogen bonds between chains of polymers. If hydrogen bonds are formed
inside the repeat unit, the glass transition temperature is abruptly decreased.
Let us consider in brief some other properties. The densities of solid samples
are approximately equal, the lowest density being possessed by the structure Fs5, in
which only a single OH-group is present. Consequently, this structure possesses the
highest onset temperature of intense thermal degradation. It should be noted that, in
some cases, the onset temperature of intense thermal degradation lies below the glass
transition temperature or is close to it, which is the reason that experimental
determination of Tg for natural polymers is rather difficult.
The surface energy for 16 calculated structures changes in quite wide ranges,
the lowest surface energy being typical of the structure Fs5 containing the smallest
number of OH-groups, and the highest value is typical of the structure Fs8 containing
the greatest number of these groups.
Dielectric constants are approximately equal, the solubility parameter is quite
high compared with synthetic polymers, refractive indices are approximately equal,
and the temperature coefficient of volumetric expansion in the glassy state
significantly depends upon the chemical structure of the fragment and changes
antibatically to the glass transition temperature.
Further on, let us consider the properties of hemicelluloses composed from the
above-mentioned fragments. The chemical composition of 8 hemicelluloses and
disposition of fragments in them are shown in Table A-1-3. The main properties of
hemicelluloses of the given composition are shown in Table A-1-4.
All properties were calculated on the basis of the molar content of fragments.
Actually, the glass transition temperature of hemicelluloses of various chemical
structures varies from 332 to 517 K, and the onset temperature of intense thermal
degradation from 517 to 580 K.
The main characteristics indicated in Table A-1-4 show that the values of
these temperatures for hemicelluloses C4 are very similar. Of special attention is the
extremely low permeability of films from hemicelluloses to various gases.
Experimental testing of the calculation results was performed on the basis of
1,4--D-glucomannan, the structure of which (F8) is shown on Table A-1-3. The
thermomechanical curve for the initial preparation containing an uncontrolled amount
of water is shown in Figure A-1-1. After drying in vacuum at a temperature of 80C
during two hours, the form of the thermomechanical curve changed a little (Figure A1-2). The form of Debye patterns (Figure A-1-3) also changes during such thermal
processing in vacuum. They indicate that the glucomannan specimens studied were
practically amorphous. The calculated value of the glass transition temperature (356
K) with the accuracy, typical of such calculations, coincides with the experimental
value which, according to the thermomechanical curve (see Figure A-1-2), is 333 K.
It is not inconcievable that after further drying the glass transition temperature
might be somewhat increased. The second peak of deformation on the
thermomechanical curve appears at a temperature close to the temperature of intense
thermal degradation and is probably associated with it.

591

Needless to say that the results of all the calculations performed may be
considered only as preliminary ones, and in future they must be related to not only
with the primary, but also with the secondary structure of chemicelluoses.
Table A-1-1
Chemical formulae of the fragments of hemicellulose structures
Designation of structures
Chemical formulae of structures fragments
fragments
1
2
CH2OH
Fs1
O
O

OH
OH

Fs2

CH2OH
O
O

OH

HO

Fs3
O
O

OH
OH

Fs4

COOCH3
O
O
OH
OH

Fs5
O
O

COOCH3
OH

Fs6
O
O

OH
O
OH
HO
OH

592

1
Fs7

2
COOH
O
O

OH
O

OH
HO
OH

Fs8
O
O

OH

OH

OH
OH

Fs9
O
O

OH
O

COOCH3
HO
COOCH3

Fs10
O
O

OH

COOH
O

O
HO
OCH3
OH

593

1
Fs11

2
OH
HO
OH
O

CH2OH
CH2
O
O

OH

HO

CH2OH

Fs12

O
O

OH

OH
OH

OH
O

CH2OH

Fs13
O
O

OH

OH

H3CO
OH
O
COOH

Fs14

CH2OH
O

HO
OH

OH O
CH2
HO

HO

O
OH

CH2OH
O

OH

594

1
Fs15

2
O
O

O
OH
O

OCH3
OH

OH
OH

O
OH

Fs16

CH2OH

CH2OH
O

HO

COOH

HO

OH

OH
OH

OH
OH

O
CH2

HO

OH

OH

HO

O
CH2

CH2
HO

O
OH

OH

OH

Table A-1-2
Properties of hemicellulose structure fragments
Values of properties of structures fragments
Designation of structures
Vi , 3 , 3 Tg, K Td, K
properties
MM
Vm, cm3/mol
g/cm
i
Fs1
162
116
132
1.39
541
556
Fs2
162
116
132
1.39
319
556
Fs3
132
95
107
1.40
511
571
Fs4
190
135
152
1.41
459
585
Fs5
174
128
145
1.36
403
607
Fs6
264
189
214
1.40
574
527
Fs7
308
212
240
1.45
641
520
Fs8
250
174
197
1.44
427
520
Fs9
348
256
290
1.36
436
570
Fs10
322
229
259
1.41
518
530
Fs11
324
233
263
1.39
431
511
Fs12
324
233
263
1.39
432
511
Fs13
322
229
259
1.41
518
530
Fs14
486
349
395
1.39
388
525
Fs15
454
324
367
1.40
552
517
Fs16
929
661
748
1.41
437
513
Values of properties of structures fragments
Designation of
E*,
E,
,
h,
dd,
d,
,
structures properties
kcal/mol rel. un. rel. un. rel. un. rel. un. kcal/mol
dyn/cm
Fs1
45.5
23.9
0.72
0.00
0.28
3.36
49.6
Fs2
45.5
23.9
0.72
0.00
0.28
3.36
49.6
Fs3
54.8
18.2
0.63
0.07
0.30
3.34
37.96

595

Fs4
Fs5
Fs6
Fs7
Fs8
Fs9
Fs10
Fs11
Fs12
Fs13
Fs14
Fs15
Fs16
Designation of
structures
properties
Fs1
Fs2
Fs3
Fs4
Fs5
Fs6
Fs7
Fs8
Fs9
Fs10
Fs11
Fs12
Fs13
Fs14
Fs15
Fs16

46.4
37.9
53.0
58.9
62.1
36.5
49.9
45.5
45.5
49.9
45.5
50.3
46.6
Edd+h,
kcal/mol
37.60
22.20
27.70
30.18
21.53
64.07
80.77
39.00
48.40
63.65
84.83
84.83
63.65
70.17
99.57
157.40

Designation
of structures
properties

L104,
K1

Fs1
Fs2
Fs3
Fs4
Fs5
Fs6
Fs7
Fs8
Fs9
Fs10
Fs11
Fs12
Fs13
Fs14
Fs15
Fs16

4.46
4.46
4.71
5.25
5.97
4.20
3.76
3.59
5.52
4.65
4.04
4.04
4.65
4.18
4.37
3.90

Designation of
structures
properties

P,
cm3/mol

Fs1
Fs2

51.2
51.2

21.5
0.54
0.11
0.35
3.49
16.8
0.34
0.21
0.44
3.33
35.2
0.66
0.03
0.31
3.34
43.3
0.67
0.05
0.28
3.56
37.8
0.61
0.13
0.26
3.45
32.3
0.36
0.18
0.46
3.33
39.7
0.58
0.09
0.33
3.43
47.8
0.72
0.00
0.28
3.36
47.8
0.72
0.00
0.28
3.36
39.7
0.58
0.09
0.33
3.43
71.7
0.72
0.00
0.28
3.36
56.7
0.61
0.06
0.33
3.40
139.1
0.70
0.03
0.27
3.38
Values of properties of structures fragments
Ed,
n,
U,
,
kcal/mol
rel. un.
kcal/mol
(cal/cm3)1/2
12.0
14.30
1.49
27.9
12.0
14.30
1.49
27.9
10.26
13.85
1.49
26.4
15.97
12.65
1.48
23.7
17.23
11.43
1.48
20.8
21.15
13.63
1.49
29.7
22.83
14.29
1.48
33.1
18.74
14.72
1.48
34.7
35.45
11.23
1.48
22.6
26.87
13.16
1.48
26.8
24.60
14.34
1.49
30.8
24.60
14.34
1.49
30.8
26.87
13.16
1.48
26.8
36.61
14.34
1.49
29.8
37.76
13.23
1.48
28.5
69.68
14.51
1.49
31.9
Values of properties of structures fragments
n l,
R,
C ps ,
C pl ,
rel. un.
cm3/mol
cal/(moldeg)
cal/(moldeg)
467
52.7
67.6
33.37
487
52.7
67.6
33.37
431
41.5
54.0
27.23
435
57.6
76.1
38.12
376
52.3
69.5
36.59
609
83.8
107.3
54.45
707
95.7
123.7
60.61
693
77.9
100.3
49.83
512
105.4
138.3
73.18
586
99.9
129.4
65.30
678
106.1
134.5
66.74
678
106.1
134.5
66.74
586
99.9
129.4
65.30
750
159.5
201.3
100.11
700
142.1
182.7
92.57
995
302.7
381.4
189.45
Values of properties of structures fragments
C,
,
PO 2 ,
PCO 2 ,
B
D
DU
DU
104
104

1.55
1.55

0.03
0.03

0.07
0.07

46.15
38.76
85.22
103.60
90.93
83.85
90.52
109.43
109.43
90.52
158.65
137.33
321.32

G104,
K4
1.78
3.01
1.88
2.09
2.38
1.67
1.50
1.43
2.20
1.85
2.22
2.22
1.85
1.66
1.74
1.55
Ml
79034
46602
56980
82740
65434
160900
217850
110198
178375
188853
219817
219817
188853
245324
318105
653567
PN 2 ,
DU
0.00
0.00

596

Fs3
Fs4
Fs5
Fs6
Fs7
Fs8
Fs9
Fs10
Fs11
Fs12
Fs13
Fs14
Fs15
Fs16

41.4
61.1
56.0
82.9
97.8
78.3
112.0
102.5
102.3
102.3
102.5
153.5
144.0
292.6

132
91
104
108
99
113
95
108
95
95
108
98
101
96

1.39
1.70
1.63
1.99
2.16
1.93
2.32
2.21
2.21
2.21
2.21
2.72
2.63
3.76

Note: MM is the molecular mass; Vm is the molar volume;

0.04
0.22
1.38
0.11
0.02
0.00
3.17
0.11
0.03
0.03
0.11
0.03
0.15
0.01

V i

0.09
0.56
4.12
0.27
0.03
0.00
10.23
0.26
0.07
0.07
0.26
0.07
0.37
0.03

0.01
0.03
0.24
0.02
0.00
0.00
0.60
0.02
0.00
0.00
0.02
0.00
0.02
0.00

is the Van-der-Waals volume; is

the density; Tg is the glass transition temperature; Td is the onset temperature of intense thermal
degradation; is the surface energy; E* is the cohesive energy; h is the relation of hydrogen bonds
energy to total cohesive energy; dd is the relation of dipoledipole interaction energy to total cohesive
energy; d is the relation of dispersion interaction energy to total cohesive energy; is the dielectric
constant; E is the total energy of intermolecular interaction; Edd+h is the energy of dipoledipole
interaction and hydrogen bonds; Ed is the energy of dispersion interaction; is the solubility parameter;
n is the refractive index; U is the activation energy of low-temperature -transition; G is the
coefficient of thermal expansion in the glassy state; L is the coefficient of thermal expansion in the
rubbery state; nl is the polymerization degree of the polymer when the rubbery state appears; C ps is the
molar heat capacity in the glassy state; C pl is the molar heat capacity in the rubbery state; R is the
molar refraction; Ml is the molecular mass of polymer when the rubbery state appears; P is the
polarizability; C is the stress-optical coefficient; is the dipole moment; PO 2 , PCO 2 , PN 2 are the
permeabilities by oxygen, carbon dioxide and nitrogen, respectively (the unit of measurement of
permeability 1 DU = 0.451010 cm2s1atm1).















Figure A-1-1. Thermomechanical curve of initial 1,4--D-glucomannan.

597

Table A-1-3
Designations of
hemicelluloses
1
C1

Fragmentary composition and structural formulae of hemicelluloses


Fragmentary composition and name Structural formulae of hemicelluloses
of natural biopolymer
2
3
(Fs12)3(Fs2)42(Fs1)23(Fs11)5
aspen glucomannan [208]

OH
OH
OH
O

CH 2O H
CH 2O H

CH2 O H
O

OH

CH2

CH2 O H
O
OH

OH

O
O

OH

42

OH

O
OH

OH

23

OH
OH

OH
O

CH2 O H

C2

(Fs12)2(Fs2)13(Fs1)5(Fs12)3
spruce glucomannan [116]

CH2 O H

CH 2O H
O

O
OH

OH

CH 2OH

CH2 O H
O

OH

O
13

OH

OH

O
OH

OH
OH

OH

OH

OH

O
2

CH2 O H

597

CH2 O H

OH

598

2
(Fs10)19(Fs3)125(Fs6)5
aspen 4-0-methylglucuronoxylan
[117]

598

1
C3

OH

OH

O
OH

OH

125

COOH
O

OH

OH

OH

OCH3
OH
C4

OH

19

Fs3Fs10Fs3Fs6Fs8Fs15
spruce arabinoglucuronoxylan [118]
O
O

O
O

OH

O
O

OH

OH

OH

OH
OH

O
O

OH

OCH3

O
OH

OH

OH
OH

OH

OH

OCH3

COOH

OH

OH

OH

OH
OH

OH

O
OH

COOH

599

1
C5

2
Fs7(Fs4)50Fs7
Galactouronan from plants [130]

COOH

COOCH3

COOH

O
OH

OH

O
OH

OH

50

O
OH

OH

OH

OH

OH

OH
C6

(Fs9)8(Fs5)21(Fs3)52(Fs13)15(Fs5)3
birch xylan [118]
O

O
O
OH

CO OCH 3
OH

O
21

OH
OH

OH
COOCH3 3

52
OH

COO CH3

OH

H3CO
OH

OH
O
COO CH3 8

COOH

15

599

600

2
Fs16Fs14
larch arabinogalactan [60]

3
CH2OH

CH2OH
O

OH

CH2OH
O

OH

OH

OH

CH2OH

CH2OH

OH

OH

O
OH

O
OH

OH

OH

CH2OH
O

OH

CH2OH
O

OH

OH

CH2OH
O

O
OH

OH

OH

CH2OH
O

OH

CH2OH
O

CH2OH

OH

OH

CH2OH
O

O
CH2

OH

OH

O
CH2

OH

[Fs2Fs1(Fs2)4Fs1Fs2]48
1,4--D-glucomannan [2]

OH

OH

OH

C8

OH

CH2

OH

OH

CH2
OH

OH

OH

OH

600

1
C7

OH

OH

O
O

OH

O
OH

OH

OH

48

Note: chemical formulae of hemicelluloses and the sample of 1,4--D-glucomannan shown in the Table were presented by Doctors of Biology V.D. Shcherbukhin and A.V.
Bolobova.

601

Table A-1-4
Properties of hemicelluloses
Properties of hemicelluloses
,
Tg,
Td,
,
G104,
(cal/cm3)
K
K
dyn/cm
K-1
1/2

Designations of
PO 2 ,
PCO 2 ,
Vi ,
,
hemiceli
3
DU
DU
g/cm
luloses
3
C1
146
1.39
332 546
2.90
14.3
45.5
0.03
0.07
C2
160
1.39
367 538
2.62
14.3
45.5
0.03
0.07
C3
130
1.40
499 557
1.92
13.7
53.4
0.05
0.12
C4
211
1.41
517 530
1.83
13.6
53.3
0.05
0.11
C5
156
1.42
464 580
2.07
12.8
47.2
0.19
0.47
C6
154
1.38
435 566
2.21
12.8
47.0
0.21
0.54
C7
571
1.40
419 517
1.62
14.5
46.2
0.02
0.04
C8
1,052
1.39
356 556
2.70
14.3
45.5
0.03
0.07
Note: Designations of properties of hemicelluloses correspond to those used in Table A-1-2.

PN 2 ,
DU
0.00
0.00
0.01
0.01
0.03
0.03
0.00
0.00

Figure A-1-2. Thermomechanical curve of dry 1,4--D-glucomannan.

Figure A-1-3. Debye patterns of powder-like initial (a) and dry (b) specimens of 1,4--D glucomannan

Appendix 2. Examples of solving the reverse problem of


polymer synthesis

The reverse task concludes in synthesis, i.e. in finding a chemical formula or


formulae of polymers whose properties must fall within the given range. Let
poly(ether ketones) be required to be synthesized possessing the glass transition
temperature above 493 K and simultaneously with the melting point below 653 K (the
latter is important for successful processing). There are no limits on the values of the
rest of properties (solubility parameter, density, dielectric constant, refractive index).
The given task will be solved with the help of the program operating with the
finest basic fragments. Let us select the number of basic fragments in repeat units,
equal to 7. The solution of the given task yields the chemical formulae of poly(ether
ketones), shown in Table A-2-1 (450 000 structures were calculated, from which only
24 structures fell within the required range of properties).
Let us widen a little the range of properties by introduction of the lower border
for the glass transition temperature, equal to 483 K. In this case, from the same
number of calculated structures 84 structures of poly(ether ketones) were selected, the
chemical formulae of which are shown in Table A-2-2.

603

Table A-2-1
Chemical formulae of poly(ether ketones) with the glass transition temperature Tg above 493 K and the melting point Tm below 653 K
Properties of poly(ether ketones)
Numbers of poly(ether ketones)
Chemical formulae of poly(ether ketones)
n,
Tg,
Tm,
,
,
,
structures
rel. un.
K
K
(cal/cm3)1/2
g/cm3
rel. un.
1
2
3
4
5
6
7
8
1
495
632
9.78
1.24
2.95
1.62
COpC6H4OpC6H4phtC10H6mC6H4
2
495
632
9.78
1.24
2.95
1.62
COpC6H4OpC6H4phtmC6H4ClOH6
3
498
652
9.63
1.20
2.83
1.59
COpC6H4OpC6H4flClOH6mC6H4
4
498
652
9.63
1.20
2.83
1.59
COpC6H4OpC6H4flmC6H4C10H6
5
495
632
9.78
1.24
2.95
1.62
COpC6H4OpC6H4C10H6phtmC6H4
6
498
652
9.63
1.20
2.83
1.59
COpC6H4OpC6H4C10H6flmC6H4
7
495
632
9.78
1.24
2.95
1.62
COpC6H4OpC6H4mC6H4phtC10H6
8
498
652
9.63
1.20
2.83
1.59
COpC6H4OpC6H4mC6H4flC10H6
9
495
632
9.78
1.24
2.95
1.62
COC10H6phtpC6H4pC6H4mC6H4
10
495
632
9.78
1.24
2.95
1.62
COC10H6phtmC6H4pC6H4pC6H4
11
498
652
9.63
1.20
2.83
1.59
COC10H6flpC6H4pC6H4mC6H4
12
498
652
9.63
1.20
2.83
1.59
COC10H6flmC6H4pC6H4pC6H4
13
495
632
9.78
1.24
2.95
1.62
COC10H6pC6H4pC6H4phtmC6H4
14
498
652
9.63
1.20
2.83
1.59
COC10H6pC6H4pC6H4flmC6H4
15
495
632
9.78
1.24
2.95
1.62
COC10H6mC6H4phtpC6H4pC6H4
16
498
652
9.63
1.20
2.83
1.59
COC10H6mC6H4flpC6H4pC6H4
17
495
632
9.78
1.24
2.95
1.62
COmC6H4phtpC6H4pC6H4C10H6
18
495
632
9.78
1.24
2.95
1.62
COmC6H4phtC10H6pC6H4pC6H4
19
489
652
9.63
1.20
2.83
1.59
COmC6H4flpC6H4pC6H4C10H6
20
498
652
9.63
1.20
2.83
1.59
COmC6H4flC10H6pC6H4pC6H4
21
495
632
9.78
1.24
2.95
1.62
COmC6H4pC6H4pC6H4phtC10H6
22
498
652
9.63
1.20
2.83
1.59
COmC6H4pC6H4pC6H4flC10H6
23
495
632
9.78
1.24
2.95
1.62
COmC6H4C10H6phtpC6H4pC6H4
24
498
652
9.63
1.20
2.83
1.59
COmC6H4C10H6flpC6H4pC6H4
Note: Designations are the same as in the Note to Table A-2-2.
603

604

26

489

550

9.71

1.24

2.96

1.62

COpC6H4pC6H4mC6H4phtmC6H4

604

Table A-2-2
Chemical formulae of poly(ether ketones) with the glass transition temperature Tg above 483 K and the melting point Tm below 653 K
Properties of poly(ether ketones)
Numbers of poly(ether ketones)
Chemical formulae of poly(ether ketones)
n,
Tg,
Tm,
,
,
,
structures
rel. un.
K
K
(cal/cm 3)1/2
g/cm3
rel. un.
1
2
3
4
5
6
7
8
1
489
626
9.71
1.24
2.96
1.62
COmC6H4pC6H4phtmC6H4pC6H4
2
489
626
9.71
1.24
2.96
1.62
COmC6H4pC6H4phtpC6H4mC6H4
3
493
648
9.55
1.19
2.83
1.58
COmC6H4pC6H4flmC6H4pC6H4
4
493
648
9.55
1.19
2.83
1.58
COmC6H4pC6H4flmC6H4pC6H4
5
489
626
9.71
1.24
2.96
1.62
COmC6H4pC6H4mC6H4phtpC6H4
6
493
648
9.55
1.19
2.83
1.58
COmC6H4pC6H4mC6H4flpC6H4
7
489
626
9.71
1.24
2.96
1.62
COmC6H4pC6H4pC6H4phtmC6H4
8
493
648
9.55
1.19
2.83
1.58
COmC6H4pC6H4pC6H4flmC6H4
9
489
626
9.71
1.24
2.96
1.62
COpC6H4mC6H4phtmC6H4pC6H4
10
489
626
9.71
1.24
2.96
1.62
COpC6H4mC6H4phtpC6H4mC6H4
11
493
648
9.55
1.19
2.83
1.58
COpC6H4mC6H4flmC6H4pC6H4
12
493
648
9.55
1.19
2.83
1.58
COpC6H4mC6H4flpC6H4mC6H4
13
498
626
9.71
1.24
2.96
1.62
COpC6H4mC6H4mC6H4phtpC6H4
14
493
648
9.55
1.19
2.83
1.58
COpC6H4mC6H4mC6H4flpC6H4
15
489
626
9.71
1.24
2.96
1.62
COpC6H4mC6H4pC6H4phtmC6H4
16
493
648
9.55
1.19
2.83
1.58
COpC6H4mC6H4pC6H4flmC6H4
17
495
632
9.78
1.24
2.95
1.62
COpC6H4pC6H4phtC10H6mC6H4
18
495
632
9.78
1.24
2.95
1.62
COpC6H4pC6H4phtmC6H4C10H6
19
498
550
9.71
1.24
2.96
1.62
COpC6H4pC6H4phtmC6H4mC6H4
20
498
652
9.63
1.20
2.83
1.59
COpC6H4pC6H4flC10H6mC6H4
21
498
652
9.63
1.20
2.83
1.59
COpC6H4pC6H4flmC6H4C10H6
22
493
578
9.55
1.19
2.83
1.58
COpC6H4pC6H4flmC6H4mC6H4
23
495
632
9.78
1.24
2.95
1.62
COpC6H4pC6H4C10H6phbmC6H4
24
498
652
9.63
1.20
2.83
1.59
COpC6H4pC6H4C10H6flmC6H4
25
495
632
9.78
1.24
2.95
1.62
COpC6H4pC6H4mC6H4phtC10H6

605

2
498
493
495
495
498
498
495
498
495
498
489
489
495
480
495
489
489
489
493
493
498
493
498
493
493
493
489
493
489
493
495

3
652
578
632
632
652
652
632
652
632
652
626
626
632
550
632
550
626
626
648
648
652
578
652
578
648
648
626
648
626
648
632

4
9.63
9.55
9.78
9.78
9.63
9.63
9.78
9.63
9.78
6.63
9.71
9.71
9.78
9.71
9.78
9.71
9.71
9.71
9.55
9.55
9.63
9.55
9.63
9.55
9.55
9.55
9.71
9.55
9.71
9.55
9.78

5
1.20
1.19
1.24
1.24
1.20
1.20
1.24
1.20
1.24
1.20
1.24
1.24
1.24
1.24
1.24
1.24
1.24
1.24
1.19
1.19
1.20
1.19
1.20
1.19
1.19
1.19
1.24
1.19
1.24
1.19
1.24

6
2.83
2.83
2.95
2.95
2.83
2.83
2.95
2.83
2.95
2.83
2.96
2.96
2.95
2.96
2.95
2.96
2.96
2.96
2.83
2.83
2.83
2.83
2.83
2.83
2.83
2.83
2.96
2.83
2.96
2.83
2.95

7
1.59
1.58
1.62
1.62
1.59
1.59
1.62
1.59
1.62
1.59
1.62
1.62
1.62
1.62
1.62
1.62
1.62
1.62
1.58
1.58
1.59
1.58
1.59
1.58
1.58
1.58
1.62
1.58
1.62
1.58
1.62

8
COpC6H4pC6H4mC6H4flC10H6
COpC6H4pC6H4mC6H4flmC6H4
COC10H6phtpC6H4OpC6H4mC6H4
COC10H6phtmC6H4pC6H4OpC6H4
COC10H6flpC6H4pC6H4mC6H4
COC10H6flmC6H4pC6H4OpC6H4
COC10H6pC6H4pC6H4phtmC6H4
COC10H6pC6H4pC6H4flmC6H4
COC10H6mC6H4phtpC6H4OpC6H4
COC10H6mC6H4flpC6H4OpC6H4
COmC6H4phtmC6H4pC6H4pC6H4
COmC6H4phtpC6H4mC6H4pC6H4
COmC6H4phtpC6H4pC6H4C10H6
COmC6H4phtpC6H4pC6H4mC6H4
COmC6H4phtC10H6pC6H4pC6H4
COmC6H4phtmC6H4pC6H4pC6H4
COmC6H4phtpC6H4mC6H4pC6H4
COmC6H4phtpC6H4pC6H4mC6H4
COmC6H4flmC6H4pC6H4pC6H4
COmC6H4flpC6H4mC6H4pC6H4
COmC6H4flpC6H4pC6H4C10H6
COmC6H4flpC6H4pC6H4mC6H4
COmC6H4flC10H6pC6H4pC6H4
COmC6H4flmC6H4pC6H4pC6H4
COmC6H4flpC6H4mC6H4pC6H4
COmC6H4flpC6H4pC6H4mC6H4
COmC6H4mC6H4pC6H4phtpC6H4
COmC6H4mC6H4pC6H4flpC6H4
COmC6H4pC6H4mC6H4phtpC6H4
COmC6H4pC6H4mC6H4flpC6H4
COmC6H4pC6H4pC6H4phtC10H6

605

1
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57

606

pht

O,
C

8
COmC6H4pC6H4pC6H4phtmC6H4
COmC6H4pC6H4pC6H4flC10H6
COmC6H4pC6H4pC6H4flmC6H4
COmC6H4C10H6phtpC6H4OpC6H4
COmC6H4C10H6flpC6H4OpC6H4
COmC6H4mC6H4phtpC6H4OpC6H4
COmC6H4mC6H4flpC6H4OpC6H4
COmC6H4pC6H4phtmC6H4OpC6H4
COmC6H4pC6H4phtpC6H4OmC6H4
COmC6H4mC6H4flmC6H4OpC6H4
COmC6H4pC6H4flpC6H4OmC6H4
COpC6H4phtmC6H4pC6H4mC6H4
COpC6H4phtpC6H4mC6H4mC6H4
COpC6H4phtmC6H4mC6H4pC6H4
COpC6H4phtmC6H4pC6H4mC6H4
COpC6H4flmC6H4pC6H4mC6H4
COpC6H4flpC6H4mC6H4mC6H4
COpC6H4flmC6H4mC6H4pC6H4
COpC6H4flmC6H4pC6H4mC6H4
COpC6H4mC6H4pC6H4phtmC6H4
COpC6H4mC6H4pC6H4flmC6H4
COpC6H4pC6H4mC6H4phtmC6H4
COpC6H4pC6H4mC6H4flmC6H4
COpC6H4mC6H4phtmC6H4OpC6H4
COpC6H4mC6H4phtpC6H4OmC6H4
COpC6H4mC6H4flmC6H4pC6H4
COpC6H4mC6H4flpC6H4OmC6H4

fl

, mC6H4

, pC6H4

2) Tg is the glass transition temperature; Tm is the melting point; is the solubility parameter; is the density; is the dielectric constant; n is the refractive index.

606

1
2
3
4
5
6
7
58
489
550
9.71
1.24
2.96
1.62
59
498
652
9.63
1.20
2.83
1.59
60
493
578
9.55
1.19
2.83
1.58
61
495
632
9.78
1.24
2.95
1.62
62
498
652
9.63
1.20
2.83
1.59
63
489
550
9.71
1.24
2.96
1.62
64
493
578
9.55
1.19
2.83
1.58
65
489
626
9.71
1.24
2.96
1.62
66
489
626
9.71
1.24
2.96
1.62
67
493
648
9.55
1.19
2.83
1.58
68
493
648
9.55
1.19
2.83
1.58
69
489
626
9.71
1.24
2.96
1.62
70
489
626
9.71
1.24
2.96
1.62
71
489
626
9.71
1.24
2.96
1.62
72
489
626
9.71
1.24
2.96
1.62
73
493
648
9.55
1.19
2.83
1.58
74
493
648
9.55
1.19
2.83
1.58
75
493
648
9.55
1.19
2.83
1.58
76
493
648
9.55
1.19
2.83
1.58
77
489
626
9.71
1.24
2.96
1.62
78
493
648
9.55
1.19
2.83
1.58
79
489
626
9.71
1.24
2.96
1.62
80
493
648
9.55
1.19
2.83
1.58
81
489
626
9.71
1.24
2.96
1.62
82
489
626
9.71
1.24
2.96
1.62
83
493
648
9.55
1.19
2.83
1.58
84
493
648
9.55
1.19
2.83
1.58
Note: 1) The following designations of the basic fragments are used in chemical formulae of poly(ether ketones):

Appendix 3. The example of solving the complex problem


analysis of the chemical structure of phenol
formaldehyde resin

The example of solving the complex problem of polymer synthesis belongs to


the analysis of the chemical structure of one of the representatives of polymer
networks phenol-formaldehyde resin widely applied, in particular, to production of
pressed woody articles.
Before analyzing the structure of a polymer network based on phenol
formaldehyde resin, it should be noted that although phenol formaldehyde resins
(PFR) are one of the oldest representatives of high cross-linked polymer networks,
their structure is still of great interest [201]. There is a broad list of works devoted to
the analysis of the structure and properties of PFR. However, there is still no general
point of view on the chemical structure of the network formed by these resins. The
reason is, apparently, in the fact that due to proceeding of the PFR curing the network
structure obtained may be different. Another reason is associated with the fact that
OH-groups capable of forming hydrogen bonds are present in the network structure.
These hydrogen bonds may be displayed in both intermolecular interaction of these
groups disposed in neighboring PFR chains and at the sacrifice of the same interaction
OH-groups disposed in the same repeat unit.

Figure A-3-1. Schematic representation of a network of hydrogen bonds (a) appearing between
neighboring chains and (b) within the same repeat unit.

The meaning of the notions of intermolecular interaction and intramolecular


interaction in the context of the approach under discussion should be explained with
the help of a simplified scheme represented in Figure A-3-1. Figure A-3-1a
demonstrates the case corresponding to the intermolecular interaction, which leads to
the formation of a network of physical (hydrogen) bonds linking chains. Clearly, the
formation of this network promotes an increase of the glass transition temperature.
The case displayed in Figure A-3-1b reflects the situation when two OH-groups are in
the same unit of the linear fragment of PFR, and hydrogen bonding then occurs in the
frames of the same unit, and the network of physical bonds between neighboring ones
is not formed. This problem was analyzed above in detail on the example of
poly(imidophenyl quinoxalines) and poly(amidophenyl quinoxalines). It stands to
reason that various types of intermolecular interaction (hydrogen bonding) causing a
significant influence on PFR properties, as it will be shown below, may coexist in
reality.
In cured PFR, according to the existing data, the following structures
(repeating fragments of the network) may exist:

608

Structure 1:

CH2

CH2

CH2
OH

This structure is included only in the network the repeat unit of which is
marked by dotted lines.
The glass transition temperature of the network will be calculated using
equation (IV.85).
According to the definition of the chemical structure of the cross-linked point
of the network, the structure of the cross-linked point is formed by a selection of
atoms, including the one from which the chain is branched, and the neighboring atoms
with their substituents, as well. For structure 1 of the network, the structure of the
cross-linked point is marked by dotted lines:
CH2

OH
Clearly, the whole cross-linked point contains three such selections of atoms.
Then for the structure 1, we obtain*:

Vi = 3(VC,19 + VC,12 + 2VH,124) + 2VC,18 + 2VH,124 + VC,20 + VO,133


i
1
+ VH,125 = 3(8.4 + 12.6 + 22.0) + 212.7 + 22.0 + 11.65 + 5.2 + 4.7 = 101.1 3;

ai Vi + b j = aC(2VC,18 + VC,20) + aH(2VH,124 + VC,125) + aO,sVO,133 + bh


i

1
= [0.021(212.7 + 11.65) + 19.93(22.0 + 4.7) 7.035.2 139.6]103 = 71.6103 3
K1;

K i Vi = KC(VC,19 + VC,12) + KH21/2VH,124 = [1.15(8.4 + 12.6) +


i
1
2.30721/22.0]103 = 21.5103 3K1.

Substituting values of Vi , ai Vi + b j , and K i Vi into


j
i
1 i
i
1
1
equation (IV.85), we obtain the following value of the glass transition temperature of
the network in the case of the ideal structure 1:

Here and below, numbers of atoms correspond to their numbers in Table 3.

609

Tg =

101.1
103 = 743 K.
21.5 3 + 71.6

(A-3-1)

The calculated value of Tg is significantly above the range of experimental


values, Tg,exper, shown in Table A-3-1 and, consequently, the real PFR network will
differ by structure from the ideal structure 1.
Table A-3-1
Experimental values of the characteristics of cured PFR
Property name
Unit
Limits of measurements
Glass transition temperature, Tg,exper
K
343433
1.241.33
g/cm3
Density, (exper.)
Refractive index, n (exper.)
rel. un.
1.571.63
K1
Coefficient of volumetric expansion, G (exper.)
(1.802.25)104
s
2948
cal/(moldeg)
Molar heat capacity,
(exper.)

Cp

Structure 2:
CH2

CH2
HO
CH2
OH

This possible structure of PFR relates to a linear polymer (the repeat unit is
marked by dotted lines).
For this structure,
V

i = 3VC,19 + 2VC,18 + 1/22VC,12 + VC,20 + VC,40 + 41/2VH,124 + 2VH,124


i

+ 2VH,124 + 2VH,125 + VO,133 + VO,132 = 38.4 + 212.7 + 1/2212.6 + 11.65 + 16.0


+ 41/22.0 + 22.0 + 22.0 + 24.7 + 5.2 + 5.6 = 123.05 3;
ai Vi = aC(3VC,19 + 2VC,18 + 1/22VC,12 + VC,20 + VC,40) + aH(41/2VH,124 +
i

2VH,124 + 2VH,124 + 2VH,125) + aO,s(VO,133 + VO,132) = [0.021(38.4 + 212.7 +


1/2212.6 + 11.65 + 16.0) + 19.98(41/22.0 + 22.0 + 22.0 + 24.7) + 7.03(5.2 +
5.6)]103 = 505.4103 3K1.
If both OH-groups participate in the formation of hydrogen bonds between
chains (as shown in Figure A-3-1), then

bj
j

= 2bh + bM;

bj

= (139.62 + 16.0)103 = 263.2103 3K1.

Substitution of all values of parameters into equation (IV.41) leads to the


following value of Tg of the ideal structure 2:
Tg =

123.05
103 = 508 K.
505.4 263.2

(A-3-2)

610

Similar to the structure 1, this value is higher than the experimental


temperature interval in which the -transition is observed for PFR.
Structure 3:

CH2

CH2
OH

This structure forms only linear chains, the repeat unit of which is marked by
dotted lines. For structure 3, we obtain:

Vi = 2VC,19 + VC,20 + 3VC,18 + 1/22VC,12 + 1/24VH,124 + VH,125 +


i
3
3VH,124 + VO,133 = 28.4 + 11.65 + 312.7 + 1/2212.6 + 1/242.0 + 4.7 + 32.0 +
5.2 = 99.05 3;

ai Vi = aC(2VC,19 + VC,20 + 3VC,18 + 1/22VC,12) + aH(1/24VH,124 +


i
3
VH,125 + 3VH,124) + aO,sVO,133 = [0.021(28.4 + 11.65 + 312.7 + 1/2212.6) +
19.98(1/242.0 + 4.7+ 32.0) + 7.03 5.2]103 = 331.9103 3K1.
b j = 2bh + bM; b j = (139.6 + 16.0)103 = 123.6103 3K1.
j

Substitution of all these values of parameters into equation (IV.41) leads to the
following value of Tg of the ideal structure 3:
Tg =

99.05
103 = 476 K.
331.9 123.6

(A-3-3)

This value of Tg is somewhat above the temperature range of the -transition


for PFR.
Structure 4:

CH2

CH2
O

This structure forms a network, the repeating fragment of which is marked by


dotted lines. The structure of the cross-linked point of this network is the following
(marked by dotted lines):

H2C

CH2
O

611

For structure 4, we obtain:

Vi = 2VC,19 + VC,20 + 1/22VC,12 + 3VC,18 + 3VH,124 + 41/2VH,124 +


i
4
VO,131 = 38.4 + 11.65 + 1/2212.6 + 312.7 + 32.0 + 41/22.0 + 2.1 = 91.25 3;

ai Vi = aC(3VC,18) + aH(3VH,124) = (0.021312.7 + 19.9832.0)103 =


i
4
120.7103 3K1;
b j = 0;
j

K i Vi = KC(2VC,19 + VC,20 + 1/22VC,12) + KH41/2VH,124 + KOVO,131 =


i
4
[1.15(28.4 + 11.65 + 1/2212.6) + 2.30741/22.0 + 0.0582.1]103 = 56.5103 3K
1
.
Substituting these values into equation (IV.85), we obtain the following value
of Tg for the ideal network, composed from the structure 4:
Tg =

91.25
103 = 515 K.
56.5 + 120.7

(A-3-4)

Hence, the glass transition temperature of this cross-linked polymer network


would be quite high compared with Tg,exper.
Structure 5:
CH2

CH2
HO

CH2
OH
From the viewpoint of the gross-formula, this structure is identical to the
structure 2, but in the present case, hydrogen bonding appears within the single repeat
unit. Such a structure is described in ref. [99].
If hydrogen bonds are formed between neighboring OH-groups disposed
within the same repeat unit, then, as mentioned above, the physical network between
chains stipulated by hydrogen bonds is not formed. Then calculating the glass
transition temperature, the value of b j is calculated in the following way: because
j

this structure is linear, the constant bM must be taken into calculation when
considering the m-substitution in an aromatic cycle; moreover, there are two branches
from the aromatic cycle, which requires introduction of two constants bd. Then b j
j

= bM + 2bd = [16.0 255.4]103 = 94.8103 3K1. Substituting values

ai Vi
i

and

bj

Vi ,
i

into equation (IV.85), the following value of the glass transition

temperature is obtained for the structure 5:

612

Tg =

123.05
103 = 300 K.
505.4 94.8

(A-3-5)

This value of Tg falls below the range of the -transition for cured PFR.
As indicated, no calculated values of the glass transition temperature of the
ideal structures of PFR discussed falls within the range of experimental values of the
glass transition temperature, Tg,exper. It may therefore be assumed that there is a more
complicated structure of cured PFR than the considered ideal ones; this has been
indicated by many investigators. Most probably, the real structure of PFR consists of
the selection of the considered ideal structures.
To state the most probablr real PFR structure, let us calculate, besides Tg, the
following characteristics of cured PFR: density , refractive index n, coefficient of
volumetric expansion G, and molar heat capacity C ps . Selection of these characteristics is stipulated by both their importance and sensitivity to changes of the structural
formula of the substance and the fact that for various ideal PFR structures, as the glass
transition temperature increases, values of other characteristics may increase or
decrease, which is important for search of the most probable structure of PFR by the
method of experiment design (see below).
The calculated values of PFR density for structures 15, obtained due to
equation (II.6), are indicated in Table A-3-2.
Table A-3-2
Calculated values of Tg, , n, G and C sp for the ideal structures of cured PFR
Ideal structure of PFR
Structure 1
Structure 2
Structure 3
Structure 4
Structure 5

Tg, K

, g/cm3

n, rel. un.

G104, K1

743
507
476
515
300

1.253
1.250
1.205
1.301
1.250

1.621
1.583
1.604
1.645
1.583

1.30
3.20
2.02
1.86
3.20

C sp ,
cal/(moldeg)
31.9
42.4
31.4
26.1
42.4

Let us calculate the refractive index n by equation (VIII.2), which contains the
molar refraction summed up from the refraction indices of separate atoms and
constants for different types of bonds (double, triple).
For structure 1, we obtain (values of Ri are taken from [28]):
R1 = 3(RC + 1/2RC + RH) + 2RC + 2RH + RC + RO + RH + 3R = 3(2.418 + 1/22.418
+ 1.10) + 22.418 + 21.10 + 2.418 + 1.525 + 1.10 + 31.733 = 31.459 cm3/mol.
For structure 2:
R2 = 8RC + 8RH + 2RO + 3R = 82.418 + 81.10 + 21.525 + 31.733 = 36.393
cm3/mol.
For structure 3:
R3 = 7RC + 6RH + RO + 3R = 72.418 + 61.10 + 1.525 + 31.733 = 30.25 cm3/mol.
For structure 4:
R4 = 7RC + 5RH + RO< + 3R = 72.418 + 51.10 + 1.643 + 31.733 = 29.268 cm3/mol.
For structure 5, the value R5 coincides with the value R2. Values of n for
various structures calculated by equation (VIII.2) are shown in Table A-3-2.
Heat capacity C ps in the glassy state of the polymer was calculated by
correlation (XVI.1).

613

The value G is directly connected with the glass transition temperature Tg by


the correlation:

G =

0.096
.
Tg ( K )

(A-3-6)

The calculated values of C ps and G for all five ideal structures are shown in
Table A-3-2.
Comparison of the calculated values of PFR characteristics for possible ideal
structures, shown in Table A-3-2, with their experimental values (see Table A-3-1)
enables us to conclude that cured PFR does not represent an ideal network, but
contains a set of structures; this set can be determined using the method of experiment
design for multi-component systems with composing simplex networks and
polynomial composition property models. For independent variables X of these
models, let us assume the above-considered ideal structures the combination of which
will determine the real structural composition of cured PFR: x1 for structure 1; x2 for
structure 2; x3 for structure 3; and x4 for structure 4. For the response function Y, the
following properties of cured PFR will be taken: glass transition temperature, Tg (K);
density, (g/cm3); refractive index, n (rel. un.); coefficient of volumetric expansion,
G (K1); and molar heat capacity, C ps (cal/(moldeg)).
As indicated, instead of the ideal structure 2, the structure 5 is assumed as the
independent variable; this structure differs by the existence of hydrogen bonding of
OH-groups within the same repeat unit which, with all other characteristics being the
same, provides for lower glass transition temperatures (see Table A-3-2).
The dependence of the response functions Y on variables X will be searched
for in the form of the second order model [70]:
Y = ij + ijxixj;
where

i = y i;

1 i 4;

1 j 4,

ij = 4yij 2yi 2yj.

(A-3-7)
(A-3-8)

Design of the experiment and appropriate simplex lattice are shown in Table
A-3-3 and Figure A-3-2a, respectively.
Table A-3-3
Matrix of the experimental design and calculated values of PFR at the experimental design
Coordinates of
Calculated values of PFR
Point of the experiment design points
experiment
C sp ,
Tg,
n,
,
G104,
design
x1
x2
x3
x4
3
1
rel.
un.
K
g/cm
K
cal/(moldeg)
y1
1
0
0
0
743
1.253
1.621
1.30
31.90
y2
0
1
0
0
300
1.250
1.583
3.20
42.40
y3
0
0
1
0
476
1.205
1.604
2.02
31.40
y4
0
0
0
1
515
1.301
1.645
1.86
26.10
y12

1/2
0
0
410
1.251
1.600
2.34
37.20
y23
0
1/2
1/2
0
359
1.232
1.594
2.67
36.90
y13

0
1/2
0
582
1.232
1.613
1.65
31.65
y14

0
0
1/2
615
1.276
1.633
1.56
29.00
y24
0
1/2
0
1/2
364
1.272
1.614
2.64
34.25
y34
0
0
1/2
1/2
494
1.254
1.624
1.94
28.75

614

Figure A-3-2. Designation of responses at the points of simplex lattice: a) experimental design points;
b) control design points.

As indicated by the diagram, the experimental design points with the single
index, yi, are located in the tetrahedron apexes, with two indices, yij in the middle of
the tetrahedron edges possessing apexes, i and j. Therewith, values of the response
functions Tg, , n, G, and C ps in the points of the experimental design yi, located in
the apexes of the tetrahedron, are calculated by equations (IV.85), (II.6), (VIII.2),
(XVI.1) and (A-3-6), and in intermediate points of the experiment, i.e. with two or
more indices at y, by the equations:
for the glass transition temperature,

1 Vi + 3 Vi + 4 Vi +
i
1
i
3
i
4

Tg =

1 i Vi + b j + K i Vi + 3 i Vi + b j +
j
j
1
i
1 i
i
3

+ 5 Vi
i
5


+ 4 i Vi + b j + K i Vi + 5 i Vi + b j
j
j
i
4
4 i
i
5

for the density,

0.681(1 M 1 + 3 M 3 + 4 M 4 + 5 M 5 )

(A-3-10)


N A 1 Vi + 3 Vi + 4 Vi + 5 Vi
1
i
3
i
4
i
5
i

; (A-3-11)


N A 1 Vi + 3 Vi + 4 Vi + 5 Vi
i
1
i
3
i
4
i
5

for the refractive index,

n2 1
=
n2 + 2

(A-3-9)

0.681( 1 R1 + 3 R3 + 4 R4 + 5 R5 )

for the molar heat capacity,

615

C ps = 1 C ps ,i Vi + 3 C ps ,i Vi + 4 C ps ,i Vi +
i
1
i
3
i
4

+ 5 C ps ,i Vi + C 0 ;
i
5

(A-3-12)

C ps = 1C ps ,1 + 3C ps ,3 + 4 C ps , 4 + 5C ps ,5 + C 0 .
The coefficient of volumetric expansion is calculated by correlation (A-3-6),
the value Tg being determined from equation (A-3-9). In correlations (A-3-9) (A-312), 1, 3, 4, and 5 are the molar parts of the structures 1, 3, 4, and 5, respectively;
therewith, 1 = x1, 3 = x3, 4 = x4, and 5 = x5.
Using the response values in the design points shown in Table A-3-3 and
equations (A-3-8) for calculation of the coefficient for the second order model (A-37), polynomial models for cured PFR are found, in which insignificant coefficients are
neglected:
for the glass transition temperature, K:
Tg = 743x1 + 300x2 + 476x3 + 515x4 446x1x2 116x1x3
56x1x4 116x2x3 174x2x4;

for the density, g/cm3:

= 1.253x1 + 1.250x2 + 1.205x3 + 1.301x4 + 0.012x1x3 +


+ 0.018x2x3 0.014x2x4;

(A-3-15)

for the coefficient of volumetric expansion, K1:

G = (1.3x1 + 3.2x2 + 2.02x3 + 1.86x4 + 0.36x1x2 +


+ 0.24x2x3 + 0.44x2x4)104;

(A-3-14)

for the refractive index, rel. un.:


n = 1.621x1 + 1.583x2 + 1.604x3 + 1.646x4 0.009x1x2;

(A-3-13)

(A-3-16)

for the molar heat capacity, cal/(moldeg):


C ps = 31.9x1 + 42.4x2 + 31.4x3 + 26.1x4 + 0.2x1x2.

(A-3-17)

Using polynomial models (A-3-13) (A-3-17) obtained and equations


(IV.85), (II.6), (VIII.2), (XVI.1), and (A-3-6) in the control design points shown in
Figure A-3-2b (the design point with three indices, yijk, locates in the middle of the
tetrahedron designe with apexes i, j and k), the properties of cured PFR are calculated.
As indicated in Table A-3-2, the relative error of the properties prognosis for
cured PFR does not exceed 1.7%, which is quite acceptable for application to
polynomial models both in order to determine the real structure of cured PFR and the
degree of influence of the ideal structure type on the PFR properties. In particular,
indicated as an example in Figure A-3-3 is graphic interpretation of the polynomial
model of the glass transition temperature for PFR (A-3-13) in the area of change of

616

Control design
points

Coordinates of control design


points

Values of PFR properties in control design points

x1

x2

x3

x4

Tg/T g

n/n

(G/G)104

y123

1/3

1/3

1/3

y124

1/3

1/3

1/3

y134

1/3

1/3

1/3

y234

1/3

1/3

1/3

428
431
436
442
559
558
394
397

1.238
1.259
1.265
1.266
1.253
1.254
1.252
1.253

1.603
1.602
1.617
1.615
1.623
1.624
1.611
1.611

2.240
2.235
2.200
2.200
1.720
1.714
2.440
2.436

C sp C sp
35.23
35.23
33.47
33.47
29.80
29.80
33.3
33.3

Relative error of the predicted properties of


PFR, %
n
G
Tg

Cs

0.7

1.7

0.1

0.2

1.4

0.1

0.1

0.2

0.1

0.1

0.4

0.8

0.1

0.2

Notes:
1. Shown in the numerator are properties of PFR calculated by equations (109), (6), (VII.2), (XVI.1), and (A-3-6). The denominator indicates the properties calculated by
polynomial models (A-3-13) (A-3-17).
2. Relative error of the predicted properties of PFR is calculated by the formula:
y ijjk =

y ijk y ijk
100% .

y ijk

616

Table A-3-4
Errors of the predicted properties of cured PFRs

617

the ideal structures 1, 3 and 5 ratio (in the absence of the ideal structure 4). As
indicated, isolines of the glass transition temperature of PFR represent straight lines,
suggesting that contributions of separate ideal structures into formation of properties
of the real desired PFR structure are additive and, the more so, the glass transition of
PFR is preferably determined by the relative concentration of the ideal structures 1
and 5.

Figure A-3-3. Isolines of the glass transition temperature Tg of PFR in the region of ideal structures x1,
x2 and x3 (x4 = 0).

Searching for the real structure of cured PFR includes the determination of the
ratio between contents of the ideal structures 1, 3, 4, and 5, at which border conditions
on the experimental values of the resin properties, indicated in Table A-3-1, are
fulfilled. The acceptable selection of these ideal structures x (at xi = 1, 0 xi 1,
i

i = 1 4 ) is set on in the following sequence. When the concentration of the structure


4 is fixed (x4 = const) for every tested property of the resin (Tg, , n, G, and C ps ),
regions of acceptable selections of the remaining three ideal structures are determined
in the plane of change of variables x1, x2 and x3. The desired value x1x4 =const is
obtained by the consequent superimposition of these regions.
The example of this procedure at x4 = const is indicated in Figure A-3-4. As
indicated, such indices of cured PFR as the refractive index (Figure A-3-4c) and the
molar heat capacity (Figure A-3-4e) cause no influence on the formation of the region
of the acceptable selection of the ideal structures x1x4 =0 (Figure A-3-4d).
Apparently, the influence of these characteristics of PFR is evident only for ideal
structure 4 and increases with the concentration of the latter (Figure A-3-5).

618

Figure A-3-4. Schematic representation of the search procedure for an acceptable set x of ideal PFR
7
structures for x4 = 0. Showing domains of the acceptable sets of structures: (a)
[[ =

[[ = for the density; (c)


Q
G
[[ = for the refractive index; (d)
x1x4 =0 for the coefficient of volumetric
&
expansion; (e)
[ [ = for the molar heat capacity; (f)
x x =0 for the ideal PFR

for the glass transition temperature of cured PFR: (b)

 

1 4

structures with the coordinates of vertices T1 (0.35; 0.34; 0.31; 0.00); T2 (0.34; 0.35;
0.31; 0.00); T3 (0.55; 0.45; 0.00; 0.00).

The region of acceptable selection of the ideal structures of cured PFR is


determined by consecutive sorting out of the regions x1x4 =const at variation of the
structure x4 content with step 0.01 from 0 to 1. Consequently, it is obtained that that
the desired region x represents an improper figure and the coordinates of apexes of
this figure are shown in Figure A-3-6. Disposition of x regions in the factor space of
ideal structures of PFR enables us to conclude that the real structure of cured PFR
may be represented only by the combination of linear (structures 3 and 5) and network
(structures 1 and 4) fragments of the polymeric chain; this has also been confirmed by
other investigators [99].

Figure A-3-5. Effect of the ideal structure 4 on the change of the domain x of acceptable set of ideal
structures of cured PFR for the refractive index (I) and for the molar heat capacity (II) at
x4 = 0 (a); 0.2 (b); 0.4 (c).

619

For the quantitative estimation of the properties of PFR containing various


types of structures, let us substitute all calculated parameters into equations (A-3-13)
(A-3-17). Consequently, we obtain the following expressions for calculation of the
glass transition temperature Tg, density , refractive index n, coefficient of volumetric
expansion G, and molar heat capacity C ps :
Tg =

1 101.1 + 5 123.05 + 3 99.05 + 4 91.25 3


10 ;
1 136.1 + 5 410.6 + 3 208.3 + 4 177.2

0.681(1 112 + 5 136 + 3 106 + 4 105)


103 ;
0.6023(1 101.1 + 5 123.05 + 3 99.05 + 4 91.25)

(A-3-18)

(A-3-19)

n 2 1 0.681(1 31.459 + 5 36.393 + 3 30.25 + 4 29.268) 3


10 ; (A-3-20)
=
n 2 + 2 0.6023(1 101.1 + 5 123.05 + 3 99.05 + 4 91.25)

G = 0.0096/Tg;
C ps = 131.9 + 542.4 + 331.4 + 426.1,

(A-3-21)
(A-3-22)

where 1, 3, 4, and 5 are the molar parts of the structures 1, 3, 4 and 5,


respectively.
One of the possible relations between relative content of ideal structures,
responded to the region of admissible selection (Figure A-3-6) has the following
form: 1 = 0.345, 5 = 0.345, 3 = 0.2, and 4 = 0.11.

Figure A-3-6. Coordinates of vetrices of the domain of admissible set of ideal structures x for cured
PFR: T1 (0.35; 0.34; 0.31; 0.00); T2 (0.34; 0.35; 0.31; 0.00); T3 (0.55; 0.45; 0.00; 0.00);
T4 (0.00; 0.22; 0.13; 0.65).

Substituting these values into correlations (A-3-18) (A-3-22), for the real
network we obtain: Tg = 429 K; = 1.248 g/cm3; n = 1.6057; G = 2.238104 K1;
C ps = 34.8 cal/(moldeg); as indicated (see Figure A-3-3), all these values fall within
the range of experimental values.

620

Taking into account the presence of a narrow region of the acceptable set of
ideal structures for cured PFR (see Figure A-3-6), it may be concluded that the real
structure of the network obtained is close to the most probable one.
Hence, the approach described in the Section IV.4, combined with the method
of experiment design, enables, firstly, to made definite conclusions about the structure
of cross-linked systems, basing on the comparison of experimental and calculated
characteristics of the network. Secondly, the examined problem of the analysis of
phenol formaldehyde resin structure relates to the mixed problem, because the direct
problem is solved in the initial stage the properties of ideal structures of PFR are
determined by their chemical structure. In the second stage of investigations, solution
of the reverse problem is considered the chemical structure of PFR is searched,
which provides for the desirable set of properties of this resin. In this case, such set of
properties is given by their experimental values shown in Table A-3-1. In the present
example, note also the application of the approach to solving the reverse problem of
polymers synthesis, different from the one described in Appendix 2.

Appendix 4. Application of the approach to multicomponent


copolymers

Usually, analysis of the structure and properties of copolymers was performed


for the systems containing two components. It has been was performed much less
frequently for triple copolymers. In this appendix, applicability of the approach,
described in the present monograph, to analysis of the structure and properties of
copolymers containing from three to five copolymers will be discussed.
Simultaneously, let us compare experimental and calculated values of the physical
characteristics for both homopolymers and homopolymer-based multicomponent
copolymers.
To analyze the properties of multicomponent copolymers, six monomers are
chosen in ref. [39], used in reactions of polymerization and copolymerization: methyl
methacrylate (MMA), butyl methacrylate (BMA), butyl acrylate (BA), n-nonyl
acrylate (NA), 2-ethylhexyl acrylate (EHA), and n-heptyl acrylate (HA).
Homopolymers and copolymers based on these compounds are fully amorphous highmolecular substances, which are easily formed in a block and suitable for
investigation of thermal and optical characteristics. Indicated in Table A-4-1 is the
composition of five copolymers synthesized in ref. [39]. Calculation of the glass
transition temperature Tg for multicomponent copolymers is performed by equations
(IV.55) and (IV.57). The first of these equations for calculation of the glass transition
temperature of copolymers requires no understanding of the experimental glass
transition temperatures of homopolymers. Equation (IV.57) contains values of glass
transition temperature of homopolymers, for calculation of Tg of copolymers
experimental values of Tg for appropriate homopolymers being used.
Table A-4-1
Copolymer
1
2
3
4
5

MMA
0.333
0.250
0.250
0.200
0.250

Composition of copolymers
Molar parts of units in copolymer
BMA
BA
NA
0.333
0.333

0.250
0.250
0.250
0.250
0.250

0.200
0.200
0.200
0.250
0.250

EHA

0.250
0.200

HA

0.250

Density is calculated by equations (II.7), (A-4-1) and (A-4-2). The first of


them is approximate, obtained on the base of suggestion that the coefficient of
molecular packing is approximately the same for all polymers, and equals kavg = 0.681
for all block polymers at room temperature. However, for polymer and copolymers
existing in the rubbery state at room temperature, it is desirable to take into account
temperature dependences of the coefficient of molecular packing, k, in the density
calculation, because the difference of the real coefficient of packing for them from the
average value kavg, increases with a decrease of the glass transition temperature.
Equations (A-4-1) and (A-4-2) are obtained with regard to this circumstance and
enable, as indicated below, to calculate more accurately the density of polymers and
copolymers. For the latter, we obtain:

622

Tg, K
Td, K
C, B
, g/cm3
G104, K-1
3.30
1.17 1.23
378
2.70
623
2.85
1.17; 1.17
377
2.55
629
PBMA

300
1.48
2.82
17.7 18.4
1.06
1.09; 1.06
287
1.50
2.71
18.2
27.2
3.05
627
16.6
PBA

221
1.47
18.2 18.4
30.7
2.80
1.04
1.11; 1.05
242
1.50
2.76
18.4
28.0
3.05
627
32.2
PNA

215
1.04; 0.97
219
1.51
2.58
17.5
25.3
3.05
626
40.5
PEHA



1.05; 0.99
242
1.51
2.60
17.5
3.05
626
39.4
PHA

220
1.06; 0.99
225
1.51
2.63
17.7
26.1
3.05
626
38.1
Notes: 1. * - shown in the numerator are experimental values of physical characteristics; indicated in the denominator are calculated values. In the case calculated values,
the first of them was obtained by equation (II.7), and the second by equations (A-4-1) and (A-4-2);
2. is the density; Tg is the glass transition temperature; n is the refractive index; is the dielectric constant; is the solubility parameter; is the surface energy; G is the
thermal coefficient of volumetric expansion in the glassy state; C is the stress-optical coefficient.
Homopolymer
PMMA

622

Table A-4-2
Calculated and experimental characteristics of homopolymers*
N
, (J/cm3)1/2
, dyn/cm

1.49; 1.55
3.10
18.6
39.0 40.2
1.49
2.94
19.0
31.0

623

k =n

G20 =

0.667 k M k
k =1

k =n

[1 + G ( 293 Tg )]N A k Vi
k =1
i
k

, 293 < Tg;

(A-4-1)

, 293 > Tg.

(A-4-2)

k =n

G20 =

0.667 k M k
k =1

k =n

[1 + L (293 Tg )]N A k Vi
k =1
i
k

Besides the glass transition temperature and density, the following


characteristics are calculated for copolymers: the refractive index n, the stress-optical
coefficient C, temperature of the onset of intense thermal degradation Td, the
dielectric constant , the solubility parameter , the surface energy . Calculations
were performed using equations (VIII.3), (VII.7), (VIII.16), (IX.4), (XII.4), and
(XIII.39), respectively. First of all, let us compare the calculated and experimental
values of the properties of homopolymers. These data are indicated in Table A-4-2. In
most cases, good coincidence between experimental and calculated data is observed.
Separately, we must dwell on the calculation of such characteristic of homo- and
copolymers as density , for which two values are shown in the Table. The first of
them is determined with the help of equation (II.7), and the second by correlations
(A-4-1) and (A-4-2), which take into account the temperature dependence of the
coefficient of molecular packing. In the case of a glassy polymer (poly(methyl
methacrylate)), the calculated value of density obtained by equation (II.7) coincides
with the experimental value. For homopolymers with low glass transition
temperatures, which are in the rubbery state at room temperature, consideration of the
temperature dependence to the coefficient of molecular packing leads to lower values
of density compared with those determined on the base of the value kavg. Therewith,
the data on density values coincide with experimental ones.
The glass transition temperature, determined from equation (IV.41) with the
accuracy usual for such calculations, coincides with the experimental values. The
refractive index for homopolymers is calculated from equation (VIII.1), as well as
from equation (VIII.6), which takes into account the temperature dependence of the
coefficient of molecular packing. The latter circumstance enables us to calculate the
refractive index with high accuracy.
With regard to such characteristics as the solubility parameter, surface energy,
temperature of the onset of intense thermal degradation, the thermal coefficient of
volumetric expansion in the glassy state, and the stress-optical coefficient, they
coincide with the experimental ones with the accuracy usual for such calculations.
The properties of multi-component copolymers are indicated in Table A-4-3.
Here, the coincidence of the calculated and experimental characteristics is
approximately the same as for homopolymers. The density calculated from equation
(II.7) using the average value of the coefficient of molecular packing is somewhat
higher than the values, determined experimentally. This is stipulated by the fact that
all copolymers possess the glass transition temperature below room temperature.
Values of density calculated by correlation (A-4-2) with regard to the temperature
dependence of the coefficient of molecular packing, is somewhat lower than these
calculated by equation (II.7), and coincide well with the experimental values of .

624

Table A-4-3
Physical characteristics of homopolymers*
Van-der-Waals volume,
n
Tg, K
Td, K
Copolymer
C, B
, g/cm3

1
125

563
298
1.09
1.50
15.5
273;268
627
1.12;1.08
2
148

573
237
1.04
1.50
21.6
247;243
627
1.09;1.03
3
143

570
233
1.04
1.50
21.3
258;255
627
1.09;1.04
4
158

610
240
1.02
1.50
25.2
244;240
627
1.08;1.02
5
139

570
243
1.06
1.50
21.0
252;247
627
1.10;1.04
Notes: 1. * Indicated in the numerator are experimental values, and in the denominator calculated
ones; for density , the first value is obtained from equation (II.7), and the second from equation (A4-2); for Tg, the first value is calculated by equation (IV.55), and the second by correlation (IV.57),
using experimental values of Tg for homopolymers.
2. is the density; n is the refractive index; C is the stress-optical coefficient; Tg is the glass transition
temperature; Td is the onset temperature of intense thermal degradation.

For all copolymers, the refractive index n is approximately the same, and is
associated with the close values of the refractive index for all homopolymers,
mentioned above. The stress-optical coefficient is also almost identical for all
copolymers, excluding the copolymer 1 with increased concentration of MMA units
possessing a negative value of C.
The glass transition temperature of copolymers indicated in Table A-4-3 is
calculated by equations (IV.55) and (IV.57). The former requires no understanding of
experimental values of the glass transition temperature of homopolymers, and the
latter uses these values. By and large good coincidence of calculated and experimental
values of Td is observed. In most cases, when experimental values of Tg for
homopolymers are used and substituted to equation (IV.57), correspondence of the
calculation to the experiment becomes somewhat better. As for temperature of the
onset of rapid thermal degradation Td, good coincidence is observed for the
copolymer 4, for which the error is 2.7%. For other copolymers, the calculated values
deviated from the experimental ones for this characteristic by 50 to higher values
which gives 8%. This phenomenon has not as yet been explained. It should be taken
into account only that the characteristic such as the dependence of the temperature of
the onset of intense thermal degradation on the presence of admixtures and other
reasons is stronger. The calculated values of this characteristic are determined for the
ideal polymeric system.
Hence, the possibility of predicting various characteristics of multi-component
copolymers is obvious. Therewith, the approach used does not require any
experimentally determined values. The prognosis is performed on the basis of only
the chemical structure of the multi-component copolymer.

Appendix 5. Influence of strong intermolecular interaction


occurring between two dissimilar polymers on
their miscibility

When two or more polymers are mixed, the situation may appear when an
additional strong intermolecular interaction occurs between chains of these polymers,
which is not displayed between macromolecules of each of polymers, taken
separately. These may be hydrogen bonds or strong dipoledipole interactions.
Analysis of the influence of these interactions on the glass transition temperature will
be shown below.
In a series of works devoted to the analysis of miscibility of polymers and the
properties of mixtures obtained, the main attention is paid to these special interactions
occurring between macromolecular chains of the mixed polymers. The scheme of this
interaction is demonstrated by the following example:

F3C

CF3
O
H
O

Let us consider in detail examples of such interactions and experimental


methods of their determination. Mixtures of styrene copolymer with vinylphenylhexafluoromethylcarbynol with the polymers as polycarbonate based on bisphenol A,
poly(butyl methacrylate), poly-2,6-dimethyl-1,4-phenylene oxide may represent an
example of this significant event [209]. Introduction of hydroxyl groups to the
mentioned polymer leads to the formation of hydrogen bonds between the copolymer
and polycarbonate, which causes a sharp increase of miscibility. The thermograms
obtained by the DSC method for polystyrene mixtures with polycarbonate show
clearly two glass transition temperatures indicating immiscibility of these polymers.
The situation is changed abruptly when polystyrene introduced to the mixture is
substituted by a copolymer of the above given structure: thermograms indicate the
single glass transition temperature (which indicates miscibility). Therewith, the glass
transition temperature regularly increases with the concentration of polycarbonate.
Mixtures of copolymers of the above-mentioned structure with poly(ethylene oxide)
(PEO), which represents the solid polymer, also display this effect [210]. Crystallinity
of PEO in the mixture varies in the manner by which the melting point decreases. The
formation of hydrogen bonds between PEO and copolymer has been studied in
dependence on temperature. As temperature increases, hydrogen bonds between
copolymer and PEO dissociate, but are restored when cooled down. Even in the case
of solid polymer such as PEO, the presence of hydrogen bonding between chains of

626

mixed polymers improves their miscibility, suppresses crystallization, and forms onephase systems.
Behavior of styrene copolymer mixtures with vinylphenylhexafluoromethyl
carbynol was studied for such polymers as poly(vinyl acetate), poly(methyl
methacrylate), poly(ethyl methacrylate), poly-n-(butyl methacrylate), poly(methyl
vinyl ether), poly-2,6-dimethyl-1,4-phenylene oxide, polycarbonate based on
bisphenol A, styrene copolymer with acrylonitrile, as well as with amorphous and
solid polyesters and polyamides [193]. For these systems, hydrogen bonding
influences the miscibility of components, which has been confirmed by measurements
of the glass transition temperature, as well as by FTIR-spectroscopy. Ref. [165]
represents a detailed investigation of the miscibility of this polymer with a series of
aliphatic polyamides, such as nylon-6,12 and N,N-dimethyl-substituted nylon-6,12.
The miscibility criterion was again assumed to be general glass transition temperature
for the mixture, therewith, the polymers were used, for which the glass transition
temperature is significantly different from that for polystyrene and copolymer. For
example, the glass transition temperature of nylon-6,12 is 46C, and the melting point
is estimated in the range of 206215C. To estimate the influence of the crystallinity
degree on miscibility, besides nylon-6,12, N,N-dimethyl-substituted nylon-6,12 was
used, as well as copolymers with different concentration of it. The main conclusions
made according to the results of ref. [165] are that hydrogen bonds between
components of the mixture are formed and these bonds dissociate at heating and occur
again at cooling down. Therewith, the introduction of a small amount of carbonyl
groups to polystyrene improved the miscibility of such immiscible polymers as
polystyrene and polyamide. Consequently, the copolymer under consideration may be
a compatibilizer.
Ref. [192] represents the study of the miscibility of substituted phenol
condensed resins with poly(methyl methacrylate) (PMMA). The following resins
were studied:
OH

where R = H; t-butyl; NO2; Cl.


This work indicates again that formation of the hydrogen bonds between
components of the mixture plays a significant role in miscibility. Thermodynamic
parameters of dissociation of hydrogen bonds, such as enthalpy and entropy, were
determined. The most significant influence is caused by such substituents as NO2 and
Cl. All mixtures of these phenol resins with poly(methyl methacrylate) displayed the
single glass transition temperature that meant that they are miscible well. Therewith,
the dependences of the glass transition temperature on the composition were
corresponded to three different cases:
1. The glass transition temperature of mixtures is higher than the mean glass
transition temperature.
2. The glass transition temperature of mixtures is always lower than the mean glass
transition temperature.

627

3. Dependence of the glass transition temperature on the composition is of the Sshape in relation to the mean dependence.
The authors of ref. [192] suggest description of the mixture behavior
appropriate to the cases 1 and 2 by the following correlation:
Tg = W1Tg1 + W2Tg2 + qW1W2,

(A-5-1)

where Tg1 and Tg2 are the glass transition temperatures of polymers 1 and 2,
respectively; W1 and W2 are their weight parts; the parameter q may be interpreted as
the contribution of hydrogen bonds, which may be considered as pseudo-crosslinks.
Therewith, the value of q characterizes intensity of the hydrogen bonding.
The experiments showed that when NO2 and t-butyl are used as substituents,
the values of q are negative and approximately equal by the absolute value. If
substituents are H and Cl, then the values of q are positive (all the above-said is true at
substitution of an aromatic cycle in the para-position).
Consequently, mixtures of phenol resins possessing t-butyl and NO2-groups
with poly(methyl methacrylate) display the behavior corresponding to the second
case. When there is no substitution, the behavior of the system corresponds to the
third case. Finally, when Cl is the substituent, behavior of the mixture corresponds to
the first case.
For the third case, the S-shape of dependence of the glass transition
temperature Tg on the composition may be described by the following correlation:
Tg =

W1Tg1 + kW2Tg2
W1 + kW2

+ qW1W2 ,

(A-5-2)

which is the most general one compared with expression (A-5-1).


Thereafter [190], correlation (A-5-2) was applied to description of the
dependence of the glass transition temperature on the weight part of the components
for mixtures of substituted phenol resins (see above) with such polymers as poly(ethyl
methacrylate) and poly(methyl methacrylate) F and t-butyl being substituents on the
phenol resin, and behavior of the mixtures based on non-substituted phenol resin was
also studied.
All three cases described above were obtained, and all parameters of equation
(A-5-2) were found.
Studied in refs. [170] and [171] was the phenomenon of hydrogen bonding
between macromolecules in mixtures of polymers by the FTIR-spectroscopy, and
using the Patterson and Robald theory [173] and by obtaining diagrams with the
lowest threshold temperature of dissolution. The solubility and miscibility of
poly(ethyl oxazoline)
(

N CH2

CH2

)n

C2H5C O
were studied; poly(ethyl oxazoline) is the isomer of poly-N,N-dimethylacrylamide
(

CH2

CH

)n

O CN(CH3)2

628

To estimate the miscibility of poly(ethyl oxazoline) with other polymers, two


series of experiments were performed. In the first series, the films based on mixtures
of pairs of polymers were obtained, one of which was poly(ethyl oxazoline). In the
second series of experiments, complexes based on the same pairs of polymers were
obtained. These complexes were obtained by mixing polymeric solutions with further
separation of residue, its drying in vacuum during a long time to a constant mass. It
was found that the composition of the complex differs from that of the initial mixture,
the composition of the complex being correspondent to the ratio of polar groups
leading to a strong intermolecular interaction.
The glass transition temperature of poly(ethyl oxazoline) mixed with
polyacrylic acid is below the additive values, whereas for the complexes, the glass
transition temperature is above them. The authors associate this with the formation of
a network of physical bonds in the case of complexes.
The miscibility of poly(vinyl methyl ether) with styrene copolymers with
methyl methacrylate [136] was also studied. The threshold composition of the
copolymer, at which it is miscible with poly(vinyl methyl ether), was found. Such
copolymer must contain about 60% (mol) of polystyrene. Phase diagrams and socalled windows of miscibility for these mixtures were obtained.
Besides the glass transition temperature, the thermal stability of mixtures was
also studied [221] (on the example of poly-p-hydroxystyrene mixtures with poly(vinyl
pyrrolidone) and poly(ethylene oxazoline)). It was found that during heating a small
loss of polyhydroxystyrene mass (6%) in the temperature range of 200250C is
caused by the cross-linking reaction leading to formation of ether bonds. Importance
of hydrogen bonds formation between chains of polymers is also noted.
Intermolecular complexes may be formed directly during polymerization
[135]. For example, this is manifested in photopolymerization of acrylic acid, in
which poly(ethylene oxide) is dissolved. The glass transition temperature of the
complexes exceeds not only the glass transition temperature of mixtures, but also
glass transition temperatures of the original components. These complexes are
dissolved in such solvents as dimethylformamide and dimethylsulfoxide, and
definitely swell in water and methanol; however, they do not swell in dioxane. If the
number of groups able to form hydrogen bonds is decreased by copolymerization of
acrylic acid with methyl methacrylate, the ability to complex formation is preserved,
although it is somewhat reduced.
Thermal oxidation also depends [191] on the formation of hydrogen bonds
between chains of polymers (for example, for mixtures of poly(vinyl methyl ether)
and modified polystyrene). Modified polystyrene represented a styrene copolymer
containing 2.5% (mol) of hexofluoro-2-propylstyrene, i.e. it contained hydroxyl
groups inclined to formation of hydrogen bonds. As mentioned many times above,
this circumstance provided for improved miscibility of the components in the mixture.
However, in addition, phenol groups in the copolymer caused an increase of the
system resistance to thermal oxidation, i.e. these groups acted as antioxidants. This
was expressed in an increase of the induction period of poly(vinyl methyl ether)
thermal oxidation, and the rate of the process decreased. The complex formation
described above was also observed between poly-N,N-dimethylacrylamide and
phenoloformaldehyde resin [218]. The complexes were precipitated from solutions in
acetone, ethyl acetate and dioxane. Therewith, as before, the composition of these
complexes corresponded to the molar ratio of the components. The glass transition
temperature of complexes was higher than the glass transition temperatures of the
original components. Poly(dimethyl acrylamide) also formed complexes with

629

p-methoxyphenoloformaldehyde resin, therewith the glass transition temperature


increases significantly compared with that of each of the components [216, 217].
In ref. [162] were synthesized interpenetrating networks (IPN) based on
miscible polymers poly-1-hydroxy-2,6-methylphenylene and poly(methyl methacrylate). Two types of cross-linked agents were used at various temperatures:
hexamethylene tetramine and 1,3-dioxolane. As indicated in the cited work, the
intensity of hydrogen bonding in mixtures and IPN is determined by variation of the
cross-linking temperature and decrease of the concentration of groups able to form
hydrogen bonds. The concentration of these groups was varied by using methyl
methacrylate copolymer with styrene. Using FTIR-spectroscopy, it was shown that to
preserve miscibility in these mixtures, it is necessary to ensure that the value of
intermolecular interaction, stipulated by hydrogen bonds, is not lower than a definite
threshold value.
Hydrogen bonding influences the phase behavior of mixtures of polymers
[163]. Methyl methacrylate copolymers with styrene and phenoloformaldehyde resin,
in which hydroxyl groups are partially methylated, were chosen as objects for such
investigation. Miscibility diagrams were composed and so-called windows of
miscibility were found where full miscibility and microphase separation were
observed in dependence on temperature and concentration of one of the components.
Ref. [222] studied the miscibility of mixtures of homopolymers with
copolymers. The first pair represented a mixture of poly-4-hydroxystyrene with
n-butyl acrylate copolymer with t-butyl methacrylate. The mixtures mentioned are
miscible, when butyl acrylate concentration in the copolymer is 64% or higher. The
second pair represented poly-t-butyl acrylate and styrene copolymer with 4hydroxystyrene. For this pair, the window of miscibility existed at 4-hydroxystyrene
concentration in copolymer between 28 and 66% (mol). The glass transition
temperature of miscible mixtures of poly-t-butyl acrylate and styrene copolymer with
4-hydroxystyrene is significantly lower than the additive values and, which is of
special interest, it is extremely weakly dependent on the mixture composition, being
approximately equal to the glass transition temperature of copolymers themselves.
Such behavior may be explained only by the formation of hydrogen bonds between
components of the mixture, which is confirmed experimentally by IR-spectroscopy
methods.
Future development of these ideas may be observed in ref. [116], in which
hydrogen bonding of IPN, obtained from a miscible polymeric mixture of poly-1hydroxyl-2,6-methylenephenylene and poly(methyl methacrylate), was studied.
Copolymers of methyl methacrylate and styrene were also used in order to decrease
the amount of carbonyl groups. To obtain IPN, various cross-linking agents were
used, such as hexamethylene tetramine and 1,3-dioxolane. Cross-linking was
performed at various temperatures, by which the intensity of hydrogen bonding in
IPN was varied. The miscible mixture of the two polymers displayed thermal
reversibility in relation to hydrogen bonding. Semi-IPN and IPN prepared at
temperatures above the glass transition temperature of the mixture contained no
initial amount of hydrogen bonds after cooling down to room temperature; but when
poly(methyl methacrylate) copolymer with styrene is used instead of pure
poly(methyl methacrylate) (i.e. when the amount of carbonyl groups is reduced),
semi-IPN did not form a single phase at a significant decrease of the concentration of
these groups. However, semi-IPN and IPN synthesized at comparatively low
temperatures (below the glass transition temperature of the mixture), preserved high

630

level of hydrogen bonds amount compared with the ones, synthesized at high
temperatures, and the single phase [196].
In ref. [186], the authors turned to new objects, such as polybenzimidazoles
and poly-4-vinylpyridine. The main object of investigation was poly-2,2-(mphenylene)-5,5-bibenzimidazole

NH

NH

C
n

All mixtures of different compositions displayed the single glass transition


temperature that testified about their miscibility of components and, therewith, the
glass transition temperature was above the additive values. As in the previous works,
the authors explain this by the formation of hydrogen bonds taking place according to
the scheme:
N

C
N

N
H

)n
The formation of hydrogen bonds is confirmed by experimental FTIRspectroscopy.
The main practical conclusion which can be drawn on the basis of the works
cited is that miscibility can be improved by introduction of a small amount of
functional groups, hydroxyl ones, for example, into one of the components, which
lead to the formation of hydrogen bonds between polymers the mixture components.
Therewith, domains of various size may be formed, and to decrease their size it is
enough, for example, to introduce 4.4% (mol) of hydroxyl groups into polystyrene so
that it would become miscible with poly-n-butyl acrylate. Preliminary investigations
of these regions of relaxation and the size of domains by the NMR method were
performed in ref. [158]. In studies [159, 178, 195, 207, 223], these investigations were
continued.
Complexes based on poly-4-hydroxystyrene and poly-N,N-dimethyl
acrylamide were studied by the NMR 13C method with cross-polarization and sample
rotation at the magic angle. The size of inhomogeneities determined by this method
was found to be equal to 2.5 nm [207]. Further on, this investigation was performed
for IPN. It is found that, in this case, the size of aggregates is smaller than 2.2 nm
[195].
The influence of PMMA microtacticity on its miscibility with a styrene
copolymer with vinyl phenol was studied by the NMR method [159]. It was found
that syndiotactic PMMA is miscible more efficiently with the copolymer mentioned
and formed a single phase in a wide range of mixture compositions based on both
(

CH CH2

631

methyl methacrylates. As indicated by the NMR method, for most composites, the
size of microinhomogeneities is about 2 nm. Investigations of this kind have also been
carried out with poly-4-vinylphenyl dimethylsilanol and its copolymers with styrene
[178, 185]. In the present case, a significant influence of hydrogen bonding between
phenolic and phosphate groups was also observed, as indicated by IR-spectroscopy,
NMR 31P and NMR 13C methods [224].
Studied in ref. [137] were mixtures of poly(ethylene oxide) and poly(butyl
methacrylate) with modified poly(butyl methacrylate) containing groups, able to form
hydrogen bonds [179 184]. Introduction of 4-hydroxy-4,4-bis-3-fluoromethylbutyl
group
CH3
Si

(CH2)3
F3C C

CF3

OH
n
even into a siloxane polymer possessing a low surface energy makes it possible
obtaining mixtures miscible with poly(ethylene oxide) and poly(butyl methacrylate).
Estimated in ref. [151] is the parameter of the polymerpolymer interaction
based on the study of diffusion of water in the mixtures. In all cases, this parameter
became more negative, which indicated better miscibility. Correlations between the
diffusion coefficient and the specific volume of the polymer were found.
It should be noted [194] that functionalization of polystyrene at the sacrifice of
introduction of fluoralkylcarbonyl or hydroxyl groups in it improves the miscibility of
polystyrene with other polymers and increases its thermal stability, oxidation
resistance, and decreases combustibility; this is evident in its mixtures with other
polymers.
The miscibility of a series of polyamides with other polymers was studied
[128, 138, 157]. In ref. [138], poly-p-phenylene terephthalamide sulfonyl was chosen
as the test sample:
HN

NH CO
SO3H

CO
n

Based on its mixtures with poly(vinyl pyrrolidone), poly-4-vinylpyridine and


polyvinyl alcohol, the so-called molecular composites were obtained, in which the
rigid-chain polyamide sulfonyl was the reinforcing element. A significant influence of
hydrogen bonding between elements of the molecular composite was detected.
Therewith, for mixtures with poly(vinyl pyridine) and IPN, the glass transition
temperature was below the additive values that, according to the authors, testified
about a small number of contacts between macromolecules.
An attempt was also made to modify nylon-6 by preparing its mixtures with a
small amount of phenoloformaldehyde resin [157]. Addition of phenoloformaldehyde
resin in amounts of 12% increases the elasticity coefficient and decreases water
absorption. Therewith, larger spherulites were produced. Increase of the
phenoloformaldehyde resin content in the composite induces no improvement of
properties. Consequently, it is concluded that the miscibility of nylon-6 with

632

phenoloformaldehyde resin is rather restricted and falls within the range up to 3% of


phenoloformaldehyde resin. The aim of the work [128] was study of the mechanism
of polymer-analogous reaction between poly(trichlorobutadiene) (PTCB) and
aliphatic diamines: t-butylamine, diethylamine and triethylamine, which model
chemical processes proceeding in separate fragments of polymeric chains during
PTCB interaction with branched poly(ethylene imine).
With the help of FTIR methods and electronic spectroscopy, it was shown that
when poly-1,1,2-trichlorobutadiene interacts with amines, hydrogen-bonded ionic
complexes with the charge transfer are formed accompanying reactions of allyl
chlorine substitution by amino group and dehydration. The contribution of each of
these processes to the total conversion of functional groups of the polymer depends on
amine nature (the N-substitution degree) and the type of solvent. The primary and
secondary amines are more inclined to form stable hydrogen-bonded complexes with
polymers, whereas the tertiary amines cause mostly dehydrochlorination of the
polymer and formation of polyene sequences of different length. Dehydrochlorination
prevails in dioxane, and substitution and complex formation in chloroform.
Hence, for more objective prediction of the miscibility of polymers with each
other, as well as for analyzing the reasons for miscibility, special attention should be
paid not only to the properties of individual components, but also to the specific
intermolecular interaction between them.
As the results of numerous works indicate, there are several ways of
improving the miscibility of polymers:
1. Selection of polymeric pairs or modification of polymers, which are performed for
strong intermolecular interaction occurring between heterogeneous polymeric
chains (for example, hydrogen bonding). This problem will be analyzed in detail
below.
2. The second way concludes in performing chemical reactions between components
of the mixture that finally leads to obtaining of so-called interpolymers, described
in detail in ref. [215].
3. The third way concludes in introducing compatibilizers into the system, i.e. lowmolecular and high-molecular compounds possessing functional groups, which
strengthen the specific intermolecular interaction between chains. Copolymers
may also be used as compatibilizers,.
Let us discuss in more detail the problems associated with strengthening of the
specific intermolecular interaction between mixed polymers.
If such an interaction does not occur, the glass transition temperature of a
homogeneous mixture of miscible polymers is calculated from equation (IV.57). It
contains the glass transition temperatures of homopolymers based on the components
1 and 2, the values of which may be taken as both calculated and experimental for
estimating the glass transition temperature of the mixture. In another form, equation
(IV.57) is reduced to expression (IV.55), which contains sets of constants for the
components 1 and 2, associated with the energy of intermolecular interaction.
Suppose now that hydrogen bonding occurs in the case, polymer 2 is added to
polymer 1. In this case, constant bh = 140103 3K1, which characterizes the

contribution of hydrogen bonds, must be introduced to the value ai Vi + b j


j
i
1
for the component 1. For this case from equation (IV.55), we obtain:

633

1 Vi + 2 Vi
i
1
i
2
Tg =
, (A-5-3)

1 ai Vi + b j + 2 2bh + 2 ai Vi + b j + 21 2 0.03
j
j
i
1
i
2
because a part of repeat units of the polymer 1 appeared bonded to the component 2
by hydrogen bonds. This equation was deduced on the basis of the assumption that
two repeat units are linked by hydrogen bonds, one of them being belonged to the
polymer 1, and the second to the polymer 2. Taking into account equation (IV.57)
and the expression 1 = (1 2), it should be written down that

Tg =

(1 2 ) Vi

Vi
1
(1 2 ) i
Tg,1

+ 2 Vi
i
1
i
2
,

Vi
i
2
+ 2(1 2 ) 2bh + 2
+ 2(1 2 ) 2 0.03
Tg,2

(A-5-4)

where Tg,1 and Tg,2 are the glass transition temperatures (calculated or experimental)
for the components 1 and 2, respectively.
Let us now analyze the dependence of the glass transition temperature on the
mixture composition, reflected by equations (A-5-3) and (A-5-4). Three different
cases will be considered:
1) The Van-der-Waals volumes of repeat unit of polymers 1 and 2 are approximately

equal: Vi Vi .
i
1 i
2
2) The Van-der-Waals volume of repeat unit of the polymer 1 is significantly lower

than for the polymer 2: Vi << Vi .


i
1
i
2
3) The Van-der-Waals volume of repeat unit of the polymer 1 is significantly higher

than for the polymer 2: Vi >> Vi .


i
1
i
2

Initial values of Vi , Vi , Tg,1, Tg,2, ai Vi + b j , and


j
i
1 i
2
i
1

ai Vi + b j for the mixtures corresponded to all three cases mentioned are


i

2
shown in Table A-5-1. Indicated in Figure A-5-1,a are dependences of the glass
transition temperature of mixtures on the content of component 2 for the case, when

Vi Vi . Therewith, these values vary in a wide range. When the Van i


1 i
2

der-Waals volume of the repeat units is low, the value of ai Vi + b j is also


j
i
2

634

low, therewith, the cases are considered when the glass transition temperature is
practically the same for both components of the mixture. Since the value bh, which
characterizes the influence of hydrogen bonds on Tg, is quite high, it may be observed
from equations (A-5-3) and (A-5-4) that the smaller is the Van-der-Waals volume, the
more the glass transition temperature exceeds the value appropriate to the usual case,
at which additional specific interaction between components is absent. This affects the
run of Tg dependences on 2, indicated in Figure A-5-1,a. When the Van-der-Waals
volume of repeat units of both components increases, the effect of additional
interaction decreases, and the glass transition temperature of the mixture is just a little
higher than the value of Tg calculated by equations (IV.55) and (IV.57).
Table A-5-1
Initial values of parameters for calculation of the glass transition temperature of miscible
mixtures of homopolymers

Number of
Vi ,
Vi ,
, L 'L +
-M ,
, L 'L +
-M ,

Figure
curve in

i
1
i
2
M
M
L

L

Figure A-5-1
3
3
3
1
3
1

K
K
260
293
115
110
A-5-1a
1
588
665
260
250
2
814
931
360
350
3
1041
1197
460
450
4
543
279
240
105
A-5-1b
1
769
279
340
105
2
995
279
440
105
3
1222
279
540
105
4
260
598
115
225
A-5-1c
1
260
904
115
340
2
260
1170
115
440
3
260
1436
115
540
4
Note: Tg,1 = 376 K; Tg,2 = 442 K.

Figure A-5-1,b is appropriate to the case, when Vi << Vi . Here


i
1
i
2
the glass transition temperature is significantly higher than the value of Tg calculated
by expression (A-5-3). It should be noted that even if additional hydrogen bonding is
absent, the glass transition temperature of the mixture exceeds the value of Tg
calculated on the basis of the molar parts of components.

Figure A-5-1,c describes the situation when Vi >> Vi . In this


i
1
i
2
case, influence of the additional hydrogen bonding decreases, because the total energy
of dispersion interaction significantly exceeds the energy of hydrogen bonds. Such
situation also appears in the cases, when the excessive intermolecular interaction
between components at their mixing is the dipoledipole interaction, occurred owing
to the presence of polar groups in the repeat unit. In this case, the value bh in
equations (A-5-3) and (A-5-4) is substituted by the value bd:

635

Figure A-5-1. Dependence of the glass transition temperature Tg on concentration of the second
component 2 for miscible mixtures of two polymers when formation of hydrogen

bonding between them takes place: a) Vi Vi ; b) Vi <<

i
1 i
2
i
1

Vi ; c) Vi >> Vi ; numbers on curves correspond to the


i
2
i
1
i
2

numbers in the curves shown in Table A-5-1. The curves 1 4 correspond to the
situation when no additional hydrogen bonding between the components takes place.

636

1 Vi + 2 Vi

i
1
i
2
Tg =
;

1 ai Vi + b j + 2 2 bd + 2 ai Vi + b j + 21 2 0.03
j
j
i
1
i
2

Tg =

+ 2 Vi

i
1
i
2
.

V
i
i
2
+ 2(1 2 ) 2 0.03
+ 2(1 2 ) 2 bd + 2
Tg,2

(A-5-5)

(1 2 ) Vi

V
i

1
(1 2 ) i
Tg,1

(A-5-6)

When Vi Vi , the influence of additional intermolecular


i
1
i
2
interaction on the dependence of Tg on the composition is also displayed, but it
appears significantly lower than at occurrence of hydrogen bonds between
components (Figure A-5-2,a). The reason is that the value bd, characterizing influence
of the dipoledipole interaction on the glass transition temperature, is lower by the
absolute value than the value bh.
If the Van-der-Waals volume of the component 2 is significantly higher than
that of the component 1, positive deviation of Tg values from the average values also
takes place, however, the effect is not so high as in the case of occurrence of hydrogen
bonds (Figure A-5-2,b). If the Van-der-Waals volume of the repeat unit of the
component 1 is much higher than of the component 2, the influence of additional
intermolecular interaction is not so high, and dependences of the glass transition
temperatures on the composition are quite close to those calculated by equation
(IV.55), i.e. with no regard to additional intermolecular interaction between
components. This is clearly indicated in Figure A-5-2,c.
It is common knowledge that besides the dependences of glass transition
temperatures on the composition of the mixtures with maxima (curve 1) and minima
(curve 2), the S-shaped dependences are encountered quite often (curve 3) (Figure A5-3). This is associated with the fact that one of the components of the mixture
represents a copolymer containing a limited number of polar groups, able to form
hydrogen bonds. Let us consider this situation in more detail. Suppose that a
copolymer contains a limited amount of repeat units capable of hydrogen bonding
(copolymer 2). For example, copolymer 2 may be presented in the form:
(CH2

CH)m

(*)

(CH2

CH)n

OH
(**)

637

Figure A-5-2. Similar to the situation in Figure A-5-1 but with occurrence of additional dipoledipole
interaction between components.

638

Let us designate the molar part of the repeat unit (**) as . Suppose that the
counter-polymer (polymer 1) is able to form hydrogen bonds with the repeat unit (**),
contained in the copolymer 2. Applying these designations, equations (A-5-3) and (A5-4) are reduced to the following forms:

Tg =

(1 2 ) Vi

(1 2 ) ai Vi + b j + 2(1 2 ) 2 2 2 bh +
j
i
1

+ 2 Vi

i
2

+ 2 ai Vi + b j + 2(1 2 ) 2 0.03

j
i
2

(  ) '

%J =
( 

'

)
L

%J

(A-5-7)

+ (  )    - + 

'

%J

(A-5-8)

+  '

+ (  )  
L

Figure A-5-3. Schematic representation of typical deviation of the glass transition temperature from its
average values for the mixtures of miscible polymers (see text).

As above, let us discuss a series of cases:


1) The Van-der-Waals volume of the repeat unit is approximately the same.
Initial values of all parameters necessary for the calculation are shown in
Table A-5-2. During calculation, let us vary content of the component (**) for the
copolymer 2. Calculation results are shown in Figure A-5-4,a. Clearly, the part of the

639

components (**) able to form hydrogen bonds is quite moderate and, therewith,
S-shaped dependences of the glass transition temperature of the mixtures on the
composition are formed. These dependences were observed in many works (refs.
[180, 192, 193, 209]).
Table A-5-2
Initial values of parameters for calculating the glass transition temperature of miscible (Tg,1 =
376 K, Tg,2 = 442 K) mixtures of homopolymer (1) and copolymer (2)

, 3K1

, 3K1
V , 3
V , 3
,
'
,
'

+
i
i

L
L
L
L
M
M

i
1
i
2
M
M
L

L


110

115

293

260

Figure A-5-4. Dependence of the glass transition temperature Tg on concentration of the second
component 2 for miscible mixtures of homopolymer and copolymer (see text) when:
a) hydrogen bonding between components exists; b) dipoledipole interaction between
components prevail.

2) This case corresponds to formation of strong dipoledipole interactions


between copolymer 2 and polymer 1.

640

For calculation of the glass transition temperature, appropriate equations are


deduced from equations (A-5-7) and (A-5-8):

 ) ' +


%J =

(  ) , ' + - + (  )    - +

+  '




+  , ' + - + (  )  

(  ) ' +

'
'



+
+ (  )    - + 
(  )

(A-5-9)

%J

%J

%J

(A-5-10)

+  '


+ (  )  
L

The calculation results obtained by equations (A-5-9) and (A-5-10) are


indicated in Figure A-5-4,b. Calculated dependences of Tg on the part of copolymer 2
were obtained under various parts of the component (**) in the copolymer. It should
be noted that occurrence of the S-shape of Tg dependence on the composition is
displayed only in the case, when the copolymer 2 contains a significant part of the
component (**). However, in all the cases, these dependences differ from these
obtained with no regard to occurrence of an additional dipoledipole interaction
between components of the mixture. In relation to variations of the glass transition
temperatures occurred as a result of excessive intermolecular interaction between
components, the calculated data indicate that these interactions are of a significant
importance. The values of Tg differ from those, when no additional intermolecular
interaction occurs. Note also the important circumstance that all equations (A-5-3)
(A-5-10) applied to the analysis of the glass transition temperature dependence on the
mixture composition displays no one fitting parameter and, hence, all calculations
are performed on the basis of the chemical structure of the mixture components.
Let us consider behavior of a series of particular mixtures of two miscible
polymers. The first of them represents a mixture of poly-n-butyl methacrylate
(PBMA) with styrene copolymer with 4-vinylphenyl dimethylsilanol (St-VPDMS)
[180]:

641

(CH2

CH)m

H3C

(CH2

CH)n

Si CH3
OH

Firstly, it is necessary to calculate glass transition temperatures for St-VPDMS


copolymer at different concentration of VPDMS. Such calculations were performed
by equation (IV.55), and their results are presented in Table A-5-3. Good coincidence
of calculated and experimental data is observed. For further calculations of glass
transition temperatures of PBMA mixtures with the given copolymer, it is necessary

to know values of , ' + -


and , ' + -
determined

!

$9
'!$
from equation (IV.55). It is also necessary to determine the same values for
copolymers of all compositions. These values are also shown in Table A-5-3. They
were used for future calculation of the glass transition temperature of a series of
PBMA and St-VPDMS mixtures of different compositions. Indicated in Figure A-5-5
are the calculation results for two mixtures in the form of dependences of the glass
transition temperature on the composition. These dependences for miscible mixtures
containing St-VPDMS-34 and St-VPDMS-9 were obtained by equation (A-5-7). For
every mixture, quite good coincidence with the experimental data is observed (see
Figure A-5-5,a and b).
L

Table A-5-3
Parameters of copolymers of styrene with vinylphenyldimethylsilanol
Tg (exp.), C
Designations
The part of
Tg (calc.), C

Vi ,

VPDMS,
, L 'L +
-M
3

% (mol)
M

3
3 -1
10 , K
PS
100
105
103
110
293
PBMA
100
29
19
148
516
PVPDMS-2
1.9
97
103
111
295
PVPDMS-4
4.1
98
103
113
301
PVPDMS-9
8.7
98
103
116
309
PVPDMS-P
11.4
99
103
117
311
PVPDMS-18
18.2
101
104
122
324
PVPDMS-34
33.9
115
109
133
348
PVPDMS-60
60.0
121
122
151
382
PVPDMS-100
100
153
178
418

Moreover, let us analyze dependence of Tg on composition for PBMA mixture


with St-VPDMS-9 with the help o equation (A-5-8). Therewith, let us also introduce
experimental data into this equation. Consequently, we obtain a good coincidence of
calculated and experimental data (see Figure A-5-5,c). Note that glass transition
temperatures of mixtures are below the average values, when concentration of
modified styrene (i.e. St-VPDMS) in the mixture is low. In the case, when
concentration of this component in the mixture increases, glass transition

642

temperatures exceed the average values. Such exceeding is explained by formation of


a great amount of hydrogen bonds between components at their mixing.

Figure A-5-5. Dependence of the glass transition temperature Tg on the mole fraction of copolymer 2
for miscible mixtures of PS with VPDMS/PBMA: a) PVPDMS-34; b) PVPDMS-9. 1
calculated curve; 2 experimental dependence; 3 linear dependence.

Let us now consider behavior of another mixture obtained at the mixing of


poly-(2,2-m-phenylene-5,5-bibenzimidazol) (PBI):
N

NH

NH

and poly-4-vinylpyridine:

643

[CH2

CH]

N
These mixtures were studied in ref. [186]. It has been indicated that hydrogen
bonds are formed between residues of 4-vinylpyridine and NH-groups in PBI. This is
the reason that homogeneous mixtures based on these two polymers are formed.
Experimental dependences of the glass transition temperatures on the composition of
these mixtures, taken from ref. [186], are indicated in Figure A-5-6. Glass transition
temperatures of the mixtures exceed values of Tg for the original components.
Calculated dependences of Tg on composition, determined by equation (A-5-8), are
also indicated in Figure A-5-6. It is obvious from this figure that experimental points
fit well the calculated curve, when the molar fraction of PBI in the mixture does not
exceed 0.3. Deviation of the experimental points from the calculated curve at high
concentration of PBI is associated with the onset of poly-4-vinylpyridine degradation
at heating over 375. Note one more time that all calculations are performed by the
equations containing no fitting parameters. Hence, all the results are obtained only on
the base of chemical structure of the components of the mixture.

Figure A-5-6. Dependence of the glass transition temperature Tg on the composition 2 for miscible
mixtures of poly(vinyl pyridine) with poly(benzimidazole). 1 calculated curve; 2
experimental dependence; 3 linear dependence.

Let us make some conclusions from the above analysis. In the case of miscible
mixtures of two homopolymers, the dependence of the glass transition temperature on
the composition of the mixture may obtain different forms /shapes/. If the Van-derWaals volumes of the repeat unit of homopolymers are approximately equal, and no
additional strong intermolecular interaction occurs between the components, the glass
transition temperature falls below the average values. If the Van-der-Waals volume of

644

the unit of homopolymer 1 is significantly smaller than that of homopolymer 2,


positive deviations of the glass transition temperature from the average value take
place even if the additional intermolecular interactions are absent. But if such
interactions take place, a positive deviation of Tg values from the average values
always takes place.
In the case of homopolymer mixtures with a copolymer containing a limited
amount of polar groups, capable for strong intermolecular interaction with other
components (dipoledipole interaction, hydrogen bonds), dependences of Tg on the
composition are S-shaped. All the mentioned features of behavior of such miscible
mixtures may be described with the help of the equations mentioned above, which
allow estimation of the properties of the mixtures based on the chemical structure of
the components. It is self-evident that the present approach cannot pretend to be
general, and more general investigations are required to take into account the
influence of various specific interactions and fine features of the structure of
components of mixtures on their thermomechanical behavior.

Appendix 6. On formation of super-molecular structures in


amorphous polymers

1 Scheme of formation of the super-molecular structure


The experimental data obtained up to the present time, as well as theoretical
investigations of polymers enable us to appoint a scheme of the formation of supermolecular structures and, yielding from this scheme, to make an attempt in future to
set a connection between the chemical structure of a polymer, its super-molecular
structure and physical properties [236]. The question about super-molecular structures
and their forms was originally formulated in ref. [237]. Further on, various models of
super-molecular structure appeared [238, 239], which allow an explanation of a series
of properties of amorphous polymers in solution, melt and solid state. Simultaneously
with model ideas, numerous experiments with obtaining super-molecular structure by
various methods were performed, with the method of electron microscopy being the
chief method. The most typical situation in amorphous polymers observed is the
picture of globules the sizes of which correspond to aggregates of tens and hundreds
of macromolecules.
Let us dwell on problems of super-molecular structure formation and
description of this process with the help of an analytical apparatus, developed in refs.
[240 242]. Before passing to the essence of the problem, let us note in brief the main
results of these works, which will be necessary for us for future discussion. Suggested
by Lifshitz in ref. [240] was the equation:
g = exp{*(n)/RT},
n = 2exp{*(n)/RT},
N = nd 3 r ,

(A-6-1)
(A-6-2)

where g is the integral operator; g = g (x x) (x)d 3 x ; g(x x) is the

conditional probability of existence of two units in points x and x, respectively; and


are the characteristic function and the characteristic value of equation (A-6-1),
respectively; n(x) is the function of the smoothed density which defines the
macroscopic state of the polymeric globule; * is the chemical potential of the system
of broken chains; N is the total number of units of the macromolecule. Besides
description of the coil globule transition, this equation allows analysis of the
structure of various physical factors (temperature, solvent quality, interaction with the
solvent). Basing on the analysis of solutions of equation (A-6-1), a phase diagram of
the system of interacting particles of the polymeric macromolecule in solution was
composed (Figure A-6-1), which, further on, will be widely used for explaining the
formation of the super-molecular structure.
As the starting state, consider a diluted solution of a polymer, in which
overlapping of coils is absent, and a monomacromolecular transition is possible. It
may be either a diluted solution, specially prepared for this purpose, or a solution of
propagating chains of the polymer in the medium in which synthesis proceeds. The
synthesis may proceed both in the monomer medium and in the medium of a specially

646

selected solvent (then, in the first case, the amount of macromolecules formed at the
onset of the reaction will be low, and the solution may be considered diluted). The
solution temperature T will be lower than some critical temperature Tc (which will be
determined below), but is greater than the triple point T (Figure A-6-1). Then,
macromolecules with molecular mass above the BA line (Figure A-6-1) will transit
into the globular state with a dense center and loose periphery, and macromolecules
with the molecular weight below the BA line will remain in the coil state.

Figure A-6-1. Phase diagram of a system of interacting particles (beads) of the chain in solution. EDC
lines separate the area of globule without periphery, EDBA globules with periphery,
CBA the coil. T is the temperature of the triple point; Tcr is the critical temperature of
broken chains.
Figure A-6-2. Scheme of formation of the initial element of super-molecular structure: 1, 2 centers of
globules of macromolecules; 3, 4 peripheries of globules of macromolecules; 5 bond
globule (or coil).

To solve the task set, let us limit by the volumetric approximation when the
density of a macromolecule at the nucleusperiphery interface changes jump-like to
zero (periphery density). Let us discuss an interaction of two such globules with the
periphery when they approach each other due to diffusion motion or sedimentation
(interaction of coils is out of interest for us, because in this case either globules with a
periphery may appear, if the total number of units in coils will appear above the coil
globule transition on the phase diagram, or a coil with higher molecular weight is
formed, i.e. the problem is reduced to the same cases, which will be discussed below).
When peripheries of globules of two macromolecules interpenetrate, units of
peripheries will interact with each other. Therewith, the following variants are
possible. If the total number of units in peripheries under given temperature is above
the coilglobule phase transition line (Figure A-6-1), a bond globule with periphery is
formed (Figure A-6-2). If the total number of units in peripheries is low and appears
below the phase transition line, a bond coil is formed at the capture of periphery units
of one globule by periphery units of another one, which at a future decrease of
temperature transits either to the bond globule with the periphery or to the bond
globule without the periphery (Figure A-6-1).
Hence, in any case, the initial element forming the super-molecular structure
will be two globules connected with each other by a globule or a bond coil.
Depending on the size of the globule center and bond globule (or coil), these elements
of super-molecular structure in the limit form two most typical types of elements of
the super-molecular structure.

647

If a bond globule (coil) is small compared with nucleus of globules of two


initial macromolecules, i.e. r << R (Figure A-6-2), an element of the super-molecular
structure of the dumbbell type is formed. If a bond globule (coil) is comparable with
globules of original macromolecules, i.e. r ~ R (Figure A-6-2), an element of the
super-molecular structure of the ellipsoid type is formed. The rest of elements of the
super-molecular structure will be of an intermediate type between the ellipsoid and
the dumbbell. When the element of the super-molecular structure formed interacts
with each other or with globules of macromolecules, new bond globules or coils
appear, which lead to formation of super-macromolecules composed of globules
macromolecules.
Depending upon the size of bond globules (coils), these super-macromolecules
will possess different mobility. In the case of small bond globules (coils), their
mobility will be maximal. Consequently, the super-macromolecule forms a giant
globule representing the final element of the super-molecular structure. In the case
when bond globules are large (the original element of the super-molecular structure,
on average, approaches the ellipsoid), the mobility of separate elements of the supermacromolecule will be low, and a drawn fibrillar super-molecular structure is
obtained. The size of the super-molecular structure will be defined by the length of
the super-macromolecule.
Let us estimate the length of a segment1 of the super-macromolecule in the
case of a fibrillar super-molecular structure (in the case of the globular supermolecular structure, the size of the segment equals the size of the globulemacromolecule). Making estimations, we will proceed from the assumption that the
sizes of super-macromolecular chains of globular and fibrillar super-molecular
structures are equal. The number of globulesmacromolecules by radius nR and length
nL of cylinder of the fibrillar structure may be determined from the condition:
2
Vfib = Rfib
Lfib , Rfib = nRRmol, Lfib = 2nLRmol, Vglob =

3
3 Rmol n ,

where n is the total number of globulesmacromolecules in the final element of the


super-molecular structure. From the condition Vfib = Vglob, it follows that
2nL nR2 = 4 3 n .

(A-6-3)

Let us deduce the second condition for nL considering a super-macromolecule,


which forms the fibrillar super-molecular structure, as the PorodKratki chain.
Suggest that a/L << 1, where a and L are sizes of the segment and the whole supermacromolecule, respectively. Then, according to [243], the length of the Porod

( )

Kratki chain will be h 2 = 2La, where L = 2nRmol, a = Rmol, and h 2

1/ 2

=2nLRmol.

Then the condition for h may change to:


2nL =

2 n .

(A-6-4)

For the segment of super-macromolecule, we assume its part, the ends of which may travel
independently, i.e. this is the same definition as for chains of macromolecules, but in relation to
superchains.

648

From conditions (A-6-3) and (A-6-4), it follows that = 8n 9nR4 . The number
of globulesmacromolecules n will be calculated on the base of the size of the supermolecular structure globule. The calculations will be performed on the example of
polyarylate F-2, for which the electron-microscopic image of the film, taken from
refs. [4, 244], is displayed in Figure A-6-3. Let us determine the number of
macromolecules in super-globules. As indicated in Figure (A-6-3), super-globules are
principally of two types comparatively small and homogeneous, and larger. The
former possess the radius Rcr = 125 , and the latter ~5001200 . Polyarylate
studied has M = 40 000, and polymerization degree is equal to 90. The Van-der-Waals
volume of the repeat unit Vi is 390.2 2; taking into account that the coefficient
i

of molecular packing in the film is ~0.695, the volume of a single unit is 390.2/
0.695 = 561 3. The volume of the macromolecule consisting of 90 units equals
50 860 3. At Rcr = 125 the number of macromolecules forming the super-globule
will be ~150, and at Rcr = 500 it is ~10 000.

Figure A-6-3. Electronmicroscopic image of polyarylate F-2 film according to data from refs.
[4, 244].

The mean radius of the globule, Dglob, determined by the microphotography


method [4, 244], is ~800 . The number n appropriate to it at Rmol ~ 20 is
n ~ 8 000. The value nR will be determined on the basis of the mean diameter of fibril
Dfib ~ 192 ; nR ~ 5. Then 11. From this it is clear that the transversal size of the
fibril will be similar by order of magnitude of the segment length of the supermacromolecule. The latter result is quite natural, because the condition a/L << 1 is
true at 11, and in this case, Dfib ~ a.
In accordance with the above-discussed ideas, the fibrillar or globular
structure may be formed by selecting the molecular mass distribution, solvent quality,
temperature solution and rigidity of molecules. To analyze the influence of the abovecounted factors on the type of super-molecular structure separately, let us consider
changing of the periphery thickness depending on the properties of macromolecules
and their solutions. According to refs. [240, 241], when passing the coilglobule
transition, the dependence of the globule density n on the coordinate x at T > Tc will
obtain the form indicated in Figure A-6-4. As volumetric approximation will be
considered further, the dependence for the periphery thickness obtained below may be
also used in the temperature range T < T Tc. Considering the globule consisting of

649

a dense center with density nn and a periphery, let us determine the periphery
thickness as the area where density changes linearly with distance x, counted from the
dn
x , where Rn is the radius of the central zone. As
center border. Then, n(x) = nn
dx Rn
nn is known, and the coil density equals zero, then to determine the periphery,

dn
dx Rn

must be found. From expression (A-6-2), it may be found that


dn
d ln
2n
=

.
n d * dx
dx

1
RT dn

Figure A-6-4. Dependence of the globule density n on coordinate x (volumetric approximation is


cross-hatched).

For the derivative dn/dx to be independent on x, it must be taken ~ ex.


Then,
dn
2n
=
.
n d *
dx Rn

1
RT dn
Let us consider the initial equation (A-6-1). For the condition ~ ex to be
met in the area of the periphery, it is necessary that exp{*(n)/RT} = const. For the
d2
area of periphery, let us present the operator g in the form g 1 + a 2 2 , as it was
dx
made in ref. [240]. Then equation (A-6-1) will obtain the form:
a2

d 2
const 1
+ (1 const ) = 0 , =
2
dx
a2

and
dn
const 1
= 2n( Rn ) = 2n( Rn )
.
dx
a2

650

The dependence of the periphery density on coordinate x will be expressed as


follows:
2n( Rn )
n ( x ) = nn
const 1x .
a
The periphery thickness 2 l will be found from the conditions n(Rn) = nn/2 and n(2l) =
0, and will be equal
2l =

a
.
const 1

(A-6-5)

In the volumetric approximation, when profile of the globule is substituted by a jog,


the periphery thickness will be equal to l.
Let us estimate const in expression (A-6-5) in the volumetric approximation.
B2
As in the volumetric approximation, according to [242], = 1 +
, where B and C
4C
are the second and the third virial coefficients in the equation of the globule state, at
-temperature the periphery size is l = a / 2 const 1 . The size of the globule Rglob is
composed of the center size Rn and the periphery size. That is why,
Rglob = Rn +

a
.
2 const 1

When T = , the globule transforms into a coil, Rn = 0, and 2Rglob = Na 2 , i.e. the
square size of the globule equals to the mean square of the coil size (amplitude of
fluctuations). That is why, const = 1 + 1/N. Hence, when T = , free energy and
density of the globule, according to [242], will be equal to zero, and the periphery
thickness 2l = N1/2a. When T < , the value 1/N may be neglected, and const 1.
N 1/ 2 a aC 1/ 2

When T < , the size of the center will be Rn =


, and its density is
b(T )
2
3

1/ 2

N

4 3 1/ 3 C
nn =
= 1 a N 1 +
,
4 R 3
n
3
3
N b( T )

i.e. the density of the center increases as temperature decreases.


To analyze the influence of temperature, solvent quality, and chain rigidity on
the periphery thickness and, respectively, on the formation of the super-molecular
structure, let us consider expression (A-6-5) at T < . The problem of the sequence of
effects (heatingcooling, rate of solvent elimination, etc.) on the system and hysteresis
phenomena associated with it will not be discussed. As in this temperature range
const 1, nglob = B/2C (according to ref. [242]), the expression for the periphery
thickness may obtain the form:
l=

a
.
2nglob B

(A-6-6)

651

Rigid-chain polymers possessing large value of the segment, a, will be


characterized by extended periphery l, and all other factors being the same, the
fibrillar structure will be typical of them. In the limit, when the segment becomes
equal to the macromolecule by length, and the latter obtains a stick-like form, the idea
of the center becomes senseless, and the super-molecular structure may be of the
fibrillar type only. The border of rigidity, out of which globules will not be already
formed, may be determined from the phase diagram (Figure A-6-1) for the equivalent
chain, macroscopic characteristics of which coincide with the characteristics of the
original macromolecule [245].
Let us now estimate the influence of the solvent quality, all other factors being
the same (constant concentration and temperature). As virial coefficients A2 are used
in studying solutions of polymers, determined from the concentration dependence of
osmotic pressure and light scattering, to analyze the influence of the solvent quality, it
will be assumed that A2 ~ B. For the periphery to form, according to expression
(A-6-6), the value of B must be negative, i.e. a bad solvent is favorable for forming
the periphery. The worse is the thermodynamic quality of the solvent (the absolute
value of B is higher), the smaller is the periphery size. If the solvent quality is
significantly worse, the size of the periphery will be minimal and polymer will be
precipitated out as compact globules without a periphery.
The problem of formation of a super-molecular structure in a polymeric
substance obtained from the solution without precipitation of the polymer will be
largely of interest for us. It follows from expression (A-6-6) that when the film is cast
from a bad solvent (large absolute value of B), globules with periphery l of small size
will be formed. This, in its turn, will lead to formation of a globular super-molecular
structure, and the sample will possess increased friability. Experimentally, this fact is
confirmed well. When the quality of the solvent is increased (|B| 0), the size of
periphery will increase, and the probability of formation of a fibrillar structure
increases. Therewith, samples possess better ability to deformation. When the
thermodynamic quality of the solvent is increased significantly, the value of B
becomes positive, and then the globule with center and periphery is not formed at all,
and at formation of a solid macromolecules remain in the form of coils. In this
connection, the experimental data become clear; according to these data, the addition
of a small amount of a polymer precipitator to the solution improves the mechanical
properties of films.
Let us now discuss the influence of temperature on the character of formation
of the super-molecular structure. If the absolute value of the second negative virial
coefficient B increases as temperature decreases, then the quality of the solvent
deteriorates; such solution has the upper critical temperature of dissolution.
Therewith, according to correlation (A-6-6), the size of the periphery l decreases, and
a compact globular structure will be formed during the formation of a film from the
solution. This will lead to deterioration of the mechanical properties of the film. In the
case when the absolute value of the second negative virial coefficient B increases with
temperature, the quality of the solvent deteriorates; such solvent displays the lower
critical temperature of dissolution. According to expression (A-6-6), in this case, the
size of the periphery l will also be low, which will lead to the formation of a compact
globular super-molecular structure. Consequently, in the first case, to eliminate
formation of the globular super-molecular structure and to obtain globules with
extended periphery, able to form bond globules and formation of a fibrillar structure,
the film must be cast under increased temperature. In the second case, the film must

652

be cast at reduced temperature. Then, according to expression (A-6-6), the absolute


value of B will be lower, and the size of periphery l will be greater.
It is common knowledge that when the transitions occurs (glass transition
temperature, melting point and temperature of degradation), the super-molecular
structures weakly affect the appropriate critical temperatures. However, if the
problems of plasticization and transfer (diffusion, conductivity, etc.) are considered,
the contribution of the super-molecular structure may appear decisive. For example,
let us consider plasticization of polymer. According to existing notions, plasticization
is subdivided into two types: intrastructural one (by Zhurkov and KarginMalinski)
and interstructural one (by Kozlov). Reasoning from the above-developed notions on
the super-molecular structure, plasticization via the globule center corresponds to the
intrastructural plasticization, when the plasticizer penetrating into the globule center,
changes the temperature of the coilglobule transition (more accurately, the glass
transition temperature). Therewith, the glass transition temperature will change
smoothly with the concentration of the plasticizer, because in most cases (especially,
in globular super-molecular structures), the volume of the globule center significantly
exceeds the periphery volume.
In the case of interstructural plasticization, the process of plasticization
proceeds at the sacrifice of peripheries, because the plasticizer does not penetrate
inside the center. Therewith, peripheries of those globules will be saturated by the
plasticizers, which are located on the surface of superglobules. As the volume of these
peripheries represents a small part of the volume of the superglobule center, saturation
of the periphery by a plasticizer will proceed faster than in the case of intrastructural
plasticization, which is observed experimentally and is indicated by an abrupt
decrease of the glass transition temperature.
The mechanism of interstructural plasticization suggested enables plasticization and anti-plasticization phenomena to be explained. As the elasticity modulus
decreases during plasticization at the given temperature, within the framework of the
model of the super-molecular structure considered this event may be explained by the
fact that the periphery extends under the effect of the plasticizer, and the contribution
of the more dense center to the elastic properties decreases. When anti-plasticization
proceeds, the periphery thickness decreases first, contribution of the center to the
elastic properties increases, and the elasticity modulus increases. Further increase of
the plasticizer concentration indicates the same effect of it as in the case of
plasticization.
Doping of polymer (for example, polyacetylene by doping components I2 and
AsF5) is, apparently, analogous to interstructural plasticization. That is why in spite of
the low concentration of the dopant components in proportion to the total volume of
the polymer, 100% doping of the periphery appears, and the total conductivity of the
polymer will be determined by the conductivity of I2 and AsF5 vapors multiplied by
the concentration of these components in the polymer (because the doped periphery
represents a mixture of gases of units and doping components). Therewith, transition
to the dielectric state will proceed due to collapse of the periphery. Consequently,
atoms of doping components appear disseminated in the solid matrix. Due to the same
reasons, the diffusion coefficient of the polymer will be equal to the diffusion
coefficient of gas of the penetrating substance multiplied by the concentration of
peripheries. When temperature decreases, the coefficient of diffusion similar to
conductivity must display a minimum in the temperature range of the metaldielectric
transition.

653

A diluted solution at temperatures below the -temperature was considered


below as the onset state for forming a super-molecular structure. Therewith, a
significant role affecting the stability of the super-molecular structure as the polymer
concentration increases is played by surface tension, appearing in both the globule
itself and in the solvent affecting the globule after the coilglobule transition. If a
good solvent at temperatures above the -point is taken for the onset state, formation
of the super-molecular structure is rather problematic and requires further
consideration with regard to the data obtained in ref. [246].
It should be noted that the majority of super-molecular structures were
obtained from solutions of polymers at room temperature, whereas their
-temperatures exceed the room temperature by 1020 and more [247]. That is why
the mechanism of formation of a super-molecular structure discussed above may be
assumed as quite a typical one of polymers. As shown below, the mechanism of
formation of a super-molecular structure suggested may be used for explaining and
predicting the solubility of polymers.

2 Calculation method of evaluation of dimensions of elements of


super-molecular structure of polymers
The physical characteristics of polymers calculated by the approach discussed
in the monograph may be subdivided conditionally into two types. The characteristics
mainly determined by the chemical structure of the repeat unit relate to the first type.
These characteristics are the glass transition temperature, temperature of the onset of
intense thermal degradation, the refractive index, density, etc. The second type
embraces the characteristics not only by the chemical structure of the repeat unit, but
also by the geometric size of elements of the supermolecular structure. These are the
elasticity modulus, solubility, conductivity, etc. To calculate these characteristics, one
must know how to determine the geometric size of elements of the super-molecular
structure in relation to conditions of molding of the polymeric sample. In the case of
casting film samples, the super-molecular structure will depend on the chemical
structure of the solvent and the polymer, the temperature of solvent evaporation, the
rate of evaporation, adding of precipitators, etc. Before we turn to discussion of the
analysis results, several general notes must be made.
Due to its stoichiometric configuration, not every system of macromolecules is
capable of spontaneous formation of either a globular or a fibrillar structure. If we
start preparation of the sample from a solution, then the flexible chains in static
conditions counteract in size and in the corresponding region of scaling diagrams of
state the interaction of the blocks may lead to the formation of globular density
agglomerations. Semi-rigid chains, vice versa, strive to straighten as the concentration
increases, which will lead to the formation of a fibrillar super-molecular structure.
Moreover, flexible molecules in a straightening field (whatsoever in the presence of
strong gradients of rate or force) may behave themselves as rigid ones and form a
fibrillar super-molecular structure.
Selection of a bad or good solvent determines both thermodynamic and kinetic
factors. Therewith, the rate of solvent elimination (or precipitator injection) may play
the decisive role, because the system of macromolecules must manage to rearrange
itself to one or another side. Moreover, it is common knowledge that amphiphilic or
diphilic macromolecules may interact in parts separately with the precipitator. For
example [254], when poly(methyl methacrylate) is precipitated by methanol, a fine

654

suspension is formed, drying of which leads to a typically granulated (globular)


structure, because the precipitator solvates ester groups, but causes a collapse of nonpolar backbone of the chains. In contrast, saturated hydrocarbons as precipitators lead
to the formation of a continuous gel, because they precipitate, in the strict sense, only
polar groups, but not the backbone. External fields and the regularity of chains (as
well as the absolute values of the size of the repeat unit) contribute to the kinetics of
structure formation.
Following the results from ref. [94], the special attention in this Appendix will
be paid to the thermodynamic aspect of the problem of formation of super-molecular
structure proceeding at casting of a polymeric substance from solution.
Performed in ref. [255] are calculations on the determination of sizes of superstructures appearing in block-copolymers. Analogous types of elements of supermolecular structure may also be observed in homopolymers (according to technology
in ref. [255], lamellar; fibrillar, appropriate to the cylindrical one; globular,
appropriate to the spherical one). Let indicate that the original physical assumptions
used in calculating the size of super-structures in block-copolymers are maintained at
formation of the super-molecular structure according to the scheme discussed in ref.
[92]. That is why correlations from ref. [255], appropriately corrected, may be applied
to the determination of the size of elements of the super-molecular structure. In spite
of ref. [255], discussed below are super-molecular structures in the amorphous state of
the homopolymer.
The physical base of structure formation in block-copolymers is immiscibility
of blocks, their tendency to segregation. Therewith, morphology is determined by the
ratio of the molecular mass of blocks. In the case of forming a super-molecular
structure in homopolymers, it is based on super-macromolecules, formed from
globules macromolecules existing below the coil globule transition temperature2,
Tcr [92]. Below Tcr, the polymer and the solvent become immiscible, and segregation
appears (the globule presses out the solvent). We are interested in the result of
segregation at decrease of the solvent concentration, as well as in the influence of the
molecular mass on the results obtained.
When determining sizes of elements of the super-molecular structure, let
consider the final stage of appearance of the super-molecular structure, when the
concentration (volumetric part) of the solvent, cs ~ 0, and the concentration of
supermolecules, csm ~ 1. Let us consider a change of the conformational free energy
of the supermolecule of the -type, F, where = 1 in the case of the lamellar
structure, = 2 in the case of the fibrillar structure, and = 3 in the case of the
globular structure. According to ref. [255], let us confine ourselves to two summands
and write down that
F = F,s + F,e,

(A-6-7)

R2
2
kT ; H sm
is
2
H sm
the mean square distance between ends of the super-macromolecule placed into the
2
medium of the same super-macromolecules with concentration csm; H sm
= l2M/m at
csm ~ 1 according to ref. [256], where M is the molecular mass of the
where F,e is the elastic component of the free energy; F,e =

In ref. [244], such super-macromolecules were named nails consisting of many twisted globules
macromolecules.

655

supermacromolecule; m and l are the mass and the size of the globule
macromolecules; R is the size of the element of the super-molecular structure of the
-type; F,s is the surface component of the free energy appearing due to the effect
of the surface tension forces of the solvent on the super-macromolecule:
F,s = Mv R ,

(A-6-8)

where is the coefficient of the surface tension of the solvent; v is the specific
volume of the super-macromolecule. Influence of the solvent on the supermacromolecule is taken into account in the value of F,s.
Reducing F to minimum by R, let deduce the value of
2
l
R =
3kT

M2

3
v .

Taking into consideration that M/m = Nc, where Nc is the number of globules
macromolecules forming the super-macromolecule, M v = Ncl (with an accuracy of a
constant coefficient), we obtain the final expression for R:
1

2 3 2 3
R =
l Nc l .
3kT

(A-6-9)

Let us discuss now in more detail the mechanism of twisting of the supermacromolecule into a globule or a fibril in order to refine the value of the surface
tension participating in expression (A-6-9). For an example, let us consider a globular
super-molecular structure, although the result obtained may also be applied to other
types of the super-molecular structure.

Figure A-6-5. Initial stage of interaction of globulesmacromolecules and formation of supermolecular structure (the scheme is presented according to data of ref. [92]).

According to the notions developed in ref. [92], globulesmacromolecules


possess a denser center and friable periphery (Figure A-6-5). When globules
macromolecules interact, bond globules are formed in the periphery, which is clearly
observed from Figure A-6-5. If the bond globules appear on two opposite ends of
diameters of the globulesmacromolecules, no motion of elements of the
supermacromolecule would appear, because forces stipulated by the surface tension
from the side of opposite hemispheres of the globulemacromolecule would be
equilibrated. But as the radii connecting centers of globules locate at an angle , the

656

value of which is random, a non-equilibrated force appears, by which the solvent acts
on elements of the supermacromolecule. The value of this force is proportional to a
non-equilibrated square equal to the doubled square cut-off from the surface of the
globulemacromolecule by the bond globule (Figure A-6-5). Hence, the surface
component of the free energy for the globulemacromolecule may be presented in the
2
form: (Fs)glob = Sbond, where Sbond = Rb.glob.
; Rb.glob. is the radius of the bond
globule.
The density of the surface component of the free energy of the globule
macromolecule (sub-index gm), calculated per specific surface, will be equal to
F
s

= ,
2
4Rgm

Rb.glob.
; Rgm is the radius of the globulemacromolecule. From the
where = 2
Rgm

condition of equality of the density of the surface component of the free energy on the
globule of super-molecular structure and on the globulemacromolecule (because
effect of the surface tension on the globule of super-molecular structure is composed
of effects on separate globulesmacromolecules), it is obtained that
1

Fs
Fs
=
= .
2
2
4R 4Rgm
Hence, influence of the solvent on globules of the super-macromolecule may
be taken into account via the efficient coefficient of surface tension , which is
= .
The value of Nc will be found from the condition of equilibrium of volumes of
a freely linked super-macromolecular chain and a globular super-molecular structure
2 3 2
N c3 2l 3 =
l l N c
3kT
or
1

2
N c1 2 =
l .
3kT
In the case of a globular super-molecular structure,
R3
kT
=
.
l
b2
Let us estimate the value of for the case of the globular super-molecular
structure. If, for example, R3/l = 20 [92], l ~ 20 , T = 300 K, = 0.04 Nm,

657

1
=
, 2 =
800

Rb.glob.

= n ,
R

N
glob

where n is the number of units in the bond globule; N is the number of units in the
2
2
= Nlc2 , where lc is the value of the
macromolecule, because Rb.glob.
= nl2, Rgm
macromolecule segment.
Let N 10,000. Then n 24 units, and the periphery of every macromolecule
gives 12 units. In the case of fibrils, R2/l 5, and n 72 units.
These results allow refining of the model of super-molecular structure
formation. If it was suggested in ref. [92] that the fibrillar structure was formed by
elements of the ellipsoid type, it followed from the above estimations that the
dumbbell was the universal element of the super-molecular structure. As the number
of units forming the bond globule at formation of fibrillar and globular structures is
much less than the amount of units in the globulemacromolecule, a dumbbell-shape
figure is formed, when two globulesmacromolecules unite.
In the case of the fibrillar structure, the bond globule is 3-fold greater. As the
bond globule contains a low number of units, it displays rather low (entropic)
elasticity almost in the whole temperature range that enables the super-macromolecule
to adhere in the required shape.
Let us estimate R3. Because = kT/a2 [257], where a is the typical size in the
a2
2
2
liquid, then R3/lb.glob. = 2
, a2 = N s ls2 , lb.glob.
= nlun.
, and R3/l = (Ns/n)(ls/lun.),
lb.glob.
where Ns/n is the number of solvent molecules per one unit of polymer. Let us
determine the value Ns/n from the condition that the polymer unit forms a sphere, and
solvent molecules locate in the spherical layer near the sphereunit:
2

2
R
R
N s 3Run.
2 Rs + 3Run. 4 Rs2 + 8Rs3
=
= 6 un. + 12 un. + 8 .
n
R
Rs3
s
Rs

Initial data and calculation results of sizes of elements of the super-molecular structure for a
series of polymers at polymerization degree N = 100
Vi , 3
R,

R /l
Solvent
i

Chloroform
DCE
Benzene
Toluene
Dioxane

Chloroform
Benzene
Dioxane

= 96.8 3, l = 21
Poly(methyl methacrylate), Vi

i
un.
69.6
0.896
23.17
76.2
0.923
23.89
88.2
0.969
25.14
105.1
1.028
27.13
87.7
0,968
25.11

= 109.7 3, l = 22
Polystyrene, Vi

un.
69.6
0.859
22.21
88.2
0.930
24.08
87.7
0.928
24.03

487
502
527
569
527

488
530
528

658

Chloroform
Dichloroethane
Tetrachloroethane
Nitrobenzene
Tetrahydrofuran

Aromatic polyester, Vi
= 390.2 3, l = 34

i
un.
69.6
0.178
8.39
76.2
0.195
8.64
105.8
0.271
9.84
105.1
0.269
9.81
84.2
0.215
8.96
2

286
292
333
333
306

l R
R
R
R
Since s = s , then 3 = 6 + 12 s + 8 s . The value Rs/Run.
l
Run. Run.
lun. Run.
1

may be presented in the form: = Rs/Run

3
= Vi
i
s

3
Vi , where
i
un.

Vi and Vi are the Van-der-Waals volumes of the solvent molecule


i
s
i
un.
and the repeat unit of the polymer, respectively. If we assume from the experimental

Vi 2.2.
data that Ns/n = 20(Run/Rs)2, then it is obtained that Vi
i
un. i
s
On a series of particular examples, let us calculate the dependence of R3/l on
. Initial data and calculation results for three polymers poly(methyl methacrylate),
polystyrene, and phenolphthalein and isophthalic acid polyester are shown in the
Table above. It is indicated that if the Van-der-Waals volumes of the repeat unit of
polymer and the solvent molecule are similar, then ~ 1, and the ratio of the size of
the element of the super-molecular structure R3 to the globulemacromolecule size l
(R3/l) changes in a narrow range from 22 to 27. If the Van-der-Waals volume of the
solvent molecule is significantly smaller than the volume of the repeat unit of
polymer, then and R3/l are significantly lower (0.180.27 and 8.49.8, respectively).
However, if the absolute size of the element of the super-molecular structure R3 is
calculated, these values are close for all polymers (see the Table). It should be noted
on this subject that electron micrographs for various amorphous polymers, observed
in numerous works, display approximate equality of these sizes, which was one of the
reasons to ascribe this structure to artifacts and to doubt in reality its existence. As
indicated by estimations for polymers of different classes made in ref. [94], this very
picture must be observed.
In the case of fibrils, from the condition of equality of the volumes of the
cylinder formed by a freely linked chain of the supermacromolecule and a fibril, we
obtain that
2

2 3 4 3 2
l N c l L ,
l LN c =
3kT
2

where L is the length of the element of the cylinder (fibril).


1
2 2
Nc 3 =
l
3kT

2 3

, R2 =

3kTl
,
2 l 2

659

and the ratio R2/R3 =

2
3Rb.glob.
g
2
2 Rb.glob.
f

3ng
2nf

, where ng is the number of units in the bond

globule of the globular super-molecular structure; nf is the number of units in the bond
globule of the fibrillar super-molecular structure. Basing on the experimental data
[767], it is obtained that R2/R3 1/4 and ng/nf 1/6 (in the same solvent).
Let now determine the value of the free energy of the domain (supermacromolecule) F. With the value of R calculated by equation (A-6-9), expression
(A-6-7) gives
F = 3 2

Mv .
R

Let us now estimate free energy of globular and fibrillar super-molecular


structures:
Fg =
Fg
Ff

9 g
2 R3

V , Ff =

9 f
V;
2 R2

3 g R2 3 g R2 3ng R2 3ng
=
=
=
2 f R3 2 f R3 2nf R3 2nf

Fg 1
3
1
= , and
=
<< 1, i.e. the fibrillar structure is
Ff 16
2nf 2 6 4
energetically less profitable than the globular one in the current solvent. To obtain a
fibrillar structure, the following condition must be fulfilled:
Estimations give

3ng

Fg + F Ff,
where F may be associated with heat delivery in the case of solution or with
application of mechanical energy (in the solid state). To obtain a bond globule
appropriate to the fibrillar structure, a polymer must be exposed in the area of the
coilglobule transition temperature Tcr. The value of the bond globule must depend on
the difference of temperatures Tcr T and the rate of evaporation of the solvent.
To obtain the desired super-molecular structure, Tcr for the given polymer
solvent system must be known. Usually, Tcr is significantly higher than the
experiment temperature, or it rapidly shifts to the side of high temperatures at solvent
evaporation. That is why peripheries of globulesmacromolecules do not manage to
penetrate into each other. If this process is decelerated artificially, formation of a
fibrillar structure may be expected.
Obtaining ratios for R3/l and R2/R3, the ratio for R2/l may be obtained:
2
Rs
Rs
R2 ng
.
= 9 + 18
+ 12
l
nf
Run.
Run.

Let assume that ng/nf = g/f, where g is the thickness of the periphery of the
globulemacromolecule for the globular super-molecular structure; f is the thickness

660

of the periphery of the globule of macromolecule of the fibrillar super-molecular


structure.
The main factor affecting the periphery thickness is the difference of
temperatures, Tcr T. If this value is extremely high, then g. In the general case,
if the results from refs. [92, 257, 258] are used, it may be indicated that

= g.+ (R g) e

T Tcr
T

When T = Tcr, = R, and when (Tcr T)/T , g:

ng

,
R

T Tcr

1 exp
1+

g
T

T
R
7T
T
= 5 ,
7 , Tlim cr .
exp
T
T

g
cr
cr

nf

If the initial temperature of solvent evaporation at casting of the film


corresponds to the condition T > Tlim, then the polymer will transit into the globule
state. When T < Tlim, formation of a fibrillar super-molecular structure may be
expected. Therewith, it should be taken into account that when the solvent
concentration changes, Tcr will also change, and to obtain a fibrillar structure, Tcr must
always be controlled.
Determined above were the sizes of various types of super-molecular
structure, their relation to the chemical structure of the polymer and the solvent was
stated. However, the question remains, how a globule or fibril is formed from a supermacromolecule existing in a relatively diluted solution that activates formation of
closed forms, the process of super-molecular structure formation.
Data accumulated up to now enable to answer this question. Usually, the film
is formed in a definite volume at gradual evaporation of the solvent. Therewith,
temperature gradients occur between the solvent surface and the vessel bottom.
Moreover, gradients of surface tension forces occur on surfaces of elements of supermacromolecules. Consequently, a convection motion occurs in the solution, and at
some concentration of the solvent, the Benar cells are formed (either spheres or
cylinders). The moment of appearance of the Benar cells is determined by the
Marangoni number, because in thickness of the solution considered a significant role
is played by gravitationcapillary effects [259]. The type of excited motion is
determined by rigidity of the chain of super-macromolecule, i.e. by the size of the
bond globule. Note that in contrast to simple viscous liquids and their mixtures, in the
case of macromolecules, only the first modes of convection motion will be excited
(rotation of the whole sphere or cylinder), because other modes of the motion cannot
be excited due to bulk of the super-macromolecular chain. This convection motion
may ascribe the shape to super-macromolecules, which they possess in the supermolecular structure. The fact that formation of the super-molecular structure is
associated with convection motions may be observed experimentally. Introducing spin
labels into a polymer and observing the solution behavior in an alternating magnetic
field, one may observe a resonance absorption in the range of low frequencies
associated with particular motions of the super-macromolecule. This resonance

661

absorption is absent in diluted solutions, appears at the moment of super-molecular


structure formation and will last up to complete evaporation of the solvent.

3 Phase state of polymers as a result of formation of the supermolecular structure by one-cavity bond hyperboloids
To describe the amorphous state of a substance, the space of constant negative
curvature (the Lobachevski space) has been used in some works [260262]. When
structures given in the Lobachevski space are mapped on the Euclidean plane, an
irregularity appears, usually registered by X-ray analysis. Moreover, as indicated in
ref. [263], in a space with a negative curvature crystalline lattice considered in the
harmonic approximation becomes unstable at quite high temperature melting is
observed.
We have described the model of the super-molecular structure of polymers,
formed on the account of interaction (bonding) of globulesmacromolecules via
periphery (Figure A-6-5). Application of this model to the description of solubility
[264] demanded revision of the structure formed by peripheries when globules
macromolecules are bonded. It was found that this structure must possess the surface
of a negative curvature. Owing to this very negative curvature of the structure surface
(bond belt) formed between globules of the super-molecular structure, the solvent
affects the globules by forces, which (under definite conditions) cause their
detachment.
In this Section, let us determine the curvature of the surface of the bond
structure formed between globulesmacromolecules and super-molecular structures
via refinement of the structure and associate the former with parameters of the supermolecular structure (thickness of peripheries, size of elements of super-molecular
structure, globulesmacromolecules, etc.). Therewith, let us base on the results
obtained in ref. [93]. Furthermore, proceeding from particular expression for
curvature, let us analyze the influence of the parameters of the super-molecular
structure on the phase state of the polymer. Let us also determine the size of the
elementary structure (amorphone) producing the amorphous state [265]. In turn,
influence of the curvature of bond hyperboloid on the phase state of polymer appears
after potential mapping of atoms interaction (for example, the LennardJones
potential) on a surface with negative curvature (the Lobachevski space).
Consequently, additional terms appear in the interaction potential, which depend on
the space curvature.
In ref. [264], the bond region of globulesmacromolecules and globules of
super-molecular structure is considered as an element of a structure with negative
curvature. To perform further calculations, let us refine the shape of the surface of this
element. Let us assume that at quite good approximation (the latter fact is confirmed
by calculations of various characteristics of polymers stipulated by the supermolecular structure) this surface may be considered as a one-cavity hyperboloid. Then
within the framework of the globule (or fibril) of the super-molecular structure, onecavity hyperboloids form the full surface of negative curvature when crossed [266].
Let us determine curvature K of the one-cavity hyperboloid via the sizes of the
periphery and the globulemacromolecule. Therewith, let us assume that bonding
regions between globulesmacromolecules and globules of the super-molecular
structure possess the same size, because in both cases bonding is performed via the
globulemacromolecule. Figure A-6-6 depicts the one-cavity bond hyperboloid
formed by mutual penetration of peripheries of two globulesmacromolecules. The
main curvature radii R1 and R2 will be respectively equal to: R1 = CB, R2 = AB. If R1

662

and R2 are expressed via radii of globules macromolecules Rgl-m and periphery
thickness l, we obtain that

Figure A-6-6. Scheme for calculating curvature of the one-cavity bond hyperboloid: 1, 2 globules
macromolecules 1 and 2; 3 one-cavity bond hyperboloid; dotted lines mark borders of
peripheries of globules 1 and 2.

R1 = l,

R2 = Rgl-m

1 l
Rglm 4 Rglm
l

(A-6-10)

Since
l
4 Rglm

<< 1, then R2

lRglm .

The value of coefficient will be refined in future calculations. According to the


definition, the curvature value K = 1/R1R2, where R1 and R2 are determined by
formulae (A-6-10).
In real polymeric systems, the curvature of one-cavity bond hyperboloids will
fluctuate owing to molecular-mass distribution (because the thickness of the periphery
depends on the molecular mass). That is why it is necessary to average the curvature
by quite a great number of macromolecules within the framework of the element of
the super-molecular structure and to operate with the mean curvature in future.
Let us consider now the potential of interaction of atoms forming the bond
hyperboloid in the negative curvature space. As indicated in ref. [263], critical
temperatures do not depend on the metrics selection. The most important is that the
metrics must correspond to the space with a constant negative curvature. That is why,
furthermore, conformal Euclidean metrics will be used, because in this case,
expressions for potential and appropriate calculations are simpler.

663

According to ref. [263], in the present case, the potential of interaction of


atoms in the space with negative curvature is of the following form:
R

= 0 + K 0 ( R) R 2 dR ,

(A-6-11)

R0

where 0 is the desired potential in the absence of curvature; R0 is the coordinate at


which 0(R0) = 0.
Let 0 be described by the LennardJones potential:
1
2
0() = D 12 6 .

Here is the non-dimensional coordinate; = R/a, a is the distance between


neighboring lattice points. Then the potential (A-6-11) in the non-dimensional system
of coordinates may be presented in the following form:
() = 0() + Ka2

0 ( )

d ,

(A-6-12)

where 0 is the coordinate, at which 0(0) = 0, 0 = R0/a; K is the curvature of the


surface possessing negative curvature in the Lobachevski space.
Integrating equation (A-6-12), we obtain that
1
1
( ) 1
2
1
1
= 12 6 + 12 Ka 2
4 12 Ka 2
4 .
10
10
D
4 0
4 0

10 0
10 0

(A-6-13)

From the condition 0(0) = 0, we obtain that 06 = 1/2. Substituting 0 into


the expression for potential (A-6-13), we finally obtain that
1
( )
1
2
1
= 12 6 + 12 Ka 2
4 + Ka 2 .
10
D
4 0

10 0

(A-6-14)

From the condition () = 0, we find extreme values of the potential (A-614):


1
1

6 + 1 2 + Ka 2 = 0.

In the case, if K > 0, () possesses a single maximum at = 1, which


coincides with the minimum of the potential 0(). This result is not of interest. When
K < 0, () possesses the minimum at 1 = 1, and the maximum at 2 = 1/ a K , if
2 > 1. When 2 < 1, potential (A-6-14) possesses the maximum at 1 = 1 and the
minimum at 2:

664
6
3
(1 )
( 2 )
( )
4
1
= 1 + Ka 2 ,
= Ka 2 Ka 2 + Ka 2 ,
Ka 2 .
D
D
D
5
5
It is clear that at 2 > 1, i.e. at |Ka2| < 1, the atom existing in the fields of two
neighboring atoms will possess only one state of equilibrium. When 2 < 1, i.e.
|Ka2| > 1, the atom existing in the field of two neighboring atoms possesses two states
of equilibrium. Therewith, if all three atoms are of the same type, or the atom located
in the point = 1 interacts with similar (but of a different type, than in the point = 1)
atoms located in the points = 0 and = 2, the difference of energies of potential pits
in points 2 and 2 (the difference in energies of rotary isomers) E = 0. But if the
atom located in the point = 1 interacts with different atoms located in points = 0
and = 2, then 1 2 2 1 and E 0 (Figure A-6-7). Summing up potentials,
we will proceed from the condition that positions of the maximum of the potential
() of the left and the right atoms coincide (although this condition is not
obligatory).

( ) ( )

Figure A-6-7. Potential of interaction of the given atom (1 = 1) with two neighboring atoms ( = 0
and = 2) under two different conditions: 2 > 1 (a) or 2 < 1 (b).

Hence, when |Ka2| > 1, summation of potentials (A-6-14) will give two states
of equilibrium for every atoms, which is typical of systems of the orderdisorder type.
Systems of similar type display a phase transition of the second kind [267]. To
provide transitions of atom from one state of equilibrium into another, the presence of
defects (of the disclination type) in the polymer is necessary, the size of which
(angular values) may be determined via coefficients of molecular packing of
amorphous and crystalline states. When |Ka2| < 1, summation of potentials (A-6-14)
will give only one state of equilibrium for each atom, which is typical of regular
systems of the crystal type. In such systems, the phase transition of the first order
takes place [267].
Let estimate the value basing on the following ideas. According to formula
(A-6-10), K = 1 l lRglm . If assume that a ~ 2 (this corresponds to two Van-der-

Waals radii of hydrogen atom), and Rgl-m 20 , then at Ka2 = 4 l 20l ~ 1, 2/3
l ,
1

2/3

1 lun. nper . For the globule, nper ~ 24 [94],

2/3

1/(5lun.), lun. ~ 3 ,

0.45, i.e. is approximately equal to the half of the periphery thickness. That is why, R1
in expression (A-6-10) may be presented in the following form: R1 = 0.5lun. nper . If
nper > nf or nper > ngl (where nf and ngl are the numbers of units in the bond

665

hyperboloid in the case of fibrillar or globular super-molecular structure [94]), the


polymer exists in the crystalline state. On the base of data from ref. [94], let us
determine

(Ka )
(Ka )
2

l
= c
la

n
= c
na

= 3 4 2.3 .
3

Since (Ka2)c = 1, (Ka2)a = 2.3 and 2

0.7. The value 2 is appropriate to


Ka 2
the size of the molecular defect and may be estimated via the coefficient of molecular
packing in amorphous, ka, and crystalline, kc, state by the relation [94]:2 = (kc ka)1/3.
Usually in the crystalline state, the mean packing coefficient kc = 0.74, and in the
amorphous one ka = 0.68; then 2 0.45.
Let us now determine the size of amophone3. For scattering of X-rays or
electrons on the globulemacromolecule to be fully isotropic, it is necessary that the
whole surface of the globule is overlapped by cross-sections of the rotation
hyperboloid with radius R2, i.e. 4Rgl2 m = NR22 .
N=

Rgl2 m
R22

4 Rglm
l

=4

M mol
~ 40 ,
ndet

L2a = 4 NRgl2 m ~ 160Rgl2 m , Na ~ 12Rgl-m ~ 240 .


The cross size of the element of the super-molecular structure in a polymer
fibril approaches by the order of value the size of amorphone estimated above. That is
why the fibril may be approximated as a tube possessing the radius of amorphone and
filled with amorphones along the generatrix. In the globule of the super-molecular
4
4
structure, the amount of globulesmacromolecules equals Rgl3 = N m Rgl3 -m ,
3
3
Nm =

Rgl3
Rgl3 -m

5 10 2
~ 16 103 ,
~

2
10

Nm
~ 400 ,
Na m

where Nm is the number of globulesmacromolecules forming the globule of the


super-molecular structure; Nam is the number of globulesmacromolecules forming
the amorphone. Hence, the globule of the super-molecular structure contains ~400
amorphones.
Given here were statistical estimates of amorphone. Determining the number
of amorphones in the globule of the super-molecular structure, it was suggested that it
has the spherical shape. In reality, to determine its shape, it is necessary to find, from
the condition of dense packing of globules in the element of the super-molecular
structure, the shape of the bordering surface providing isotropic scattering of X-rays
3

The assumption used is excessive and guarantees fully isotropic scattering; more precise solution of
the problem requires determination of the amount of cross-sections and their projections appearing at
overlapping of bond hyperboloids under condition of formation of element of the super-molecular
structure.

666

or electrons on globulesmacromolecules. This sort of problems may be solved by a


computer.
The model of the super-molecular structure considered enables to refine the
physical meaning of elements of the nonlinear mechanical model of polymeric
substances, suggested in ref. [87]. This model is depicted in Figure A-6-8. According
to this model, relaxation processes are described by the spectrum of relaxation times,
which contains two main times (short and long) and a number of other relaxation
times dependent on these two main times. In the right part of Figure A-6-8 shown the
globulemacromolecule and radial lines connecting the center O of the globule
macromolecule with canters of neighboring globulesmacromolecules. Let us
separate the typical element, deformation of which determines the deformation
behavior of the whole polymeric substance. In Figure A-6-8, this element is marked
by the dotted line. Two thresholds may be indicated in the mentioned element: one
threshold (OA) includes a bond hyperboloid, and another one (OB) does not include it.
Therewith, OA OB, but taking into account the small thickness of the periphery, it
may be assumed that OA OB, i.e. the supposition about equal lengths of elements in
the nonlinear mechanical model made in ref. [87], may be assumed to be true.

Figure A-6-8. Element of super-molecular structure (right) and nonlinear mechanical model of
polymeric body appropriate to it (See text).

Let us estimate the angle between the axis lines of elements of the supermolecular structure (pairs of bonded globules). This may be made proceeding from
the following ideas. In the case of spatial contacts, = 4/n, where n is the number of
globules present in the first spherical layer near the present globulemacromolecule of
the radius Rgl-m. Let us determine the value of n from the following expression:
2
Rglm + 2 Ri )3 Rgl3 m
R
(
Rglm
+ 12 glm ,
n=
= 8 + 6

Ri3

R
i

Ri

where Ri is the radius of the i-th globule present in the spherical layer near the globule
of radius Rgl-m. In the general case, Rgl-m/Ri is a random value, because Ri may obtain
values depended on molecular mass and molecular-mass distribution.
Let us estimate values of n and from different border cases. When
Rgl-m/Ri = 1, n = 26, = /6;
Rgl-m/Ri 0, n = 8, = /2;

667

Rgl-m/Ri , n , = 0.
Suggesting these cases to be equally probable, let us calculate the value of as
=

1
2
.
+ + 0 =
3 6 2
9

Hence, corresponds to Rgl-m/Ri ~ 1 (more accurate, to Rgl-m/Ri = 0.63).


The model of the super-molecular structure suggested allows complete
definition of the physical meaning of relaxation mechanisms of the Alexandrov
Lazurkin elements, participating in the nonlinear model: one mechanism associated
with long relaxation times takes place in bond hyperboloids and may be described
with the help of the orderdisorder system in the isomeric approximation by
Volkenstein. The second mechanism with shorter relaxation times is analogous to the
mechanism associated with stagnation of internal rotation and is developed in the
globulemacromolecule itself (the BreslerFrenkel mechanism).
So, one of the main relaxation times (the short one) is typical of the globule of
super-molecular structure of a positive curvature (corresponds to the time in
ferroelectric of the mixing type), and the second one is typical of the bond
hyperboloid with a negative curvature (long time in the system of the orderdisorder
type).
The time and frequency dependences of polymer behavior in the transitional
region can be described satisfactorily with the help of the IsakovichChaban [268,
269] and IrzhakRozenbergRostiashvili [270] models. In the one-dimensional case,
it follows from these models that the main relaxation process is stipulated by the
mutual diffusion of five to seven atoms. In the case of the model of the suggested
super-molecular structure, these regions of rearrangement of atoms may be concretely
defined they must proceed in the bond hyperboloids.

669

References

1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.

13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.

31.

A.P. Alexandrov, Proc. I and II Conf. on High-Molecular Compounds, Moscow-Leningrad, Izd.


AN SSSR, 1945, pp. 49-59. (Rus)
O.V. Anulov, G.V. Baryishnikov, N.I. Shimanova, N.M. Mestechkina, and V.D. Shcherbukhin,
Appl. Biochem. And Microbiol., 1995, vol. 31(1), pp. 87-91. (Rus)
P.U. Arifov, S.A. Tishin, A.V. Shevchenko, and V.A. Tishin, Vysokomol. Soedin., 1987, vol.
B29(1), pp. 59-63. (Rus)
A.A. Askadskii, Physicochemistry of polyarylates. Moscow, Khimia, 1968, 214 p. (Rus)
A.A. Askadskii, Deformation of polymers. Moscow, Khimia, 1973, 448 p. (Rus)
A.A. Askadskii, Structure and properties of heat-resistant polymers. Moscow, Khimia, 1981, 320
p. (Rus)
A.A. Askadskii, Mechan. Comp. Mater., 1987, No. 3, pp. 403-409. (Rus)
A.A. Askadskii, Vysokomol. Soedin., 1989, vol. A31(10), pp. 2141-2148. (Rus)
A.A. Askadskii, Mechan. Composite Mater., 1990, No. 6, pp. 963-977. (Rus)
A.A. Askadskii, V.A. Arslanbekov, T.A. Andryushchenko, Yu.I. Matveev, and A.L. Blumenfeld,
Vysokomol. Soedin., 1989, vol. A31(8), pp. 1616-1623. (Rus)
A.A. Askadskii, R.B. Banyavichus, Z.S. Vikhauskas, A.I. Marma, and A.L. Blumenfeld,
Vysokomol. Soedin., 1988, vol. A30(8), pp. 1684-1689. (Rus)
A.A. Askadskii, V.A. Beloshenko, K.A. Byichko, V.V. Shaposhnikova, A.V. Samoryadov, O.V.
Kovriga, S.N. Salazkin, V.A. Sergeev, V.G. Slobodina, and Ya.V. Genin, Vysokomol. Soedin.,
1994, vol. A36(7), pp. 1143-1147. (Rus)
A.A. Askadskii, A.L. Blumenfeld, E.G. Galpern, and A.L. Chistyakov, Vysokomol. Soedin., vol.
A36(7), pp. 1143-1147. (Rus)
A.A. Askadskii, K.A. Byichko, O.V. Kovriga, S.V. Tsutsuran, R.B. Shleifman, and V.V.
Kurashev, Vysokomol. Soedin., 1991, vol. A33(11), pp. 2477-2486. (Rus)
A.A. Askadskii, and M.P. Valetskii, Mechan. Composite Mater., 1990, No. 3, pp. 441-446. (Rus)
A.A. Askadskii, and I.I. Vointseva, Vysokomol. Soedin., 1987, vol. A29(12), pp. 2654-2669.
(Rus)
A.A. Askadskii, E.G. Galpern, T.P. Matveeva, A.L. Chistyakov, and G.L. Slonimskii,
Vysokomol. Soedin., 1987, vol. A29(11), pp. 2433-2440. (Rus)
A.A. Askadskii, E.G. Galpern, G.V. Surov, V.V. Bulatov, G.L. Slonimskii, and V.V. Korshak,
Vysokomol. Soedin., 1989, vol. B31(4), pp. 308-317. (Rus)
A.A. Askadskii, L.M. Goleneva, and K.A. Byichko, Vysokomol. Soedin., 1995, vol. A37(5), pp.
829-841. (Rus)
A.A. Askadskii, V.V. Kazantsva, O.A. Melnik, K.A. Byichko, A.A. Sakharova, G.L. Slonymskii,
and T.M. Frunze, Vysokomol. Soedin., 1988, vol. A30(6), pp. 1285-1293. (Rus)
A.A. Askadskii, A.I. Kitaigorodskii, and G.L. Slonimskii, Vysokomol. Soedin., Vysokomol.
Soedin., 1975, vol. A17(10), pp. 2293-2300. (Rus)
A.A. Askadskii, and A.F. Klinskikh, Plasticheskie Massy, 1998, No. 4, pp. 29-33. (Rus)
A.A. Askadskii, and A.F. Klinskikh, Vysokomol. Soedin., 1999, vol. A41(1), pp. 83-85. (Rus)
A.A. Askadskii, and O.V. Kovriga, Vysokomol. Soedin., 1991, vol. A33(9), pp. 1945-1955. (Rus)
A.A. Askadskii, L.K. Kolmakova, A.A. Tager, G.L. Slonymskii, and V.V. Korshak, Vysokomol.
Soedin., 1977, vol. A19(5), pp. 1004-1013. (Rus)
A.A. Askadskii, and V.I. Kondrashchenko, Vysokomol. Soedin., 1997, vol. A39(10), pp. 16251634. (Rus)
A.A. Askadskii, V.M. Litvinov. A.A. Zhdanov, and G.L. Slonymskii, Vysokomol. Soedin., 1985,
vol. A27(11), pp. 2408-2414. (Rus)
A.A. Askadskii, and Yu.I. Matveev, Chemical structure and physical properties of polymers.
Moscow, Khimia, 1983, 248 p. (Rus)
A.A. Askadskii, Yu.I. Matveev, G.L. Slonymskii, and V.V. Korshak, Doklady AN SSSR, 1978,
vol. 238(3), pp. 592-595. (Rus)
A.A. Askadskii, Yu.I. Matveev, A.V. Pastukhov, B.A. Rozenberg, T.I. Ponomareva, N.A.
Shchegolevskaya, and A.S. Marshalkovich, Vysokomol. Soedin., 1983, vol. A25(1), pp. 56-71).
(Rus)
A.A. Askadskii, Yu.I. Matveev, and T.P. Matveeva, Vysokomol. Soedin., 1988, vol. A30(12), pp.
2542-2550. (Rus)

670

32. A.A. Askadskii, Yu.I. Matveev, and M.S. Motevosyan, Vysokomol. Soedin., 1990, vol. A32(10),
pp. 2157-2166. (Rus)
33. A.A. Askadskii, M.S. Motevosyan, and G.L. Slonymskii, Dep. VINITI, No. 6949-C, Dep. IEOC
October 2, 1985, Moscow, 26 p. (Rus)
34. A.A. Askadskii, M.S. Motevosyan, and G.L. Slonymskii, Vysokomol. Soedin., 1987, vol. A29(4),
pp. 753-760. (Rus)
35. A.A. Askadskii, A.V. Pastukhov, and A.S. Marshalkovich, Vysokomol. Soedin., 1984, vol.
A26(1), pp. 160-171. (Rus)
36. A.A. Askadskii, S.N. Prozorova, and G.L. Slonymskii, Vysokomol. Soedin., 1976, vol. A18(3),
pp. 636-647. (Rus)
37. A.A. Askadskii, and T.R. Razmadze, Vysokomol. Soedin., 1991, vol. A33(5), pp. 1141-1148.
(Rus)
38. A.A. Askadskii, T.R. Razmadze, S.N. Salazkin, V.A. Sergeev, A.V. Samoryadov, K.A. Byichko,
O.V. Kovriga, T.M. Babchinitser, and A. Varada Ragulu, Vysokomol. Soedin., 1991, vol. A33(6),
pp. 1239-1250. (Rus)
39. A.A. Askadskii, A.A. Sakharova, O.A. Melnik, and V.V. Kazantseva, Vysokomol. Soedin., 1997,
vol. A39(7), pp. 1160-1165. (Rus)
40. A.A. Askadskii, and G.L. Slonymskii, Vysokomol. Soedin., 1971, vol. A13(8), pp. 1917-1919.
(Rus)
41. A.A. Askadskii, G.L. Slonymskii, and A.I. Kitaigorodskii, Vysokomol. Soedin., 1974, vol.
A16(2), pp. 424-430. (Rus)
42. A.A. Askadskii, G.L. Slonymskii, Yu.I, Matveev, and V.V. Korshak, Doklady AN SSSR, 1975,
vol. 224(3), pp. 612-615. (Rus)
43. A.A. Askadskii, G.L. Slonymskii, Yu.I, Matveev, and V.V. Korshak, Vysokomol. Soedin., 1976,
vol. A18(9), pp. 2067-2074. (Rus)
44. A.A. Askadskii, G.V. Surov, V.V. Nemchinov, A.L. Blumenfeld, and Z.S. Vikhauskas,
Vysokomol. Soedin., 1989, vol. A31(6), pp. 1320-1327. (Rus)
45. A.A. Askadskii, G.V. Surov, V.A. Pankratov, Ts.M. Frenkel, A.A. Zhdanov, L.I. Makarova, A.S.
Marshalkovich, and L.G. Radchenko, Vysokomol. Soedin., 1990, vol. A32(7), pp. 1517-1527.
(Rus)
46. A.A. Askadskii, G.V. Surov, V.A. Pankratov, Ts.M. Frenkel, L.I. Makarov, A.A. Zhdanov, I.V.
Blagodetskikh, and A.V. Pastukhov, Vysokomol. Soedin., 1990, vol. A32(7), pp. 1528-1534.
(Rus)
47. A.A. Askadskii, G.V. Surov, A.S. Marshalkovich, V.V. Dvalishvili, V.A. Pankratov, and Ts.M.
Frenkel, Mechan. Composite Mater., 1991, No. 1, pp. 3-8. (Rus)
48. A.A. Askadskii, S.A. Tishin, V.V. Kazantseva, and O.V. Kovriga, Vysokomol. Soedin., 1990,
vol. A32(12), pp. 2437-2445. (Rus)
49. A.A. Askadskii, S.A. Tishin, M.I. Tsapovetskii, V.V. Kazantseva, O.V. Kovriga, and V.A. Tishin,
Vysokomol. Soedin., 1992, vol. A34(12), pp. 62-72. (Rus)
50. Yu.S. Barash, J. Exper. And Theor. Phys., 1980, vol. 79(6), pp. 2271-2281. (Rus)
51. G.M. Bartenev, and Yu.V. Zelenev, Physics and Mechanics of Polymers. Moscow, Vysshaya
Shkola, 1983, 391 p. (Rus)
52. G.M. Bartenev, and S.Ya. Frenkel, Physics of polymers. Ed. A.M. Elyashevich, Leningrad,
Khimia, 1990, 432 p. (Rus)
53. G.M. Bartenev, and A.G. Barteneva, Relaxation properties of polymers. Moscow, Khimia, 1992,
383 p. (Rus)
54. A.Z. Varisov, Yu.N. Kuznetsov, E.P. Prokopiev, and A.I. Filipiev, Uspekhi Khimii, 1981, vol.
50(10), pp. 1892-1923. (Rus)
55. S.V. Vinogradova, S.N. Salazkin, G.Sh. Chelidze, G.L. Slonymskii, A.A. Askadskii, K.A.
Bychko, L.I. Komarova, I.V. Zhuravleva, and V.V. Korshak, Plasticheskie Massy, 1971, No. 8,
pp. 10-13. (Rus)
56. S.V. Vinogradova, V.A. Pankratov, Ts.M. Frenkel, L.F. Larina, L.I. Komarova, and V.V.
Korshak, Vysokomol. Soedin., 1981, vol. A23(6), pp. 1238-1243. (Rus)
57. M.V. Volkenstein, Molecular optics. Moscow-Leningrad, Gos. Izd. Tekhn. I Teor. Lit., 1951, 744
p. (Rus)
58. M.V. Volkenstein, Configuration statistics of polymeric chains. Moscow-Leningrad, Izd. AN
SSSR, Leningrad dept., 1959, 466 p. (Rus)
59. B. Vunderlich, and G. Baur, Heat capacity of linear polymers (Transl. From Engl. And Germ.).
Moscow, Mir, 1972, 238 p. (Rus)
60. Hemicelluloses. Ed. V.S. Gromov and M.S. Dudkin, Riga, Zinatne, 1991, 488 p. (Rus)

671

61. Yu.K. Godovskii, Thermophysical methods of polymers study. Moscow, Khimia, 1976, 216 p.
(Rus)
62. Yu.K. Godovskii, Thermophysics of polymers. Moscow, Khimia, 1982, 280 p. (Rus)
63. V.E. Gul, Colloid. Zh., 1953, vol. 15(3), pp. 170-177. (Rus)
64. L.V. Gurvich, G.V. Karachentsev, V.N. Kondratiev, Yu.A. Lebedev, V.A. Medvedev, V.K.
Potapov, and Yu.S. Khodeev, Energies of chemical bonds break. Ionization potentials and affinity
to electron. Moscow, Nauka, 1974, 351 p. (Rus)
65. G.I. Gurevich, Zh. Tekhn. Fiz., 1947, vol. 17(12), pp. 1491-1502. (Rus)
66. V.G. Dashevskii, Conformation of organic molecules. Moscow, Khimia, 1974, 432 p. (Rus)
67. B.A. Dogadkin, and V.E. Gul, Doklady AN SSSR, 1950, vol. 70(6), pp. 1017-1019. (Rus)
68. M.M. Dubinin, Surface and porosity of adsorbents. In: Main Problems of the theory of physical
adsorption. Proc. I-st All-Union Conf. on Theor. Problems of Adsorption. Moscow, Nauka, 1970,
pp. 251-269. (Rus)
69. N.S. Enikolopov, and S.A. Volfson, Chemistry and technology of polyformaldehyde. Moscow,
Khimia, 1968, 279 p. (Rus)
70. I.G. Zedginidze, Planning of experiment for study of multi-component systems. Moscow, Nauka,
1976, 390 p. (Rus)
71. P.I. Zubov, Mechanism of formation of protein jellies. Doctor Thesis, Moscow, L.Ya. Karpov
SICPI, 1949, 174 p. (Rus)
72. P.I. Zubov, Yu.I. Matveev, A.A. Askadskii, and T.A. Andryushchenko, Vysokomol. Soedin.,
1980, vol. A22(6), pp. 1347-1358. (Rus)
73. A.A. Iliushin, and B.E. Pobedrya, Grounds of mathematical theory of thermal viscoelasticity.
Moscow, Nauka, 1970, 280 p. (Rus)
74. V.A. Kabanov, Doklady AN SSSR, 1970, vol. 195(2), pp. 402-405. (Rus)
75. A.I. Kitaigorodskii, Organic crystallochemistry. Moscow, Izd. AN SSSR, 1955, 558 p. (Rus)
76. V.S. Klimenkov, V.A. Kargin, and A.I. Kitaigorodskii, Zh. Fiz. Khim., 1953, vol. 27(8), pp.
1217-1227. (Rus)
77. M.I. Kornfeld, Elasticity and strength of liquids. Moscow-Leningrad, Gos. Izd. Tekhn. Teor. Lit.,
1951, 108 p. (Rus)
78. V.V. Korshak, Heterochain structure of polymers. Moscow, Nauka, 1977, 301 p. (Rus)
79. V.V. Korshak, S.V. Vinogradova, G.L. Slonymskii, V.A. Pankratov, A.A. Askadskii, Ts.M.
Frenkel, L.F. Larina, and K.A. Bychko, Vysokomol. Soedin., 1981, vol. A23(6), pp. 1244-1251.
(Rus)
80. V.V. Korshak, and A.A. Askadskii, Doklady AN SSSR, 1986, vol. 289(6), pp. 1412-1416. (Rus)
81. V.V. Korshak, G.L. Slonymskii, A.A. Askadskii, A.L. Rusanov, T.V. Lekae, and M.S.
Matevosyan, Vysokomol. Soedin., 1989, vol. A31(1), pp. 34-39. (Rus)
82. V.G. Krasovskii, N.A. Sadovskii, O.B. Gorbatsevich, A.M. Muzafarov, V.D. Myakushev, M.N.
Ilina, I.I. Dubovik, T.V. Strelkova, and V.S. Papkov, Vysokomol. Soedin., 1994, vol. A36(4), pp.
714-720. (Rus)
83. E.S. Krongauz, N.M. Belomoina, A.A. Askadskii, V.V. Kazantseva, K.A. Bychko, G.L.
Slonymskii, and V.V. Korshak, Vysokomol. Soedin., 1986, vol. A28(7), pp. 1473-1478. (Rus)
84. A.S. Kuzminskii, and L.V. Borkova, Zh. Fiz. Khim., 1958, vol. 32(9), pp. 2051-2060. (Rus)
85. L.D. Landau, and E.M. Lifshits, Theoretical physics. Vol. 8. Electrodynamics of continuum. 2nd
Ed., improved and added by E.M. Lifshits and L.P. Pitaevskii, Moscow, Nauka, 1982, 623 p.
(Rus)
86. L.D. Landau, and E.M. Lifshits, Theoretical physics. Vol. 7. Elasticity theory. 4th Ed., Moscow,
Nauka, 1987, 246 p. (Rus)
87. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1981, vol. A23(6), pp. 1347-1357.
(Rus)
88. Yu.I. Matveev, A.A. Askadskii, I.V. Zhuravleva, G.L. Slonymskii, and V.V. Korshak,
Vysokomol. Soedin., 1981, vol. A23(9), pp. 2013-2026. (Rus)
89. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1982, vol. A24(6), pp. 1295-1303.
(Rus)
90. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1985, vol. A27(11), pp. 2357-2362.
(Rus)
91. T.P. Matveeva, Yu.I. Matveev, and A.A. Askadskii, Mechan. Composite Mater., 1986, No. 2, pp.
201-206. (Rus)
92. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1986, vol. A28(7), pp. 1365-1372.
(Rus)
93. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1989, vol. A31(3), pp. 617-622. (Rus)

672

94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.

114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.

Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1989, vol. A31(3), pp. 526-532. (Rus)
Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1994, vol. A36(3), pp. 436-443. (Rus)
Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soedin., 1993, vol. A35(1), pp. 63-67. (Rus)
M.S. Matevosyan, A.A. Askadskii, and G.L. Slonymskii, Vysokomol. Soedin., 1987, vol. A29(4),
pp. 761-767. (Rus)
A.M. Muzafarov, E.A. Rebrov, and V.S. Papkov, Uspekhi Khimii, 1991, vol. 60(7), pp. 15961612. (Rus)
A.F. Nikolaev, Synthetic polymers and plastics on their base. 2nd Ed., Moscow-Leningrad,
Khimia, Leningrad dept., 1966, 768 p. (Rus)
R.S. Norkina, S.I. Sokolov, and N.A. Shchegolevskaya, Izv. VUZov, Khimia i Khim. Tekhnol.,
1964, vol. 7(4), pp. 645-650. (Rus)
V.A. Pankratov, A.A. Askadskii, M.M. Patsuria, Yu.S. Kochergin, S.V. Vinogradova, G.L.
Slonymskii, and V.V. Korshak, Vysokomol. Soedin., 1977, vol. A19(1), pp. 142-148. (Rus)
G.L. Slonymskii, V.V. Korshak, S.V. Vinogradova, A.I. Kitaigorodskii, A.A. Askadskii, S.N.
Salazkin, and E.M. Belavtseva, Vysokomol. Soedin., 1967, vol. A9(2), pp. 851-854. (Rus)
G.L. Slonymskii, V.V. Korshak, A.I. Mzhelskii, A.A. Askadskii, Ya.S. Vygodskii, and S.V.
Vinogradova, Doklady AN SSSR, 1968, vol. 182(4), pp. 851-854. (Rus)
G.L. Slonymskii, and V.O. Shestopal, Vysokomol. Soedin., 1978, vol. A20(8), pp. 1712-1721.
(Rus)
G.L. Slonymskii, and A.A. Askadskii, Vysokomol. Soedin., 1986, vol. A28(5), pp. 1014-1022.
(Rus)
T.A. Speranskaya, and L.N. Tarutina, Optical properties of polymers. Leningrad, Khimia,
Leningrad dept., 1976, 136 p. (Rus)
A.A. Tager, M.V. Tsilipotkina, and A.I. Suvorova, Doklady AN SSSR, 1958, vol. 120(3), pp.
570-572. (Rus)
A.A. Tager, and L.K. Kolmakova, Vysokomol. Soedin., 1980, vol. A22(3), pp. 483-496. (Rus)
B.Ya. Teitelbaum, Thermomechanical analysis of polymers. Moscow, Nauka, 1979, 234 p. (Rus)
S.A. Tishin, V.A. Tishin, and B.A. Shpilevskii, Mechan. Composite Mater., 1990, No. 1, pp. 3-8.
(Rus)
S.A. Tishin, Appl. and Tech. Exper., 1990, No. 5, pp. 171-174. (Rus)
Yu.S. Urzhumtsev, and R.D. Maksimov, Prognosis of deformation of polymeric materials. Riga,
Zinatne, 1975, 416 p. (Rus)
Achievements in the field of synthesis of elementorganic polymers. pp. 143-231. V.V.Korshak,
A.I. Kirilin et al.; Ed. V.V. Korshak, Nesmeyanovs Institute of Organo-Element Compounds, AS
USSR. Moscow, Nauka, 1988, 319 p. (Rus)
R.F. Feinman, Statistic mechanics: the course of lectures. (Transl. From Engl.) Ed. Prof. D.N.
Zubarev, 2nd Ed., Moscow, Mir, 1978, 407 p. (Rus)
M.V. Tsilipotkina, A.A. Tager, O.V. Nechaeva, A.V. Vasiliev, L.M. Bykova, S.M. Bulygina,
V.A. Vasnev, and S.N. Salazkin, Vysokomol. Soedin., 1972, vol. A14(9), pp. 2053-2057. (Rus)
V.I. Sharkov, N.I. Kuibina, and Yu.P. Solovieva, J. Appl. Chem., 1965, vol. 35(5), pp. 11191128. (Rus)
V.I. Sharkov, N.I. Kuibina, and Yu.P. Solovieva, J. Appl. Chem., 1967, vol. 40(11), pp. 26092611. (Rus)
V.I. Sharkov, and N.I. Kuibina, Chemistry of Hemicelluloses. Moscow, Lesnaya
Promyshlennost, 1972, 440 p. (Rus)
V.O. Shestopal, Phys. Solids, 1970, vol. 12(1), pp. 291-294. (Rus)
A. Einstein, Collection of scientific works. Vol. 3. Works on kinetic theory, theory of radiation
and grounds of quantum mechanics. Moscow, Nauka, 1966, 631 p. (Rus)
Experimental methods in adsorption and molecular chromatography. Ed. A.V. Kiselev and V.P.
Dreving, Moscow, Izd. Moscow Univ., 1973, 433 p. (Rus)
Encyclopedia of polymers. Vol. I. 1043 p., Moscow, Sov. Enciklop., 1972. (Rus)
P. Arifov, S. Vasserman, and S. Tishin, Phys. Stat. Sol. A-Appl. Res., 1987, vol. 102(2), pp. 565570.
A.A. Askadskii, Physical properties of polymers. Encyclopedia of Fluid Mechanics. Vol. 9.
Polymer Flow Engineering. Ed. N.P. Cheremisinoff, 1990, pp. 103-147.
A.A. Askadskii, Analysis of the structure and properties of high-crosslinked polymer networks.
Chemistry Review, vol. 16, part 3, Harwood Academic Publishers, 1992, 120 p.
A.A. Askadskii, E.G. Galpern, I.V. Stankevich, and A.L. Chistyakov, Computer simulation of
polymers with predetermined physical properties. J. Biochem. Kinet., 1992, vol. 2(4), pp. 225232.

673

127. A.A. Askadskii, A.E. Shvorak, T.M. Frenkel, T.M. Babshinitzer, C.A. Bychko, O.V. Kovriga,
V.A. Pankratov, A.V. Rajulu, J. Appl. Polymer Sci., 1995, vol. 55(8), pp. 1173-1183.
128. A.A. Askadskii, Physical properties of polymers, prediction and control. Gordon and Breach
Publishers, Amsterdam, 1996, 333 p.
129. A.A. Askadskii, and L.M. Goleneva, Macromolecular Symposia Nano-Structures in Polymer
Systems, 1966, vol. 106(9), April, pp. 9-17.
130. G.O. Aspinal, Constitution of plant cell wall Polysaccharides. Encyclopedia of Plant Physiology,
Ed. W. Tanner and E.A. Loewus, Berlin, etc., Springer Verlag, 1981, vol. 13B. Plant
Carbohydrates, 11, Extracellular Carbohydrates, pp. 5-8.
131. M. Bank, J. Leffmgwell, and C. Thies, Macromolecules, 1971, vol. 4(1), pp. 43-46.
132. R.G. Beaman, J. Polym. Sci., 1952, vol. 9(5), pp. 470-472.
133. J. Bicerano, Prediction of polymer properties. New-York, Marcel Dekker Inc., 1996, 528 p.
134. F.J. Calleja, J. Serna, J. Vicente, and M.A. Segovia, J. Appl. Phys., 1985, vol. 58(1), pp. 253-259.
135. F.-L. Chen, E.M. Pearce, and T.K. Kwei, Polymers, 1988, vol. 29(12), pp. 2285-2289.
136. Y.Y. Chien, E.M. Pierce, and T.K. Kwei, Macromolecules, 1988, vol. 21(6), pp. 1616-1619.
137. E.Y. Chu, E.M. Pearce, T.K. Kwei, T.F. Yeh, and Y. Okamoto, Makromol. Chem., Rapid
Commun., 1991, vol. 12(1), pp. 1-4.
138. Y.K. Dai, E.Y. Chu, Z.S. Xu, E.M. Pearce, Y. Okamoto, and T.K. Kwei, J. Polym. Sci., Part A,
Polym. Chem., 1994, vol. 32, pp. 397-400.
139. M.A. de Araujo, R. Stadler, and H.-J. Cantow, Polymer, 1988, vol. 29(12), pp. 2235-2243.
140. M. Eldrup, D. Lightbood, and J.N. Sherwood, Chem. Phys., 1981, vol. 63(1-2), pp. 51-58.
141. R.F. Fedors, Polymer Eng. and Sci., 1974, vol. 14(2), pp. 147-154.
142. P.J. Flory, Faraday Disc. Chem. Soc., 1979, No. 68, pp. 14-25.
143. F.M. Fowkes, Indust. Eng. Chem., 1964, vol. 56(12), p. 40.
144. T.G. Fox, and P.J. Flory, J. Appl. Phys., 1950, vol. 21(6), pp. 581-591.
145. T.G. Fox, and S. Loshaek, J. Polym. Sci., 1955, vol. 15(80), pp. 371-390.
146. R.J. Gaylord, B. Joss, J.T. Bendler, and E.A. di Marzio, Brit. Polymer J., 1985, vol. 17(2), pp.
126-128.
147. L.A. Girifalco, and R.J. Good, J. Phys. Chem., 1957, vol. 61(7), pp. 904-909.
148. R.J. Good, L.A. Girifalco, and G.A. Kraus, J. Phys. Chem., 1958, vol. 62(11), pp. 1418-1421.
149. R.J. Good, and L.A. Girifalco, J. Phys. Chem., 1960, vol. 64(5), pp. 561-565.
150. J.L. Gordon, Cohesive energy density. Encyclopedia of Polymer Science and Technology, vol. 3,
New-York, John Wiley and Sons, Inc., 1965, pp. 833-862.
151. T.C. Gsell, E.M. Pearce, and T.K. Kwei, Polymer, 1991, vol. 9, pp. 1663-1674.
152. A.M. Gupta, and C.W. Macosko, Macromolecules, 1993, vol. 26(10), pp. 2455-2463.
153. J.-L. Halary, F.C. Ben Larbi, P. Oudin, and L. Monnerie, Macromol. Chem., 1988, vol. 189(9),
pp. 2117-2124.
154. R.A. Haldon, and R. Simha, J. Appl. Phys., 1968, No. 39, pp. 1890-1898.
155. A.E. Hamielec, M. Eldrup, O.E. Mogensen, and P. Jansen, J. Macromol. Sci., 1973, vol. 9(2), pp.
305-337.
156. C.K. Hu, and G.R. Gruzalsk, Phys. Rev., 1983, vol. B27(1), pp. 1-8.
157. W. Huang, K.J. Zhu, E.M. Pearce, and T.K. Kwei, J. Appl. Sci., 1993, vol. 48, pp. 563-573.
158. L. Jong, E.M. Pearce, T.K. Kwei, and L.C. Dickinson, Macromolecules, 1990, vol. 23(24), pp.
5071-5074.
159. L. Jong, E.M. Pearce, and T.K. Kwei, Polymer, 1993, vol. 34(1), pp. 48-55.
160. G. Kanig, J. Polym. Sci., part C, 1967, pp. 1957-1964.
161. H.-L. Kim, T.K. Kwei, and E.M. Pearce, Hydrogen bonding interaction in Interpenetrating
polymer network. In: Contemporary Topics in Polymer Science, vol. 6 (Ed. B. Culberison),
Plenum, New York, 1989, pp. 687-697.
162. H.-L. Kim, E.M. Pearce, and T.K. Kwei, Macromolecules, 1989, vol. 22(8), pp. 3374-3380.
163. H.-L. Kim, E.M. Pearce, and T.K. Kwei, Macromolecules, 1989, vol. 22(8), pp. 3498-3500.
164. P. Kindl, and G. Reiter, Phys. Stat. Sol. A Appl. Res., 1987, vol. 104(2), pp. 707-713.
165. D.L. Kotzev, E.M. Pearce, and T.K. Kwei, J. Appl. Polym. Sci., 1984, vol. 29, pp. 4443-4448.
166. M. Kryszewski, J. Jachwicz, M. Malanga, and O. Vogl, Polymer, 1982, vol. 23(2), pp. 271-276.
167. K. Kun, and R. Kunin, J. Polymer Sci., 1964, vol. B2, p. 587.
168. T.K. Kwei, T. Nishi, and R.F. Roberts, Macromolecules, 1974, vol. 7(5), pp. 667-674.
169. T.K. Kwei, and H.L. Frisch, Macromolecules, 1978, vol. 11(6), pp. 1267-1271.
170. T.K. Kwei, E.M. Pearce, and B.Y. Min, Macromolecules, 1985, vol. 18(11), pp. 2326, 2327.
171. T.K. Kwei, E.M. Pearce, F. Ren, and J.P. Chen, J. Polym. Sci., Part B, Polym. Phys., 1986, vol.
24(7), pp. 1597-1609.

674

172. W.A. Lee, and G.J. Knight, Brit. Polymer J., 1970, vol. 2(2), pp. 73-80.
173. P. Lin, C. Clash, E.M. Pearce, and T.K. Kwei, J. Polym. Sci., Part B, Polym. Phys., 1988, vol. 26,
pp. 603-619.
174. Y.-H. Lin, Macromolecules, 1984, vol. 17(12), pp. 2846-2856.
175. Y.-H. Lin, Macromolecules, 1986, vol. 19(1), pp. 159-168.
176. Y.-H. Lin, Macromolecules, 1991, vol. 19(1), pp. 168-173.
177. Y.-H. Lin, Macromolecules, 1991, vol. 24(19), pp. 5346-5350.
178. S. Lu, E.M. Pearce, and T.K. Kwei, Macromolecules, 1993, vol. 26(4), pp. 3514-3518.
179. S. Lu, E.M. Pearce, and T.K. Kwei, J. Polym. Sci., Part A, Polym. Chem., 1994, vol. 32, pp.
2597-2601.
180. S. Lu, E.M. Pearce, and T.K. Kwei, J. Macromol. Sci., Pure Appl. Chem., 1994, vol. A31(11), pp.
1535-1560.
181. S. Lu, M.M. Melo, J. Zhao, E.M. Pearce, and T.K. Kwei, Macromolecules, 1995, vol. 28(14), pp.
4908-4913.
182. S. Lu, E.M. Pearce, and T.K. Kwei, Polymer Eng. Sci., 1995, vol. 35(13), pp. 1113-1116.
183. S. Lu, E.M. Pearce, and T.K. Kwei, Polym. Adv. Tech., 1996, vol. 7, pp. 553-559.
184. S. Lu, E.M. Pearce, and T.K. Kwei, J. Polym. Sci., Part A, Polym. Chem., 1996, vol. 34, pp.
3163-3172.
185. D. Luo, E.M. Pearce, and T.K. Kwei, Macromolecules, 1993, vol. 26(23), pp. 6220-6225.
186. S. Makhija, E.M. Pearce, T.K. Kwei, and F. Liu, Polym. Eng. Sci., 1990, vol. 30(13), pp. 798801.
187. B.D. Malhotra, and R.A. Pethrick, Eur. Polym. J., 1983, vol. 19(6), pp. 457-459.
188. J.E. Mark, and J.L. Sullivan, J. Chem. Phys., 1977, vol. 66(3), pp. 1006-1011.
189. R.L. Miller, and R.F. Boyer, J. Polymer Sci., Polymer Phys. Ed., 1984, vol. 22(12), pp. 20432050.
190. K. Ogawa, F. Tanaka, J. Tamura, K. Kadowaki, and K. Okamura, Macromolecules, 1987, vol.
20(5), pp. 1174-1176.
191. H. Park, E.M. Pearce, W.H. Starnes, and T.K. Kwei, J. Polym. Sci., Part A, Polym. Chem., 1990,
vol. 28, pp. 1079-1089.
192. J.R. Pennacchia, E.M. Pearce, T.K. Kwei, B.J. Bulkin, and J.-P. Chen, Macromolecules, 1986,
vol. 19(4), pp. 973-977.
193. E.M. Pearce, T.K. Kwei, and T.K. Min, J. Macromol. Sci. Chem., 1984, vol. A21(8,9), pp.
1181-1216.
194. E.M. Pearce, and T.K. Kwei, J. Macromol. Sci. Chem., 1991, vol. A28(11,12), pp. 1207-1212.
195. E.M. Pearce, and T.K. Kwei, Polym. Solutions, Blends and Interfaces, 1992, pp. 133-149.
196. E.M. Pearce, T.K. Kwei, and S. Lu, Polym. Adv. Techn., 1994, vol. 5, pp. 600-602.
197. W.A. Prest, and R.S. Porter, J. Polym. Sci., Part A-2, Polym. Phys., 1972, vol. 10(9), pp. 16391647.
198. K. Schmidt-Rohr, J. Clauss, and H.W. Spiess, Macromolecules, 1992, vol. 25(12), pp. 3273-3277.
199. H.A. Schneider, and H.J. Brekner, Polymer Bull., 1985, vol. 14(2), pp. 173-178.
200. H.A. Schneider, and B. Leikauf, Thermochim. Acta, 1987, vol. 114(1), pp. 165-170.
201. B.L. Schuerman, Int. Conference St-Petersburg Moscow, May 18-26, 1995, OL70.
202. A. Shefer, and M. Gottlieb, Macromolecules, 1992, vol. 25(15), pp. 4036-4042.
203. A.R. Shultz, and B.M. Gendron, J. Appl. Polymer Sci., 1972, vol. 16(2), pp. 461-471.
204. A.R. Shiltz, and A.L. Young, J. Appl. Polym. Sci., 1983, vol. 28(5), pp. 1677-1684.
205. R. Simha, J. Chem. Phys., 1962, vol. 37(5), p. 1003.
206. P.A. Small, J. Appl. Chem., 1953, vol. 3(2), pp. 71-80.
207. T. Suzuki, E.M. Pearce, and T.K. Kwei, Polymer, 1992, vol. 33(1), pp. 198-201.
208. T.E. Timell, Wood hemicellulose. Adv. In Carbohydrate Chem., New York, 1964, vol. 19, pp.
247-302.
209. S.P. Ting, E.M. Pearce, and T.K. Kwei, J. Polym. Sci., Polym. Lett., 1980, vol. 18(3), pp. 201209.
210. S.P. Ting, B.J. Bulkin, and E.M. Pearce, J. Polym. Sci., Part A, Polym. Chem., 1981, vol. 19(6),
pp. 1451-1473.
211. S.A. Tishin, E.S. Bakaev, and B.A. Shipilevskij, Plaste und Kautschuk, 1991, Heft 4, SS. 45-49.
212. D.A. Tomalia, A.M. Noylor, and A.W. Doddard, Angew. Chem., 1990, vol. 29(2), pp. 138-175.
213. M.M. Vaidya, K. Levon, and E.M. Pearce, J. Polym. Sci., Part B, Polym. Phys., 1995, vol. 33, pp.
2093-2108.
214. D.W. Van Krevelen, Properties of polymers. Third Edition, Elsevier, Amsterdam, 1990, p. 875.

675

215. I.I. Vointseva, and A.A. Askadskii, Interpolymers (paired polymers). Chemistry reviews, Soviet
scientific reviews. Vol. 16, part 2, Harwood Academic Publishers, Chur-Reading, Paris,
Philadelphia, Tokyo, Melbourne, 1991, p. 86.
216. L.F. Wang, E.M. Pearce, and T.K. Kwei, J. Polym. Sci., Part C, Polym. Lett., 1990, vol. 28, pp.
317-321.
217. L.F. Wang, E.M. Pearce, and T.K. Kwei, J. Polym. Sci., Part B, Polym. Phys., 1991, vol. 29, pp.
619-626.
218. T.P. Yang, E.M. Pearce, T.K. Kwei, and N.L. Yang, Macromolecules, 1989, vol. 22(4), pp. 18131818.
219. X. Yang, P.C. Painter, M.M. Coleman, E.M. Pearce, and T.K. Kwei, Macromolecules, 1992, vol.
25(8), pp. 2156-2165.
220. A.F. Yee, R.J. Bankert, K.L. Ngai, and R.W. Rendell, J. Polymer Sci., 1988, vol. 26(12), pp.
2463-2483.
221. K.J. Zhu, E.M. Pearce, and T.K. Kwei, J. Appl. Polym. Sci., 1989, vol. 37(2), pp. 573-578.
222. K.J. Zhu, S.F. Chen, T. Ho, E.M. Pearce, and T.K. Kwei, Macromolecules, 1990, vol. 23(1), pp.
150-154.
223. H. Zhuang, E.M. Pearce, and T.K. Kwei, Macromolecules, 1994, vol. 27(22), pp. 6398-6403.
224. H. Zhuang, E.M. Pearce, and T.K. Kwei, Polymer, 1995, vol. 36(11), pp. 2237-2241.
225. M. Doi, and S.F. Edwards, J. Chem. Soc. Faraday Trans. II, 1978, vol. 74(10), p. 1818.
226. M. Doi, J. Polym. Sci., Polym. Phys. Ed., 1983, vol. 21(5), p. 667.
227. M.J. Doi, J. Polym. Sci., Polym. Phys. Ed., 1980, vol. 18(5), p. 1005.
228. M. Doi, and S.F. Edwards, The Theory of Polymer Dynamics, Oxford Univ. Press, 1986.
229. D. Richter, B. Farago, R. Bitera, L.J. Fetters, J.S. Huang, and B. Ewen, Macromolecules, 1993,
vol. 26(4), p. 795.
230. S. Wu, J. Polym. Sci., Polym. Phys. Ed., 1989, vol. 27(4), p. 723.
231. K. Iwata, and S.F. Edwards, J. Chem. Phys., 1989, vol. 90(8), p. 4567.
232. Yu.I. Matveev, and A.A. Askadskii, Vysikomol. Soedin., 1993, vol. A35(1), p. 63. (Rus)
233. A.V. Vologodskii, A.V. Lukashin, M.D. Frank-Kamenetskii, and V.V. Anshelevich, Zh. Eksper. I
Teor. Fiz., 1974, vol. 66(6), p. 2153 (Rus); M.D. Frank-Kamenetskii, A.V. Lukashin, and A.V.
Vologodskii, Nature, 1975, vol. 258(3), p. 398; Frank-Kamenetskii, and A.V. Vologodskii,
Uspekhi Fiz. Nauk, 1981, vol. 139(4), p. 641. (Rus)
234. A.A. Askadskii, A.I. Kitaigorodskii, and G.L. Slonimskii, Vysokomol. Soedin., 1975, vol.
A17(10), p. 2293. (Rus)
235. A.A. Askadskii, and Yu.I. Matveev, Chemical Structure and Physical Properties of Polymers,
Moscow, Khimia, 1983. (Rus)
236. M. Nakato, T. Nakagama, Y. Nakamura, and S. Wakatsuki, Kinetics of Chain Aggregation of
Poly(Methyl Methacrylate) in Isoamyl Acetate, J. Chem. Phys., 1999, vol. 110(5), pp. 2711-2716.
237. V.A. Kargin, A.I. Kitaigorodskii, and G.L. Slonymskii, Colloid. Zh., 1957, vol. 19(2), pp. 131.
(Rus)
238. G.S.Y. Yeh, Pure and Appl. Chem., 1972, vol. 32(1/2), pp. 65.
239. S.A. Arzhakov, N.F. Bakeev, and V.A. Kabanov, Vysokomol. Soed., 1973, vol. A15(5), pp. 1154.
(Rus)
240. I.M. Lifshitz, Zh. Experim. I Teor. Fiz., 1968, vol. 55, vol. 6(12), pp. 2408. (Rus)
241. I.M. Lifshitz, and A.Yu. Grosberg, Zh. Exper. I Theor. Fiz., 1973, vol. 65(6 12), pp. 2399. (Rus)
242. A.Yu. Grosberg, and A.R. Khokhlov, In: Problems in Solid-state Physics, Ed. A.M. Prokhorov,
A.S. Prokhorov, Moscow, Mir, 1984, 368 p. (Rus)
243. T.M. Birstein, ans O.B. Ptyitsin, Conformations of Macromolecules, Moscow Leningrad, Nauka,
1964, 391 p. (Rus)
244. G.L. Slonimskii, V.V. Korshak, S.V. Vinogradova, A.I. Kitaigorodskii, A.A. Askadskii, S.N.
Salazkin, and E.M. Belavtseva, Vysokomol. Soed., 1967, vol. A9(2), pp. 402. (Rus)
245. A.P. Khokhlov, In: Model Approach in Statistic Mechanics, Moscow, Izd. MSU, 1979. (Rus)
246. T.M. Birshtein, A.M. Skvotsov, and A.A. Sariban, Vysokomol. Soed., 1977, vol. A19(1), pp. 63.
(Rus)
247. S.R. Rafikov, V.P. Budtov, and Yu.B. Monakov, Introduction into Physicochemistry of Solutions
of Polymers, Moscow, Nauka, 1978, 328 p. (Rus)
248. A.A. Askadskii, and Yu.I. Matveev, Chemical Structure and Physical Properties of Polymers,
Moscow, 1983, 248 p. (Rus)
249. T.P. Matveeva, Yu.I. Matveev, and A.A. Askadskii, Mechanics of Composite Materials, 1986, No.
2, pp. 201. (Rus)
250. A.A. Askadskii, and Yu.I. Matveev, Doklady AN SSSR, 1986, vol. 290(1), pp. 135. (Rus)

676

251. Yu.A. Matveev, and A.A. Askadskii, Vysokomol. Soed., 1987. (Rus)
252. A.A. Askadskii, A.S. Marshalovich, T.P. Matveeva, A.V. Pastukhov, and G.I. Sidorova,
Mechanics of Composite Materials, 1983, No. 5, pp. 906. (Rus)
253. S.Ya. Frenkel, Introduction into Statistic Theory of Polymerization, Leningrad, 1965. (Rus)
254. E.B. Zhulina, and T.M. Birstein, Vysokomol. Soed. 1986, vol. A28(12), pp. 2589. (Rus)
255. P. de Jen, Ideas of Scaling in Physic of Polymers, 1982, Moscow, 368 p. (Rus)
256. I.M. Lifshits, Uspekhi Fiz. Nauk, 1979, vol. 127(3), pp. 353. (Rus)
257. L.D. Landau, and E.M. Lifshits, Statistic Physics, 1964, Moscow, 568 p. (Rus)
258. G.Z. Gershuni, and E.M. Zhukhovskaya, Convective Stability of Incompressible Liquid, 1972,
Moscow, 392 p. (Rus)
259. M. Kleman, and J.E. Sados, Phys. Letters, 1979, vol. 40, p. 569; J.F. Sados, and R. Mosseri, Phys.
Mag., 1982, vol. B45, p. 467.
260. M. Kleman, J. Phys., 1982, vol. 43, p. 1389.
261. D.R. Nelson, Phys. Rev. Letters, 1983, vol. 50, p. 982.
262. V.N. Novikov, Zh. Eksper. I Teor. Fiz., 1984, vol. 87, No. 3(9), p. 1080. (Rus)
263. Yu. I. Matveev, and A.A. Askadskii, Vysokomol. Soed., 1986, vol. A28(7), p. 1365. (Rus)
264. A.A. Askadskii, and Yu.I. Matveev, Doklady AN SSSR, 1986, vol. 290(1), p. 135. (Rus)
265. R. Grigorovici, Amorphous and Liquid Semiconductors, 1974, New York, 320 p.
266. N.V. Efimov, Mathematical Encyclopedia, 1984, vol. 4, p. 163. (Rus)
267. A. Brus, and R. Kauli, Structure Phase Transitions, 1984, Moscow, 408 p. (Rus)
268. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soed., 1989, vol. A31(3). (Rus)
269. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soed., 1981, vol. A23(6), p.1347. (Rus)
270. M.A. Isakovich, and I.A. Chaban, Zh. Eksper. I Teor. Fiz., 1966, vol. 50(5), p. 1343. (Rus)
271. I.A. Chaban, Doctor Thesis, Moscow, N.N. Andreev Acoustic Institute, 1984, 324 p. (Rus)
272. V.G. Rostiashvili, A.R. Nekhoda, and V.I. Irzhak, Fluctuation Theory of Relaxation and Dynamic
Properties of Amorphous Polymers in the Area of Glass Transition, 1985, Chernogolovka, 324 p.
(Rus)
273. L.E. Nielsen, Mechanical Properties of Polymers, Reinhold, New York, 1962.
274. R.F. Boyer, Rubber Chem. Technol., 1963, vol. 63, p. 1303.
275. J.H. Sewell, Royal Aircraft Establishment, Farnborough, Hants. UK, Unpublished report.
276. L. Turunen, Kunststoffe, 1962, vol. 52, p. 672.
277. R.H. Wiley, and G.M. Brauer, J. Polym. Sci., 1948, vol. 3, p. 647.
278. P.L. Magagnini, Chim. Ind. (Milan), 1967, vol. 49, p. 1041.
279. W.J. Jackson, T.F. Gray, and J.R. Caldwell, J. Appl. Polym. Sci., 1970, vol. 14, p. 685.
280. O.G. Lewis, "Physical Constants of Linear Homopolymers", Springer-Verlag, 1968.
281. B.G. Ranby, K.S. Chan, and H. Brumberger, J. Polym. Sci., 1962, vol. 58, p. 545.
282. A. Quach, and R. Shima, J. Appl. Phys., 1971, vol. 42, p. 4592.
283. H.G. Killian, and E. Jenckel, in: K.A. Wolf (Ed.), "Struktur und Physikalisches Verhalten der
Kunststoffe", Springer-Verlag, 1962, p. 176.
284. G.A. Gordon, J. Polym. Sci., 1971, vol. A-2(9), p. 1693.
285. S.C. Temin, J. Polym. Sci., 1965, vol. 9, p. 471.
286. R.F. Boyer, Changements de Phases, Soc. Chim. Phys., Paris, 1952, p. 383.
287. W.G. Barb, J. Polym. Sci., 1959, vol. 37, p. 515.
288. S.Y. Chuang, S.J. Tao, and J.M. Wilkenfeld, 1972, vol. 43, p. 737.
289. J.N. Gayles, and W.L. Peticolas, Light Scattering Spectr. Solids, Proc. Intern. Conf., 1968, Ed.
G.B. Wright, Springer-Verlag, New York, 1969, p. 715; Chem. Abstr., 1970, vol. 72, p. 22089z.
290. B.M. Grieveson, Polymer, 1960, vol. 1, p. 499.
291. D.W. Brown, and L.A. Wall, J. Polym. Sci., 1969, vol. A-2(7), p. 601.
292. G. Allen, J. McAinsh, and G.M. Jeffs, Polymer, 1971, vol. 12, p. 85.
293. IUPAC, "Nomenclature of Organic Chemistry", Sections A and B, Third Edit., Sect. C, Second
Edit., London, Butterworths, 1969.
294. J.M. Augl, and J.V. Duffy, J. Polym. Sci., 1971, vol. A-1(9), p. 1343.
295. G.F.L. Ehlers, and K.R. Fisch, Appl. Polym. Symp., 1969, vol. 8, p. 171.
296. V.V. Korshak, V.M. Mamedov, G.E. Golubkov, and D.R. Tur, Vysokomol. Soedin., 1970, vol.
B12, p. 57. (Rus)
297. R.N. Johnson, A.G. Farnham, R.A. Glendinning, W.F. Hale, and C.N. Merriam, J. Polym. Sci.,
1967, vol. A-1(5), p. 2375.
298. J.K. Gilham, and M.B. Roller, Polym. Eng. Sci., 1971, vol. 11, p. 295.
299. W.M. Edwards, and I.M. Robinson, US Pat. 2,710,853, 14 June 1955.
300. K.W. Doak, Modern Plastics Encyclopedia, 1967, vol. 45(1A), p. 14.

677

301. J.M. Barton, and J.P. Crichley, Polymer, 1970, vol. 11, p. 212.
302. S.L. Cooper, A.D. Mair, and A.V. Tobolsky, Textile Res. J., 1965, vol. 35, p. 1110.
303. J.H. Freeman, L.W. Frost, G.M. Bower, and E.J. Traynor, Conf. Structural Plastics, Adhesives,
Filament Wound Composites, WPAFB, Dayton, Ohio, 1962, vol. 1, p. 30.
304. G.A. Bernier, and D.E. Kline, J. Appl. Sci., 1968, vol. 12, p. 593.
305. A.D. Mair, M.C. Shen, and A.V. Tobolsky, AD 604010, 1964, "High temperature polymers: HFilm and SP-Polymer", ONR Tech. Rept. RLT-83.
306. C.E. Sroog, J. Polym. Sci., 1967, vol. C16, p. 1191; data quoted from W.E. Tatum, L.E. Amborski,
C.W. Gerow, J.F. Heacock, and R.S. Mallouk, Electrical Insulation Conf., Chicago, 1963.
307. A.V. Sidorovich, and Ye.V. Kuvshinskii, Polym. Sci. USSR, 1968, vol. 10, p. 1627; (Transl'n of
Vysokomol. Soedin., 1968, vol. A10, p. 1401. (Rus)).
308. W. Wrasidlo, and J.M. Augl, J. Polym. Sci., 1969, vol. A-1(7), p. 321.
309. G.A. Progany, Polymer, 1970, vol. 11, p.66.
310. J.M. Augl, J. Polym. Sci., 1970, vol. A-1(8), p. 3145.
311. W.Wrasidlo, J. Polym. Sci., 1970, vol. A-1(8), p. 1107.
312. W. Wrasidlo, and J.M. Augl, J. Polym. Sci., 1969, vol. B7, p. 281.
313. P.W. Morgan, J. Polym. Sci., 1964, vol. A2, p. 437.
314. J.P. Critchley, V.C.R. McLoughlin, J. Thrower, and I.M. White, Brit. Polym. J., 1970, vol. 2, p.
288.
315. I.M. Dolgopol'skii, Kauchuk i Rezina, 1971, vol. 30(2), p. 31. (Rus) (transl'n in Soviet Rubber
Technol., 1971, vol. 30(2), p. 38).
316. G.L. Brode, ACS Polym. Preprints, 1965, vol. 6, p. 626.
317. R.G. Beaman, J. Appl. Polym. Sci., 1965, vol. 9, p. 3949.
318. A. Schiller, J.C. Petropoulos, and C.S.H. Chen, J. appl. Polym. Sci., 1964, vol. 8, p. 1699.
319. M.E. Baird, G.T. Goldsworthy, and C.J. Creasey, J. Polym. Sci., 1968, vol. B6, p. 737.
320. M.E. Baird, J. Polym. Sci., 1970, vol. A-2(8), p. 739.
321. A.B. Thompson, and D.W. Woods, Trans. Faraday Soc., 1956, vol. 52, p. 1383.
322. H.A. Vogel, J. Polym. Sci., 1970, vol. A-1(8), p. 2035.
323. Z. Dobkowski, B. Krajewski, and Z. Weilgosz, Polim. Tworz. Wielk., 1970, vol. 15, p. 428.
324. R.F. Clash, and L.M. Runkieweiz, Ind. Eng. Chem., 1944, vol. 36, p. 279.
325. K.E. Perepelkin, Mekh. Polim., 1971, vol. 5, p. 790. (Rus)
326. G.P. Mikhailov, A.I. Artyukhov, and V.A. Polym. Sci. USSR, 1969, vol. 11, p. 553 (transl'n of
Vysokomol. Soedin., 1969, vol. A11, p. 553).
327. F.E. Wiley, Ind. Eng. Chem., 1942, vol. 34, p. 1052.
328. W.J. Feast, W.K.R. Musgrave, and N. Reeves, J. Polym. Sci., 1971, vol. A-1(9), p. 2733.
329. N.H. Reinking, A.E. Bamabeo, and W.F. Hale, J. Appl. Polymer Sci., 1963, vol. 7, p. 2153.
330. J. Lai, and G.S. Trick, J. Polym. Sci., 1970, vol. A-1(8), p. 2339.
331. A.S. Hay, and R.F. Clark, Macromolecules, 1970, vol. 3, p. 533.
332. A.R. Blythe, and G.M. Jeffs, J. Macromol. Sci., 1969, vol. B3, p. 141.
333. D.J. Marks, Thesis, Manchester, Univ., 1961.
334. G.Allen, C. Booth, S.J. Hunt, C. Price, F. Vernon, and R.F. Warren, Polymer, 1967, vol. 8, p. 406.
335. P.G. Wapner, (supervisor: W.C. Forsman), PhD. Thesis, Univ. Pennsylvania, 1971; via Diss.
Abstr. Intern., 1971, vol. B32, p. 2332.
336. N. Doddi, W.C. Forsman, and C.C. Price, Macromolecules, 1971, vol. 4, p. 648.
337. H. Schnell, L. Bottenbruch, G. Darsow, and K. Weirauch, Brit. 1,400.00, 15, Jan. 1969.
338. P.E. Wei, and P.E. Butler, J. Polym. Sci., 1968, vol. A-1(6), p. 2461.
339. V. Frosini, E. Butta, and M. Calamia, J. Appl. Polym. Sci., 1967, vol. 11, p. 527.
340. F.S. Dainton, D,M, Evans, F.E. Hoare, and T.P. Meila, Polymer, 1962, vol. 3, p. 271.
341. D.J.H. Sandiford, J. Appl. Chem., 1958, vol. 8, p. 188.
342. K.-H. Helwege, J. Hennig, and W. Knappe, Kolloid Z.-Z. Polym., 1962, vol. 186, p. 29.
343. K.-H. Helwege, R. Hoffman, and W. Knappe, Kolloid Z.-Z. Polym., 1968, vol. 226, p. 109.
344. M. Barccaredda, E. Butta, and V. Frosini, Eur. Polym. J., 1966, vol. 2, p. 423.
345. A. Nagai, T. Ishibashi, and M. Takayanagi, Rept. Progr. Polym. Phys. (Japan), 1967, vol. 10, p.
341.
346. A.I. Marei, Ye.A. Sidorovich, G.Ye. Novikova, and E.I. Rodina, Polym. Sci. USSR, 1968, vol. 10,
p. 630 (transl'n of Vysokomol. Soedin., 1968, vol. 10, p. 542 (Rus)).
347. J.A. Faucher, and J.V. Koleske, Polymer, 1968, vol. 9, p. 44.
348. W.G. Miller, and J.H. Saunders, J. Appl. Polym. Sci., 1969, vol. 13, p. 1277.
349. Yu.K. Godovskii, and Yu.S. Lipatov, Polym. Sci. USSR, 1968, vol. 10, p. 34 (transl'n of
Vysokomol. Soedin., 1968, vol. 10, p. 32 (Rus)).

678

350. A. Conix, J. Polym. Sci., 1958, vol. 29, p. 343.


351. R.H. Gobran, and M.B. Berenbaum, ACS Meeting 13, April 1958; via E.R. Bertozzi, Rubber
Chem. Technol., 1968, vol. 41, p. 114.
352. A.V. Tobolsky, and W.J. MacKnight, in: H.F. Mark and E.H. Immergut (Eds.), "Polymer
Reviews", vol. 13: Polymeric Sulfur and Related Polymer, Interscience, 1965, p. 35.
353. G. Williams, Trans. Faraday Soc., 1963, vol. 59, p. 1397.
354. B.E. Read, Polymer, 1962, vol. 3, p. 529.
355. A. Mele, A. Delle Site, C. Bettinali, and A. Di Domenico, J. Chem. Phys., 1968, vol. 49, p. 3297.
356. J.L. Zollinger, J.R. Throckmorton, S.T. Ting, R.A. Mitsch, and D.E. Elrick, J. Macromol. Sci.,
1969, vol. A3, p. 1443.
357. W.A. Lee, Royal Aircraft Establishment, Farnsborough, Hants., UK - Unpublished results.
358. P. Johncock, M.A.H. Hewins, Royal Aircraft Establishment, Farnborough, Hants., UK Unpublished report.
359. N.H. Reinking, A.E. Bamabeo, and W.F. Hale, J. Appl. Polymer Sci., 1963, vol. 7, p. 2135.
360. K. Fujimoto, Chem. High Polym. (Japan), 1961, vol. 18, p. 415.
361. J. Heijboer, Brit. Polym., 1969, vol. 1, p. 3.
362. A.H. Willbourn, Kolloid Z. - Z. Polym., 1965, vol. 201, p. 20.
363. R.G. Saba, J.A. Sauer, and A.E. Woodward, J. Polym. Sci., 1963, vol. A1, p. 1483.
364. J.J. Stratta, F.P. Reding, and J.A. Faucher, J. Polym. Sci., 1964, vol. A2, p. 5017.
365. F.B. Cramer, and R.G. Beaman, J. Polym. Sci., 1956, vol. 21, p. 237.
366. W.W. Schartz, and R.D. Lowrey, J. Appl. Polym. Sci., 1967, vol. 11, p. 553.
367. P. Johncock, and P.A. Grattan, Royal Aircraft Establishment, Farnborough, Hants., UK Unpublished report.
368. P. Johncock, and J.D. Lee, Royal Aircraft Establishment, Farnborough, Hants., UK - Unpublished
report.
369. A.T. DiBenedetto, and K.L. Trachte, J. Appl. Polym. Sci., 1970, vol. 14, p. 2249.
370. E. Butta, Matter, Plastics, Elastomers, 1969, vol. 35, p. 1411.
371. P.A. Gratten, V.C.R. McLoughlin, and W.A. Lee, Royal Aircraft Establishment, Farnborough,
Hants., UK - Unpublished data.
372. L.W. Breed, R.L. Elliot, and M.E. Whitehead, J. Polym. Sci., 1967, vol. A-1(5), p. 2745.
373. L.T.C. Lee, and E.M. Pearce, J. Polym. Sci., 1971, vol. A-1(9), p. 557.
374. K. Saotome, and H. Komoto, J. Polym. Sci., 1966, vol. A-1(4), p. 1475.
375. J.C. Mileo, B. Sillion, and G. De Gaudemaris, Compt. Rend., 1969, vol. C268, p. 1949.
376. M.E.B. Jones, and I. Goodman, Brit. Pat. 1,128,807, 2, Oct. 1968.
377. D. Brown, M.E.B. Jones, and W.R. Maltman, J. Polym. Sci., 1968, vol. B6, p. 635.
378. V. Frosini, M. Pasquini, and E. Butta, Chim. Ind., 1971, vol. 53, p. 140.
379. E. Butta, S. de Petris, V. Frosini, and M. Pasquini, Eur. Polym. J., 1971, vol. 7, p. 387.
380. G.F.L. Ehlers, R.C. Evers, and K.R. Fisch, J. Polym. Sci, 1969, vol. A-1(7), p. 3413.
381. B.E. Jennings, M.E.B. Jones, and J.B. Rose, J. Polym. Sci, 1967, vol. C16, p. 715.
382. M.E.A. Cudby, R.G. Feasey, B.E. Jennings, M.E.B. Jones, and J.B. Rose, Polymer, 1965, vol. 6, p.
589.
383. J.E. Kurz, J.C. Woodbrey, and M. Ohta, J. Polym. Sci., 1970, vol. A-2(8), p. 1169.
384. H. Schnell, Ind. Eng. Chem., 1959, vol. 51, p. 157.
385. A.N. Perepelkin, and P.V. Kozlov, Polymer Sci. USSR, 1966, vol. 8, p. 56 (transl'n of Vysocomol.
Soedin., 1966, vol. 8, p. 56 (Rus)).
386. F.P. Reding, J.A. Faucher, and R.D. Whitman, J. Polym. Sci., 1961, vol. 54, p. S56.
387. L.J. Garfield, J. Polym. Sci., 1970, vol. C30, p. 561.
388. H. Schnell, "Chemistry and Physics of Polycarbonates", 1964, Interscience, New York, Chap. III.
389. B.F. Malichenko, V.I. Feoktistova, and A.E. Nesterov, Vysokomol. Soedin., 1969, vol. B11, p.
543. (Rus)
390. E.N. Elam, J.C. Martin, and R. Gilkey, US Pat. 3,313,777, 11 April, 1967 (Brit. 962,913).
391. W. Gawlak, R.P. Palmer, J.B. Rose, D.J. Sandford, and A. Turner-Jones, Chem. Ind., 1962, vol.
25, p. 1148.
392. A.A. D'Onofrio, Brit. Pat. 1,011,283, 24 Nov., 1965.
393. O.V. Smirnova, El Said Ali Kasan, I.P. Losev, and G.S. Kolesnikov, Polymer Sci. USSR, 1965,
vol. 7, p. 557 (transl'n of Vysokomol. Soedin., 1965, vol. 7, p. 503 (Rus)).
394. H. Schnell, Angew. Chem., 1956, vol. 68, p. 633.
395. US Patent 3,546,165, Dec. 8, 1970, assigned to P.W. Morgan.
396. D. Stefan, H.L. Williams, D.R. Renton, and M.M. Pintar, J. Macromol. Sci., 1970, vol. B4, p. 853.
397. D.W. McCall, and D.R. Falcone, Trans. Faraday Soc., 1970, vol. 66, p. 262.

679

398. S. Matsuoka, and Y. Ishida, J. Polym. Sci., 1966, vol. C14, p. 247.
399. F. Krum, and F.H. Muller, Kolloid. Z., 1959, vol. 164, p. 81.
400. G.P. Mikhailov, and M.P. Eidel'nant, Vysokomol. Soedin., 1960, vol. 2, p. 287. (Rus)
401. J. Bussink, J. Heijboer, and J.A. Prins (Ed.), " Physics of Non-Crystalline Solids", Proc. Intern.
Conf., Delft, 1964, p. 388, North Holland Publishing Co., Amsterdam, 1965.
402. B. von Falkai, and W. Rellensmann, Makromol. Chem., 1964, vol. 75, p. 112.
403. E.V. Gouinlock, E.J. Quinn, H.W. Marciniak, and R.R. Hindersinn, IUPAC Preprints, Intern.
Symp. Macromol. Chem., Toronto, Sect. A13.3, 1968.
404. H.G. Weyland, C.A.M. Hoefs, K. Yntema, and M.J. Mijs, Eur. Polym. J., 1970, vol. 6, p. 1339.
405. T. Takekoshi, ACS Polym. Preprints, 1969, vol. 10, p. 103.
406. G. Bier, Adv, Chem. Ser., 1969, vol. 91, p. 612.
407. A. Conix, IUPAC Preprints, Intern. Symp. Macromol. Chem., Montreal, Sect. D39, 1961; via O.G.
Lewis, "Physical Constants of Linear Homopolymers", Springer-Verlag, 1968.
408. C.S. Marvel, and C.H. Young, J. Amer. Chem. Soc., 1951, vol. 73, p. 1066.
409. W.A. Lee, and B. Stag, Royal Aircraft Establishment, Farnborough, Hants, UK, TR 71,223, 1971.
410. V.C.R. McLoughlin, J. Thrower, M.A.H. Hewins, J.S. Pippett, and M.A. White, Royal Aircraft
Establishment, Farnborough, Hants, UK, TR 70,160, 1970.
411. I.N. Duling, US Pat. 3,436,376, April 1, 1969.
412. H.J. Kolb, and E.F. Izard, J. Appl. Phys., 1949, vol. 20, p. 564.
413. Y. Wada, J. Phys. Soc. (Japan), 1961, vol. 16, p. 1226.
414. I.M. Ward, Textile Res. J., 1961, vol. 13, p. 650.
415. L. Mandelkern, G.M. Martin, and A. Quinn, J. Res. Nat. Bur. Std., 1957, vol. 58, p. 137.
416. G. Farrow, J. McIntosh, and I.M. Ward, Makromol. Chem., 1960, vol. 38, p. 147.
417. J.G. Smith, C.J. Kibler, and B.J. Sublett, J. Polym. Sci., 1966, vol. A-1(4), p. 1851; values from
R.M. Schulken, R.E. Boy, and R.H. Cox, J. Polym. Sci., 1964, vol. C6, p. 17.
418. I.M. Dolgopol'skii, Kh.A. Dobina, V.S. Fikhteingol'ts, S.A. Kamysheva, M.I. Sinaiskaya, L.G.
Balashova, and R.V. Zolotareva, Polymer Sci. USSR, 1967, vol. 9, p. 1723 (transl'n of
Vysokomol. Soedin, 1967, vol. A9, p. 1536 (Rus)).
419. E.V. Gouinlick, C.J. Verbanic, and G.C. Schweiker, J. Appl. Polym. Sci., 1959, vol. 1, p. 361.
420. W.R. Griffin, Rubber World, 1957, vol. 136, p. 687.
421. G.C. Schweiker, and P. Robitschek, J. Polym. Sci., 1957, vol. 24, p. 33.
422. J.H. Sewll, Royal Aircraft Establishment, Farnborough, Hants, UK, unpublished results.
423. I.J. Goldfarb, and R. McGuchan, AD 697987, 1987.
424. R.C. Evers, and G.F.L. Ehlers, J. Polym. Sci., 1969, vol. A-1(7), p. 3020.
425. A Conix, Ind. Eng. Chem., 1959, vol. 51, p. 147.
426. V. Frosini, and G. Vallebona, Chim. Ind., 1970, vol. 52, p. 499.
427. C.S. Marvel, and J.H. Johnson, J. Amer. Chem. Soc., 1950, vol. 72, p. 1674.
428. S.M. Somerville, and I.M. White, Royal Aircraft Establishment, Farnborough, Hants, UK, TR
66345, 1966.
429. I.M. White, M.A.H. Hewins, and S.M. Somerville, paper presented at 2nd Intern. Symp.
Organosilicon Chem., Univ. Bordeaux, July 9-12, 1968.
430. W.A. Lee, and I.H. Sewell, Royal Aircraft Establishment, Farnborough, Hants, UK, unpublished
results.
431. W.A. Lee, and F. Pinchin, Royal Aircraft Establishment, Farnborough, Hants, UK, unpublished
results.
432. French Pat. 1,175,362, 1959; Brit. Pat. 863, 704, 1961, via G.F.L. Ehlers, R.C. Evers, and K.R.
Fisch, J. Polym. Sci, 1969, vol. A-1(7), p. 3413.
433. W.M. Eareckson, J. Polym. Sci., 1959, vol. 40, p. 399.
434. V.V. Korshak, S.V. Vinogradova, C.L. Slonimskii, Ya.S. Vygodskii, S.N. Salazkin, A.A.
Askadskii, A.I. Mzhel'skii, and V.P. Sidorova, Polym. Sci. USSR, 1968, vol. 10, p. 2395 (transl'n
of Vysokomol. Soedin., 1968, vol. A10, p. 2058 (Rus)).
435. W.A. Lee, I.M. White, and F. Pinchin, Royal Aircraft Establishmant, Farnborough, Hants, UK,
unpublished results.
436. K.A. Gol'dgammer, G.G. Pimenov, A.I. Maklakov, V.V. Korshak, S.V. Vinogradova, P.M.
Valetskii, and A.N. Baskakov, Vysokomol. Soedin., 1968, vol. A10, p. 821 (Rus); (transl'n of
Polym. Sci. USSR, 1968, vol. 10, p. 953).
437. R.L. Combs, and R.G. Nations, J. Polym. Sci., 1970, vol. C30, p. 407 (unpublished results of T.F.
Gray).

680

438. V.V. Korshak, G.L. Berestneva, S.V. Vinogradova, A.N. Baskakov, and P.M. Valetskii, Polymer
Sci. USSR, 1968, vol. 10, p. 2300 (transl'n of Vysokomol. Soedin., 1968, vol. A10, p. 1984
(Rus)).
439. S.S. Karapetyan, A.Ya. Yakubovich, and I.L. Knunyants, Polym. Sci. USSR, 1964, vol. 6, p. 1718
(transl'n of Vysokomol. Soedin., 1964, vol. 6, p. 1550 (Rus)).
440. J.E. McIntyre, and E.C. Pugh, Brit. Pat. 838,986, 1960.
441. K. Saotome, and H. Komoto, J. Polym. Sci., 1967, vol. A-1(5), p. 119.
442. W.J. MacKnight, M. Yang, and T. Kajiyama, ACS Polym. Preprints, 1968, vol. 9, p. 860.
443. D.J. Lyman, J. Heller, and M. Borlow, Makromol. Chem., 1965, vol. 84, p. 64.
444. M.A. McCall, J.R. Caldwell, H.G. Morre, and H.M. Beard, J. Macromol. Sci., 1969, vol. A3, p.
911.
445. P.D. Schumann, E.C. Stump, and G. Westmoreland, J. Macromol. Sci., 1967, vol. B1, p. 815.
446. Yu.A. Yuzhelevskii, E.G. Kagan, A.L. Klebanskii, I.A. Zevakin, and A.V. Kharlamova,
Vysokomol. Soedin., 1969, vol. B11, p. 854. (Rus)
447. M.S. Paterson, J. Appl. Phys., 1964, vol. 35, p. 176.
448. K.A. Andrianov, T.K. Dzhashiashvili, V.V. Astakin, and G.N. Shumakavo, Vysokomol. Soedin.,
1968, vol. B10, p. 766. (Rus)
449. K.A. Andrianov, and S.E. Yakushkina, Vysokomol. Soedin., 1962, vol. 4, p. 1193. (Rus)
450. V.C.R. McLoughlin, and P.A. Grattan, Royal Aircraft Establishment, Farnborough, Hants, UK, TR
71224, 1971.
451. R.E. Burks, and J.L. Greene, N68-16509; Southern Res. Inst. Rept., 25, 1967.
452. J.A. Empen, and J.K. Stille, in: N.G. Gaylord (Ed.), "Macromolecular Synthesis", 3, Wiley, New
York, 1969, p. 15.
453. Polymer Handbook, Eds. J. Brandrup, E.H. Immergut, and E.A. Grulke, Associate Eds. A. Abe,
and D.R. Bloch, 4th Edition, 1999, Wiley, New York.
454. A. Nicco, J.P. Machon, H. Fremaux, J.Ph. Pied, B. Zindy, and M. Thiery, Eur. Polym. J., 1970,
vol. 6, p. 1427.
455. R.S. Nevin, and E.M. Pearce, J. Polym. Sci., 1965, vol. B3, p. 487.
456. W.J. Middleton, H.W. Jacobson, R.E. Putnam, H.C. Walter, D.G. Pye, and W.H. Sharkey, J.
Polym. Sci., 1965, vol. A3, p. 4115.
457. V.A. Engelhardt; report in Chem. Eng. News, 1965, vol. 43, p. 80.
458. W.H. Sharkey, W.J. Middleton, H.W. Jacobson, O.S. Acker, and H.C. Walter, Chem. Eng. News,
1963, vol. 41, p. 46.
459. J.A. Empen, and J.K. Stille, in: N.G. Gaylord (Ed.), "Macromolecular Synthesis", 3, Wiley, New
York, 1969, p. 53.
460. C.C. Price, and E.A. Blair, J. Polym. Sci., 1967, vol. A-1(5), p. 171.
461. R.M. Fitch, and D.C. Helgeson, J. Polym. Sci., 1969, vol. C22, p. 1101.
462. M. Morton, R.F. Kammereck, and L.J. Fetters, Brit. Polym. J., 1971, vol. 3, p. 120.
463. H. Komoto, Rev. Phys. Chem. (Japan), 1967, vol. 37, p. 112.
464. J. Pellon, J. Polym. Sci., 1963, vol. A1, p. 3561.
465. C.G. Overberger, and J.K. Weise, J. Amer. Chem. Soc., 1968, vol. 90, p. 3538.
466. H.G. Buhrer, and H.-G. Elias, Makromol. Chem., 1970, vol. 140, p. 41.
467. M.E.B. Jones, SPE Tech. Papers, 15, 27th ANTEC, Chicago, 1969, p. 453.
468. G. Jarrett, and M.E.B. Jones, Brit. Polym. J., 1970, vol. 2, p. 229.
469. H.Wexler, Makromol. Chem., 1968, vol. 115, p. 262.
470. R.G. Beaman, and F.B. Cramer, J. Polym. Sci., 1956, vol. 21, p. 223.
471. J.H. Magill, J. Polym. Sci., 1965, vol. A3, p. 1195.
472. B.Ke, and A.W. Sisko, J. Polym. Sci., 1961, vol. 50, p. 87.
473. Y.Nishijima, J. Selki, and T. Kawai, Rept. Progr. Polym. Phys. (Japan), 1967, vol. 10, p. 473.
474. V. Frosini, and E. Butta, J. Polym. Sci., 1971, vol. B9, p. 253.
475. D.C. Prevorsek, R.H. Butler, and K.H. Reimschuessel, J. Polym. Sci., 1971, vol. A-2(9), p. 867.
476. K. Schmieder, and K. Wolf, Kolloid Z., 1953, vol. 134, 149.
477. H. Komoto, Rev. Phys. Chem. (Japan), 1967, vol. 37, p. 105.
478. K. Saotome, and H. Komoto, J. Polym. Sci., 1966, vol. A-1(4), p. 1463.
479. H. Komoto, and K. Saotome, Chem. High Polymers (Japan), 1965, vol. 22, p. 337; Chem. Abstr.,
1965, vol. 56, p. 15045f.
480. G. Champetier, and J.P. Pied, Makromol. Chem., 1961, vol. 44/46, p. 64; Chem. Abstr., 1962, vol.
56, p. 1584c.
481. B.S. Marks, and G.C. Schweiker, J. Polym. Sci., 1960, vol. 43, p. 229.
482. D.A. Holmer, J. Polym. Sci, 1968, vol. A-1(6), p. 3177.

681

483. J.C. Mileo, B. Sillion, and G. De Gaudemaris, Compt. Rend., 1969, vol. C268, p. 2007.
484. J.B. Jackson, Polymer, 1969, vol. 10, p. 159.
485. L.C. Glover, and B.J. Lyons, ACS Polymer Preprints, 1968, vol. 9, p. 243.
486. G.A. Kuznetsov, V.D. Gerasimov, L.N. Fomenko, A.I. Maklakov, G.G. Pimenov, and L.B.
Sokolov, Polym. Sci. USSR, 1965, vol. 7, p. 1763 (transl'n of Vysokomol. Soedin., 1965, vol. 7,
1592 (Rus)).
487. G.S. Kolesnikov, O.Ya. Fedotova, V.V. Trezov, and V.N. Kuz'micheva, Polym. Sci. USSR, 1968,
vol. 10, p. 2612 (transl'n of Vysokomol. Soedin., 1968, vol. 10, p. 2248 (Rus)).
488. E. Bessler, and G. Bier, Makromol. Chem., 1969, vol. 122, p. 30.
489. N.S. Nametkin, N.V. Ushakov, and V.M. Vdovin, Polym. Sci. USSR, 1971, vol. 13, p. 31 (transl'n
of Vysokomol. Soedin., 1971, vol. A13, p. 29 (Rus)).
490. N.S. Nametkin, V.M. Vdovin, and V.I. Zav'yalov, Doklady Akad. Nauk SSSR, 1965, vol. 162, p.
824. (Rus)
491. F.S. Model, G. Redl, and E.-G. Rochow, J. Polym. Sci., 1966, vol. A-1(4), p. 639.
492. F.S. Model, and E.G. Rochow, J. Polym. Sci., 1970, vol. A-2(8), p. 999.
493. N.S. Nametkin, V.M. Vdovin, and V.I. Zav'yalov, Polym. Sci. USSR, 1965, vol. 7, p.836 (transl'n
of Vysokomol. Soedin., 1965, vol. 7, p. 757 (Rus)).
494. M. Morton, J. Elastoplastics, 1971, vol. 3, p. 112.
495. E. Jenckel, Kolloid Z., 1942, vol. 100, p. 163.
496. M.L. Dannis, J. Appl. Polym. Sci., 1959, vol. 1, p. 121.
497. R.H. Gerke, J. Polym. Sci., 1954, vol. 13, p. 295.
498. R.H.Gerke, J. Polym. Sci., 1954, vol. 13, p. 295.
499. W. Marconi, A. Mazzel, G. Ligni, and M. Bruzzone, J. Polym. Sci., 1967, vol. C16, p. 805.
500. D.I. Relyea, H.P. Smith, and A.N. Johnson, 1967, AD 657675.
501. W.R. Brown, and G.S. Park, Paint Technol., 1970, vol. 42, p.16.
502. F. Haas, K. Nutzel, S. Pampus, and D. Theisen, Rubber Chem Technol., 1970, vol. 43, p. 1116.
See also Kautschuk Gummi Kunststoffe, 1970, vol. 23, S. 502.
503. M. Takeda, K. Tanaka, and R. Nagao, J. Polym. Sci., 1962, vol. 57, p. 517.
504. I.A. Livshits, L.M. Korobova, and E.A. Sidorovich, Mekh. Polim., 1967, vol. 4, p. 596. (Rus)
505. C.G. Overberger, L.H. Arond, R.H. Wiley, and R.R. Garrett, J. Polym. Sci., 1951, vol. 7, p. 431.
506. W. Marconi, A. Mazzei, S. Cucinella, and M. Cesari, J. Polym. Sci., 1964, vol. B2, p. 4261.
507. I.A. Livshits, and L.M. Korobova, Dokl. Akad. Nauk SSSR, 1958, vol. 121, p. 474. (Rus)
508. G. Dall'asta, and P. Scaglione, Rubber Chem. Technol., 1969, vol. 42, p. 1235.
509. K.J. Clark, A.T. Jones, and D.H. Sandford, Chem. Ind., 1962, vol. 47, p. 2010.
510. G. Natta, F. Danusso, and G. Moraglio, J. Polym. Sci., 1957, vol. 25, p. 119.
511. K.R. Dunham, J. Vandenberghe, J.W.H. Faber, and L.E. Contois, J. Polym. Sci., 1963, vol. A1, p.
751.
512. G. Natta, Rend. Accad. Nazi. Lincei. Ser., 1958, vol. 8, p. 24; via Ref. 223.
513. S.F. Kurath, E. Passaglia, and R. Pariser, J. Appl. Phys., 1957, vol. 28, p. 499.
514. F.P. Reding, J.A. Faucher, and R.D. Whitman, J. Polym. Sci., 1962, vol. 57, p. 483.
515. H.-G. Elias, and O. Etter, J. Macromol. Sci., 1967, vol. A1, p. 943.
516. W.H. McCarty, and G. Parravano, J. Polym. Sci., 1965, vol. A3, p. 4029.
517. G.L. Taylor, and S. Davison, J. Polym. Sci., 1968, vol. B699.
518. S.T. Barsamyan, A.S. Apresyan, V.I. Kleiner, and L.L. Stotskaya, Soviet Plastics, 1968, vol. 3, p.
17 (transl'n of Plasticheskie Massy, 1968, vol. 3, p. 13 (Rus)).
519. R.E. Kelchner, and J.J. Aklonis, J. Polym. Sci., 1970, vol. A-2(8), p. 799.
520. A. Abe, and T. Hama, J. Polym. Sci., 1969, vol. B7, p. 427.
521. M. Takayanagi, Pure Appl. Chem., 1970, vol. 23, p. 151.
522. M. Baccaredda, and E. Butta, Chim. Ind. (Milan), 1962, vol. 44, p. 1228.
523. A.V. Tobolsky, "Properties and Structure of Polymers", 1960, Wiley, New York.
524. R.A. Jackson, S.R.D. Oldland, and A. Pajaczkowski, J. Appl. Polym. Sci., 1968, vol. 12, p. 1297.
525. J.P. Kennedy, W. Naegele, and J.J. Elliott, J. Polym. Sci., 1965, vol. B3, p. 729.
526. O.E. Van Lohuizen, and K.S. De Vries, J. Polym. Sci., 1968, vol. C16, p. 3943.
527. A.E. Woodward, Trans. N.Y. Acad. Sci., 1962, vol. 24, p. 250.
528. A.E. Woodward, J.A. Sauer, and R.A. Wall, J. Polym. Sci., 1961, vol. 50, p. 117.
529. J.H. Griffith, and B.G. Ranby, J. Polym. Sci., 1960, vol. 44, p. 369.
530. R.W. Penn, Trans. Soc. Rheol., 1963, vol. 3, p. 416.
531. W.A. Hewett, and F.E. Weir, J. Polym. Sci., 1963, vol. A1, p. 1239.
532. F.E. Karasz, H.E. Bair, and J.M. O'Reilly, Polymer, 1967, vol. 8, p. 547.
533. T.P. Melia, and A. Tyson, Makromol. Chem., 1967, vol. 109, p. 87.

682

534. I. Kirshenbaum, R.B. Isaacson, and M. Druin, J. Polym. Sci., 1965, vol. B3, p. 525.
535. N. Kawasaki, and M. Takayanagi, Rept. Progr. Polym. Phys. (Japan), 1967, vol. 10, p. 337.
536. R.G. Quynn, and B.S. Sprague, J. Polym. Sci., 1970, vol. A-2(8), p. 1971.
537. A.A. Miller, J. Polym. Sci., 1968, vol. A-2(6), p. 249.
538. G. Gianotti, G. Dall'asta, A. Valvassori, and V. Zamboni, Makromol. Chem., 1971, vol. 149, p.
117.
539. A.Turner, and F.E. Bailey, J. Polym. Sci., 1963, vol. B1, p. 601.
540. J.P. Kennedy, and R.M. Thomas, Macromol. Chem., 1963, vol. 64, p. 1.
541. J.P. Kennedy, and R.M. Thomas, Macromol. Chem., 1962, vol. 53, p. 28.
542. J.P. Kennedy, J.J. Elliott, and B. Groten, Macromol. Chem., 1964, vol. 77, p. 26.
543. G. Gianotti, and A. Capizzi, Eur. Polym. J., 1968, vol. 4, p. 677.
544. G.M. Pogosyan, G.A. Zhamkochyan, and S.G. Matsoyan, Arm. Khim. Zh., 1969, vol. 22, p. 364.
(Rus)
545. G.M. Pogosyan, L.M. Akopyan, E.V. Vanyan, and S.G. Matsoyan, Vysokomol. Soedin., 1971,
vol. B13, p. 242. (Rus)
546. W. Kern, W. Heitz, M. Jager, K. Pfitzner, and H.O. Wirth, Makromol. Chem., 1969, vol. 126, p.
73.
547. G.M. Pogosyan, L.M. Akopyan, E.V. Vanyan, and S.G. Matsoyan, Vysokomol. Soedin., 1970,
vol. B12, p. 142. (Rus)
548. K.R. Dunham, J.W.H. Faber, J. Vandenberghe, and W.F. Fowler, J. Appl. Polym. Sci., 1963, vol.
7, p. 897.
549. R. Kosfeld, Kolloid Z., 1960, vol. 172, p. 182.
550. K.H. Illers, Z. Elektrochem., 1961, vol. 65, p. 679.
551. G.M. Pogosyan, T.G. Karapetyan, and S.G. Matsoyan, Vysokomol. Soedin., 1971, vol. B13, p.
228. (Rus)
552. E.C. Chapin, J.G. Abrams, and V.L. Lyons, J. Org. Chem., 1962, vol. 27, p. 2595.
553. G.S. Kolesnikov, and G.M. Pogosyan, Izv. Akad. Nauk SSSR, Otd. Khim. Nauk, 1958, vol. 2, p.
227.
554. C.G. Overberger, C. Frazier, J. Mandelman, and H.F. Smith, J. Amer. Chem. Soc., 1953, vol. 75,
p. 3326.
555. T.E. Davies, British Plastics, 1959, vol. 32, p. 283.
556. G.M. Pogosyan, T.G. Karapetyan, and S.G. Matsoyan, Vysokomol. Soedin., 1970, vol. B12, p.
463. (Rus)
557. G.S. Kolesnikov, Izv. Akad. Nauk SSSR, Otd. Khim Nauk, 1959, p. 1333. (Rus)
558. K.R. Dumham, J. Vandenberghe, J.W.H. Faber, and W.F. Fowler, J. Appl. Polym. Sci., 1963, vol.
3, p. 1531.
559. Brit. Pat. 609,482, 1 Oct. 1948.
560. S. Krause, J.J. Gormley, N. Roman, J.A. Shetter, and W.H. Watanabe, J. Polym. Sci., 1965, vol.
A3, p. 3573.
561. J. Thrower, and M.A.H., Royal Aircraft Establishment, Farnborough, Hants, UK, TR 70056, 1970.
562. R.A. Abdrashitov, N.M. Bazhenov, M.V. Vol'kenstein, A.I. Kol'tsov, and A.S. Khachaturov,
Vysocomol. Soedin., 1963, vol. 5, p. 405; (transl'n in Polymer Sci. USSR, 1963, vol. 4, p. 1066);
Z. Phys. Chem., 1962, vol. 220, p. 413.
563. F. Danusso, and G. Polizzotti, Makromol. Chem., 1963, vol. 61, p. 157.
564. K. Tanaka, O. Yano, S. Manu, and R. Nagao, Rept. Progr. Polym. Phys. (Japan), 1969, vol. 12, p.
379.
565. R.A. Abdrashitov, N.M. Bazhenov, M.V. Vol'kensthtein, A.I. Kol'tsov, and A.S. Khachaturov,
Polym. Sci. SSSR, 1964, vol. 6, p. 2074; (transl'n of Vysocomol. Soedin., 1964, vol. 6, p. 1871).
566. US Pat. 2,651,627, 8 Sept. 1953.
567. G.M. Pogosyan, G.A. Zhamkochyan, R.A. Stepanyan, and S.G. Matsoyan, Arm. Khim. Zh., 1969,
vol. 22, p. 915. (Rus)
568. Polaroid Corp., Offic. Sci. Res. Dev. Rept. No. 4417, p. 246, Feb. 1945.
569. V. Frosini, and P.L. Magagnini, Eur. Polym. J., 1966, vol. 2, p. 129.
570. V. Frosini, and E. Butta, Mater. Sci. Eng., 1970, vol. 6, p. 274.
571. E. Rovira, and A. Eisenberg. Unpublished results quoted by: M.C. Shen, and A. Eisenberg in:
"Progress in Solid State Chemistry", Ed. by H. Reiss, 1966, vol. 3, Pergamon Press, New York, p.
407; reprint in Rubber Chem. Technol., 1970, vol. 43, p. 95.
572. G.T. Kennedy, and F. Morton, J. Chem. Soc., 1949, p. 2383.
573. L.C. Corrado, J. Chem. Phys., 1969, vol. 50, p. 2260.

683

574. W. Hodes, American Cyanamid Co. Unpublished results quoted by O.G. Lewis in "Physical
Constants of Linear Homopolymers", Springer-Verlag, 1968.
575. G.S. Kolesnikov, and G.M. Pogosyan, Izv. Akad. Nauk SSSR, Otd. Khim. Nauk, 1962, p. 2098;
Chem. Abstr., 1963, vol. 58, p. 8047a.
576. L.A. Utracki, and R Simha, Makromol. Chem., 1968, vol. 117, p. 94.
577. M. Bank, J. Leffingwell, and C, Thies, Macromolecules, 1971, vol. 4, p. 43.
578. M. Baccaredda, P.L. Magagnini, G. Pizzirani, and P. Giusti, J. Polym. Sci., 1971, vol. B9, p. 303.
579. E.C. Eberlin. Unpublished values quoted by Ref. 44.
580. E.C. Eberlin, unpublished data quoted by O.G. Lewis, "Physical Constants of Linear
Homopolymers", Springer-Verlag, 1968.
581. C.E. Rehberg, and C.H. Fisher, Ind. Eng. Chem., 1948, vol. 40, p. 1429.
582. C.E. Rehberg, W.A. Faucette, and C.H. Fisher, J. Am. Chem. Soc., 1944, vol. 44, p. 1723.
583. J.A. Shetter, J. Polym. Sci., 1963, vol. B1, p. 209.
584. J.A. Monahan, J. Polym. Sci., 1966, vol. A-1(4), p. 2381.
585. G. Pizzirani, and P.L. Magagnini, Chim. Ind. (Milan), 1968, vol. 50, p. 1218.
586. V.P. Shibayev, B.S. Petrukhin, Yu.A. Zubov, N.A. Plate, and V.A. Krgin, Polym. Sci. USSR,
1968, vol. 10, p. 258.
587. F.A. Boyer, and J.F. Abere, J. Polym. Sci., 1955, vol. 15, 537.
588. G.P. Mikhailov, and V.A. Shevelev, Polym. Sci. USSR, 1967, vol. 9, p. 2762 (transl'n of
Vysokomol. Soedin. 1967, vol. A9, p. 2442 (Rus)).
589. F.A. Bovey, J.F. Abere, G.B. Rathmann, and C.L. Sandberg, J. polym. Sci., 1955, vol. 15, p. 520.
590. A.S. Khachaturov, N.M. Bazhenov, M.V. Vol'kenstein, I.M. Dolgopol'skii, and A.I. Kol'tsov,
Soviet Rubber Technol., 1965, vol. 24(12), p. 9 (transl'n of Kauchuk i Rezina, 1965, vol. 24(12),
p. 6 (Rus)).
591. A.G. Pittman, B.A. Ludwig, and D.L. Sharp, J. Poym. Sci., 1968, vol. A-1(6), p. 1741.
592. W.R. Sorenson, and T.W. Campbell, "Preparative Methods of Polymer Chemistry", 2nd Ed., 1968,
p. 248.
593. E.P. Otocka, and T.K. Kwei, Macromolecules, 1968, vol. 1, p. 401.
594. J.R. Constanza, and J.A. Vona, J. Polym. Sci., 1966, vol. A-1(4), p. 2659.
595. F. Wuerstlin, in: H.A. Stuart (Ed.), "Die Physik der Hochpolymeren", Springer, Berlin, 1955,
Chap. 11.
596. R.A. Haldon, and R.Simha, J. Appl. Phys., 1968, vol. 39, p. 1890.
597. A.V. Tobolsky, and M.C. Shen, J. Phys. Chem., 1963, vol. 67, p. 1886.
598. S.S. Rogers, and L. Mandelkern, J. Phys. Chem., 1957, vol. 61, p. 985.
599. E.A.W. Hoff, D.W. Robinson, and A.H. Willbourn, J. Polym. Sci., 1955, vol. 18, p. 161.
600. J. Lal, and G.S. Trick, J. Polym. Sci., 1964, vol. A2, p. 4559.
601. G. Williams, and D.C. Watts, Trans. Faraday Soc., 1971, vol. 67, p. 2793.
602. W.M. Lee (advisers: F.R. Eirich and B.R. McGarvey), Dissertation, Polytechn. Inst., Brooklyn,
1967: via Diss. Abs., 1967, vol. B28, p. 1897.
603. E.N. Rostovskii, and L.D. Rubinovich, Vysokomol. Soedin., Karb. Vys. Soedin., Sb. Statei, 1963,
p. 140. (Rus)
604. M.C. Chen, J.D. String, and F.J. Matusik, J. Macromol. Sci., 1967, vol. B1, p. 15.
605. J. Janacek, and J.D. Ferry, J. Macromol. Sci., 1971, vol. B5, p. 219.
606. F. Lednicky, and J. Janacek, J. Macromol. Sci., 1971, vol. B5, p. 335.
607. H. Ochiai, H. Shindo, and H. Yamamura, J. Polym. Sci., 1971, vol. A-2(9), p. 431.
608. A.K. Schulz, J. Chim. Phys., 1956, vol. 53, p. 933.
609. N.N. Aylward, J. Polym. Sci., 1970, vol. A-1(8), p. 319.
610. T.I. Borisova, Vysokomol. Soedin., 1970, vol. A12, p. 932 (Rus) (transl'n in Polymer Sci. USSR,
1970, vol. 12, p 1060).
611. G.M. Chetyrkina, T.A. Sokolova, and M.M. Koton, Vysokomol. Soedin., Vsesoyuz. Khim.
Obshchestvo im. D.I. Mendeleeva, 1959, vol. 1, p. 248; Chem Abstr., 1959, vol. 53, p. 23059i.
612. B.E. Tate, Adv. Polym. Sci., 1967, vol. 5, p. 214 (Unpublished results of Charles Pfizer and Co.
Inc.).
613. J.B. Kinsingr, J.R. Panchak, R.L. Kelso, J.S. Bartlett, and R.K. Graham, J. Appl. Polym. Sci.,
1965, vol. 9, p. 429.
614. B.Wesslen, and R.W. Lenz, ACS Polym. Preprints, 1970, vol. 11, p. 105.
615. J.A. Powell, and R.K. Graham, J. Polym. Sci., 1965, vol. A-1(3), p. 3451.
616. J.A. Powell, and R.K. Graham, J. Polym. Sci., 1965, vol. A-1(3), p. 3451.
617. V.A. Kargin, and T.I. Sogolova, Zh. Fiz. Khim., 1949, vol. 23, p. 540. (Rus)
618. H.D. Anspon, and J.J. Baron, WADC Tech' 1 Report 57-24, Pt. 1, 1957, Pt. 11, 1958; via ref. 616.

684

619. H. Yuki, K. Hatada, T. Nijnomi, M. Hashimoto, and J. Ohshima, Polym. J. (Japan), 1971, vol. 2, p.
759.
620. R.A. Haldon, W.J. Schell, and R. Simha, J. Macromol. Sci., 1967, vol. B1, p. 759.
621. D.S. Otto, Vinyl Technol. Newsletter, 1968, vol. 5, p. 20.
622. W.J. Schell, R. Simha, and J.J. Aklonis, J. Macromol. Sci., 1969, vol. A3, p. 1297.
623. W.J. Schell, Thesis, Univ. S. California, 1969; via Diss. I Abstr. Intern., 1969, vol. B30, p. 2120.
624. L. Fishbein, and B.F. Crowe, Macromol. Chem., 1961, vol. 48, p. 221.
625. E.J. Vandenberg, R.F. Heck, and D.S. Breslow, J. Polym. Sci., 1959, vol. 41, p. 519.
626. C.E. Schildkecht, S.T. Gross, H.R. Davidson, J.M. Lambert, and A.O. Zoss, Ind. Eng. Chem.,
1948, vol. 40, p. 2104.
627. G.M. Pogosyan, E.S. Avanesyan, and S.G. Matsoyan, Arm. Khim. Zh., 1971, vol. 24, p. 694.
(Rus)
628. G. Pizzirani, P. Magagnini, and P. Giusti, J. Polym. Sci., 1971, vol. A-2(9), p. 1133.
629. J.D. Hoffman, J. Amer. Chem. Soc., 1952, vol. 74, p. 1696.
630. R.F. Boyer, J. Appl. Phys., 1954, vol. 25, p. 825.
631. F.P. Reding; private communication to Ref. 946.
632. J.D. Hoffman, and J.J. Weeks, J. Polym. Sci., 1958, vol. 28, p. 472; J. Res. Nat. Bur. Std., 1958,
vol. 60, p. 465.
633. T. Nakajima, and S. Saito, J. Polym. Sci., 1958, vol. 31, p. 423.
634. A. Nishioka, J. Polym. Sci., 1959, vol. 37, p. 163.
635. F. Krum, Kolloid Z., 1959, vol. 165, p. 77.
636. K.-H. Illers, and E. Jenckel, Kolloid Z., 1959, vol. 165, p. 84.
637. M. Baccaredda, and E. Butta, J. Polym. Sci., 1960, vol. 44, p. 421.
638. A.W. Meyers, V. Tammela, V. Stannett, and M. Szwarc, Mod. Plast., 1960, vol. 37, p. 139.
639. N.G. McGrum, J. Polym. Sci., 1962, vol. 60, p. 53.
640. A.H. Scott, D.J. Scheiber, A.J. Curtis, J.L. Lauritzen, and J.D. Hoffman, J. Res. Nat. Bur. Std.,
1962, vol. 66A, p. 269.
641. E. Passaglia, J.M. Crissman, and R.R. Stromberg, ACS Polym. Preprints, 1965, Vol. 6, p. 590.
642. J.D. Hoffman, G. Williams, and E. Passaglia, J. Polym. Sci., 1966, vol. C14, p. 173.
643. J.M. Crissman, and E. Passaglia, J. Polym. Sci., 1966, vol. C14, p. 237.
644. I.I. Perepechko, and L.A. Bodrova, Vysokomol. Soedin., 1968, vol. B10, p. 148. (Rus)
645. I.I. Perepechko, L.A. Kvacheva, L.A. Ushakov, A.Ya. Svetov, and V.A. Grechishkin, Soviet
Plastics, 1970, vol. 8, p. 39; (transl'n of Plasticheskie Massy, 1970, vol. 8, p. 43 (Rus)).
646. G.S. Kolesnikov, and M.G. Avetyan, Izv. Akad. Nauk SSSR, Otd. Khim. Nauk, 1959, p. 331.
(Rus)
647. D.W. Brown, and L.A. Wall, ACSPolym. Preprints, 1966, vol. 7, p. 1116.
648. E.M. Sullivan, E.W. Wise, and F.P. Reding, US Pat. 3,110,705, 12 Nov. 1962.
649. J.A.E. Kail, Polymer, 1965, vol. 6, p. 535.
650. H. Van Hoorn, Rheol. Acta, 1971, vol. 10. P. 208.
651. H. Hopff, and M.A. Osman, Angew. Makromol. Chem., 1969, vol. 6, p. 39.
652. M. Kinoshita, T. Irie, and M. Imoto, Makromol. Chem., 1967, vol. 110, p. 47.
653. P.L. Magagnini, and V. Frosini, Eur. Polym. J., 1966, vol. 2, p. 139.
654. A.R. Monahan, Macromolecules, 1968, vol. 1, p. 408.
655. M.A. Osman, and H.-G. Elias, J. Macromol. Sci., 1971, vol. A5, p. 805.
656. T.S. Reid, D.W. Codding, and F.A. Bovey, J. Polym. Sci., 1955, vol. 18, p. 417.
657. H.C. Haas, E.S. Emerson, and N.W. Schuler, J. Polym. Sci., 1956, vol. 22, p. 291.
658. G. Pizzirani, and P.L. Magagnini, J. Appl. Polym. Sci., 1967, vol. 11, p. 1173.
659. H. Hopff, and M.A. Osman, Makromol. Chem., 1970, vol. 135, p. 175.
660. H. Hopff, and M.A. Osman, Makromol. Chem., 1971, vol. 143, p. 289.
661. J.A. Price, M.R. Lytton, and B.G. Ranby, J. Polym. Sci., 1961, vol. 51, p. 541.
662. K.A. Movsisyan, G.T. Ovanesov, and K.A. Gasparyan, Uch. Zap. Erevan. Univ. Estestv. Nauk,
1980, vol. 2, p. 140. (Rus)
663. Y.K. Sung, D.E. Gregonis, G.A. Russell, and J.D. Andrade, Polymer, 1978, vol. 19(11), p. 1362.
664. L. Gargallo, and M. Russo, Makromol. Chem., 1975, vol. 176(9), p. 2735.
665. J.M.G. Cowie, Z. Haq, and I.J. McEwen, J. Polym. Sci., Polym. Lett. Ed., 1979, vol. 17(12), p.
771.
666. S.L. Malholtra, P. Lessard, and L.P. Blanchard, J. Macromol. Sci. A: Chem., 1981, vol. 15(1), p.
121.
667. S.L. Malholtra, P. Lessard, and L.P. Blanchard, J. Macromol. Sci. A: Chem., 1981, vol. 15(2), p.
279.

685

668. S.L. Malhotra, P. Lessard, and L.P. Blanchard, J. Macromol. Sci. A: Chem., 1981, vol. 15(8), pp.
1125, 1577.
669. J.A. Bergfjord, R.C. Penwell, and M. Stolka, J. Polym. Sci., Polym. Phys. Ed., 1979, vol. 17(4), p.
711.
670. Polymer Handbook, Eds. J. Brandrup, E.H. Immergut, and E.A. Grulke, Associate Eds. A. Abe,
and D.R. Bloch, 4th Edition, 1999, Wiley, New York.
671. U. Gaur, and B. Wunderlich, Macromolecules, 1980, vol. 13(6), p. 1618.
672. S.L. Malhotra, P. Lessard, and L.P. Blanchard, J. Macromol. Sci., Chem., 1981, vol. A15(2), p.
301.
673. W. Wrasidlo, Macromolecules, 1971, vol. 4(5), p. 642.
674. H. Aida, and H. Senda, Fukui Daigaku Kogakubu Kenkyu Hokoku, 1980, vol. 28(1), p. 95.
675. C.G. Seefried Jr., and J.V. Koleske, J. Polym. Sci., Polym. Phys. Ed., 1975, vol. 13(4), p. 851.
676. Same as ref. 1172.
677. E. Sorta, and A. De Chirico, Polymer, 1976, vol. 17(6), p. 348.
678. E.Sorta, and A. De Chirico, Polymer, 1976, vol. 17(4), p. 348.
679. The same as ref. 678.
680. K.H. Illers, Polymer, 1977, vol. 18(6), p. 551.
681. Same as ref. 680.
682. J.M. Barrales-Rienda, Ramos J. Gonzalez, and M. Sanchez. (Same as ref. 1196).
683. S.Wu, J. Appl. Polym. Sci., 1992, vol. 46, p. 619.
684. S.J. Bai, M. Dotrong, and R.C. Evers, J. Polym. Sci., Part B: Polym. Phys., 1992, vol. 30, pp.
1515.
685. M. Scandola, G. Ceccorulli, and M. Pizzoli, Macromolecules, 1992, vol. 25, p. 6441.
686. S. Koizumi, K. Tadano, Y. Tanaka, T. Shimidzu, S. Kutsumizu, and Yano, Macromolecules, 1992,
vol. 25, p, 6563.
687. T. Narasimhaswamy, S.C. Sumathi, and B.S.R. Reddy, J. Polym. Sci., Part A: Polym. Chem.,
1992, vol. 30, p. 2165.
688. S.S. Murthy, J. Polym. Sci., Part B: Polym. Phys., 1993, vol. 31, p. 475.
689. J.M.G. Cowie, and S. Miachon, Macromolecules, 1992, vol. 25, 3295.
690. D. Zaldivar, C. Peniche, A. Bulay, and J. San Roman, J. Polym. Sci., Part A: Polym. Chem., 1993,
vol. 31, p. 625.
691. E. Jabbari, and N.A. Peppas, Macromolecules, 1993, vol. 26, p. 2175.
692. E.M. Cavalcante, and M. Campos, Synth. Met., 1993, vol. 57, p. 4918.
693. V.I. Bondar, Yu.M. Kukharskii, Y.P. Yampol'skii, E.Sh. Finkelshtein, and K.L. Makovetskii, J.
Polym. Sci., Part B: Polym. Phys., 1993, vol. 31, p. 1273.
694. E. Espi, M. Alberdi, and J.J. Iruin, Macromolecules, 1993, vol. 26, p. 4586.
695. K.L. Ngai, and C.M. Roland, Macromolecules, 1993, vol. 26, p. 6824.
696. A.R. Khokhlov, and I.Ya. Erukhimovich, Macromolecules, 1993, vol. 26, p. 7195.
697. W.-Y. Chiang, and J.-Y. Lu, J. Appl. Polym. Sci., 1993, vol. 49, p. 893.
698. H. Kumazawa, J.-S. Wang, K. Naito, B. Messaoudi, and E. Sada, J. Appl. Polym. Sci., 1994, vol.
51, 1015.
699. S.-J. Chang, and H.-B. Tsai, J. Appl. Polym. Sci., 1994, vol. 51, p. 999.
700. M. Mourgeus-Martin, A. Bernes, and C. Lacabanne, J. Therm. Anal., 1993, vol. 40, p. 697.
701. D.M. Petrovic-Djakov, J.M. Filipovic, Lj.P. Vrhovac, and J.S. Velickovic, J. Therm. Anal., 1993,
vol. 40, p. 741.
702. C.L. Choy, K.W. Kwok, W.P. Leung, and F.P. Lau, J. polym. Sci., Part B: Polym Phys., 1994, vol.
32, p. 1389.
703. X. Zhang, M. Shomoda, and A. Toyoda, J. Polym. Sci., Part B: Polym. Phys., 1994, vol. 32, p.
1399.
704. G. Muller, and R. Stadler, Macromolecules, 1994, vol. 27, p. 1555.
705. C.K. Park, C.S. Ha, J.K. Lee, and W.J. Cho, J. Appl. Polym. Sci., 1994, vol. 53, p. 967.
706. C.C. Cypcar, P. Camelio, V. Lazzeri, L.J. Mathias, and B. Waegell, Macromolecules, 1996, vol.
29, p. 8954.
707. F. Koopmann, A. Burgath, R. Knischka, J. Leukel, and H. Fey, Acta Polym., 1996, vol. 47, p. 377.
708. A. Brunacci, J.M.G. Cowie, R. Ferguson, J.L.G. Ribelles, and A.V. Garayo, Macromolecules,
1996, vol. 29, p. 7976.
709. S. Matsuoka, J. Therm. Anal., 1996, vol. 46, p. 985.
710. D.D. Deppe, A. Dhinojwala, and J.M. Torkelson, Macromolecules, 1996, vol. 26, p. 3898.
711. F. Koopmann, and H. Frey, Macromolecules, 1996, vol. 29, p. 3701.
712. Y. Yuan, A. Siegmann, and M. Markis, J. Appl. Polym. Sci., 1996, vol. 60, p. 1475.

686

713. Y. Ding, and A.S. Hay, Macromolecules, 1996, vol. 26, p. 4811.
714. D. Li, and J. Brisson, Macromolecules, 1996, vol. 29, 868.
715. M. Chen, and H.W. Gibson, Macromolecules, 1996, vol. 29, p. 5502.
716. L.F. Brown, J.I. Scheinbeim, and B.A. Newman, Ferroelectrics, 1995, vol. 171, p. 321.
717. C.-P. Yang, and J.-H. Lin, J. Polym. Sci., Part A: Polym. Chem., 1996, vol. 29, p. 1943.
718. A. Spyros, and R.H. Marchessault, Macromolecules, 1996, vol. 29, p. 2479.
719. H. Takeno, S. Koizumi, H. Hasegawa, and T. Hashimoto, Macromolecules, 1996, vol. 29, p. 2440.
720. S.S. Wu, D.S. Kalika, R.R. Lamonte, and S. Makhija, J. Macromol. Sci., Part B: Phys., 1996, vol.
35, p. 157.
721. E. Perez, R. Benavente, A. Bello, J.M. Perena, and V.L. VanderHart, Macromolecules, 1995, vol.
28, p. 6211.
722. E. Perez, Z. Zhen, A. Bello, and R. Benavente, Polymer, 1994, vol. 35, p. 4794.
723. J.S. Kollodge, and R.S. Porter, Macromolecules, 1995, vol. 28, p. 4089.
724. H. Abe, I. Matsubara, Y. Doi, Y. Hoi, and A. Yamaguchi, Macromolecules, 1994, vol. 27, p. 6018.
725. K.H. Yoon, S.C. Lee, and O.O. Park, Polymer Journal, 1994, vol. 26, p. 816.
726. K. Osaki, T. Inoue, E.-J. Hwang, H. Okamoto, and O. Takiguchi, J. Non-Cryst. Solids, 1994, vol.
172/177, p. 838.
727. R.A. Huijts, and A.J. De Vries, "International Journal of Polymeric Materials", 1993, vol. 23, p.
231.
728. E.G. Kiparisova, T.A. Bykova, B.V. Lebedev, T.I. Guseva, K.A. Mager, and Yu.G. Gololobov,
Vysokomol. Soed., 1993, vol. A35(6), p. 615. (Rus)
729. T.A. Bykova, E.G. Kiparisova, B.V. Lebedev, T.I. Guseva, K.A. Mager, and Yu.G. Gololobov,
Vysokomol. Soed., 1991, vol. A33(12), p. 2602. (Rus)
730. T.A. Bykova, E.G. Kiparisova, B.V. Lebedev, K.A. Mager, and Yu.G. Gololobov, Vysokomol.
Soed., 1991, vol. A33(3), p. 614. (Rus)
731. T.A. Bykova, E.G. Kiparisova, B.V. Lebedev, N.G. Senchenya, K.A. Mager, and Yu.G.
Gololobov, 1993, vol. A35(1), p. 13. (Rus)
732. T.I. Guseva, Yu.P. Kvachev, N.G. Senchenya, and Yu.G. Gololobov, Russian Polymer News,
1997, vol. 2(3), p. 11.
733. N.G. Senchenya, M.D. Suchkova, K.A. Mager, T.I. Guseva, and Yu.G. Gololobov, Vysokomol.
Soed., 1990, vol. B32(7), p. 524. (Rus)
734. W.A. Lee, and R.A. Rutherford, In: Polymer Handbook, 2nd Ed., J. Brandrup and E.H. Immergut
(Eds.), Wiley, New York, p. III-139.
735. W.S. Bahary, D.I. Sapper, and J.H. Lane, Rubber Chem. Technol., 1967, vol. 40, p. 1529.
736. G.S. Trick, J. Appl. Polym. Sci., 1960, vol. 3, p. 253.
737. G.C. Eastmond, and J. Paprotny, J. Mater. Chem., 1996, vol. 6, p. 1459.
738. T. Takekoshi et al., J. Polym. Sci., Polym. Chem. Ed., 1980, vol. 18, p. 3069.
739. G.C. Eastmond, J. Paprotny, and I. Webster, Polymer, 1993, vol. 34, p. 2865.
740. V.S. Papkov et al., J. Polym. Sci., Polym. Chem. Ed., 1984, vol. 22, p. 3617.
741. C.L. Beatty, and F.E. Karasz, J. Polym. Sci., Polym. Phys. Ed., 1975, vol. 13, p. 971.
742. A. Matsumoto, K. Mizata, and T. Otsu, J. Polym. Sci., Polym. Chem. Ed., 1993, vol. 31, p. 2531.
743. K.R. Dunham, J. Vandenberghe, J.W.H. Faber, and W.F. Fowler, J. Appl. Polym. Sci., 1963, vol.
7, p.143.
744. A. Anton, and H.C. Beachel, J. Polym. Sci., 1969, vol. B7, p. 215.
745. B.F. Malichenko, E.V. Shedul'ko, and Yu.Yu. Kercha, Vysokomol. Soedin., 1967, vol. A9, p.
2482 (Rus) (transl'n in Polym. Sci. USSR, 1967, vol. 9, p. 2808).
746. Yu.Yu. Kercha, L.I. Ryabokon', and B.F. Malichenko, In: Synthesis and Physico-Chemistry of
Polyurethanes, Yu.S. Lipatov (Ed.), Naukova Dumka, Kiev, 1968, p. 198. (Rus)
747. J.B. Alexopoulos, J.A. Barrie, J.C. Tye, and M. Fredrickson, Polymer, 1968, vol. 9, p. 56.
748. K.E. Polmanteer, and M.J. Hunter, J. Appl. Polym. Sci., 1959, vol. 1, p. 3.
749. C.C. Price, and E.A. Blair, J. Polym. Sci., 1967, vol. A-1(5), p. 171.
750. T.G. Bassiri, A. Levy, and M. Litt, J. Polym. Sci., 1967, vol. B5, p. 871.
751. I.N. Markevich, S.I. Beilin, M.P. Teterina, G.P. Karpacheva, and B.A. Dolgoplosk, Dokl. Akad.
Nauk SSSR, 1970, vol. 191, p. 362. (Rus)
752. R.J. Schaffhauser, M.G. Shen, and A.V. Tobolsky, J. Appl. Polym. Sci., 1964, vol. 8, p. 2825.
753. M. Kaufman, and A.F. Williams, J. Appl. Chem., 1951, vol. 1, p. 489.
754. B. Vollmert, and H.Haas, Brit. Pat. 1,116, 933, June12, 1968.
755. J.H. Prager, R.M. McCurdy, and C.B. Rathmann, J. Polym. Sci., 1964, vol. A2, p. 1941.
756. R.H. Wiley, and G.M. Brauer, J. Polym. Sci., 1949, vol. 4, p. 351.

687

757. E.F. Jordan, G.R. Riser, B. Artymyshyn, W.E. Parker, J.W. Pensabene, and A.N. Wridley, J. Appl.
Polym. Sci., 1969, vol. 13, p. 1777.
758. L. Gargallo, and M. Russo, Makromol. Chem., 1975, vol. 176(9), p. 2735.
759. J. Ballesteros, G.J. Howard, and L. Teasdale, J. Macromol. Sci., Chem., 1977, vol. A11(1), p. 39.
760. A. Matsumoto, K. Mizuta, and T. Otsu, Macromolecules, 1993, vol. 26, p. 1659.
761. A.J. Milton, and A.P. Monkman, J. Phys. D: Appl. Phys., 1993, vol. 26, p. 1468.
762. G. Muller, and R. Stadler, Macromolecules, 1994, vol. 27, p. 1555.
763. P. Skoglund, and A. Fransson, J. Polym. Sci., Part B: Polym. Phys., 1994, vol. 32, p. 1999.
764. E. Perez, Z. Zhen, A. Bello, and R. Benavente, Polymer, 1994, vol. 35, p. 4794.
765. Yu.K. Govodskii, and Yu.S. Lipatov, Vysokomol. Soedin., 1968, vol. 10, p. 323. (Rus)
766. Polymer Handbook, J. Brandrup, E.H. Immergut, and E.A. Grulke Eds., A. Abe, and D.R. Bloch
Associate Eds., 1999, 4th edition, Wiley, New York.
767. Yu.I. Matveev, and A.A. Askadskii, Vysokomol. Soed., 1994, vol. B36(10), pp. 1750-1755. (Rus)
768. A.I. Bachinskii, Izbrannye Trudy, 1960, Moscow. (Rus)
769. V.V. Shaposhnikova, S.N. Salazkin, V.A. Sergeev, I.V. Blagodatskikh, L.V. Dobrovina, A.A.
Sakunts, and S.A. Pavlova, Izvestiya Akademii Nauk, Ser. Khim, 1996, No. 10, pp. 2526 2530.
(Rus)
770. V.V. Shaposhnikova, A.A. Askadskii, S.N. Salazkin, V.A. Sergeev, A.V. Samoryadov, A.P.
Krasnov, V.V. Kazantseva, and B.S. Lioznov, Vysokomol. Soedin., 1997, vol. A39(4), pp. 713
719. (Rus)
771. V.V. Shaposhnikova, S.N. Salazkin, K.I. Donetskii, G.V. Gorshkov, A.A. Askadskii, K.A.
Bychko, V.V. Kazantseva, A.V. Samoryadov, A.P. Krasnov, B.S. Lioznov, O.V. Afonicheva,
N.A. Svetlova, A.S. Kogan, and A.S. Tkachenko, Vysokomol. Soedin., 1999, vol. A41(2), pp.
217 225. (Rus)
772. M. Doi, J. Polym. Sci., Polym. Phys. Ed., 1983, vol. 21(5), p. 667; 1980, vol. 18(5), p. 1005.

689

Subject Index

A
Acetone 454
Acid,
-, acetic 454
-, butyric 454
-, isophthalic 70
-, isovaleric 454
-, polyacrylic 422
-, sebacic 70
-, n-valeric 454
Alcohol
polyvinyl 9, 39, 70
Aldehyde(s) 532
Amide(s) 533
Amine(s) 533
Amino group 632
Angle tangent
-, of dielectric losses 86
-, of mechanical losses 86
Anharmonic oscillator 405
Anhydride,
-, polysebacic acid 473
Aniline phthalein 70
Aniline phthalein polypyromellitimide 70
Anisotropy,
-, polarizability 426
Annihilation
-, positrons 52
Arabinogalactane 599
Arabinoglucuronoxylane 597
Arsenic 17
Artifact(s) 509
B
Barrier,
-, energetic 88
-, potential 88
Benzene 454, 516 518
Benzoyl peroxide 560
Birefrigrance 426
Bispheno-A-polycarbonate 70, 87, 449, 625
Bisphenol 358, 555
Bond(s)
hydrogen 63, 101, 399, 588
intrachain 588
intermolecular 101, 399
interchain 588
intraunit 588
physical 102
Van-der-Waals 101
Boron 17, 63
Bromine 17
Bromobenzene 454
Butanale 454
Butylethyl ether 454

n-Butanol 454
t-Butyl 626, 627
n-Butylacetate 454
t-Butylamine 644
Butyl acrylate 621
Butyl methacrylate 621
C
Carbon 17, 63, 399
Chlorine 17, 64
Chlorobenzene 454
Chloroform 516
Coefficient
bond elasticity 459
deformation-optical 426, 427
depolarization 446
diffusion 480
of heat capacity increase 562
of elasticity of rotary-isomeric subsystem 459
of linear expansion 59, 60
of molecular packing 29 34
of molecules packing in the surface
layer 530
of oscillator anharmonism 405
of packing, partial 398
of the reaction equation, stoichiometric 103
of the repeating unit friction 62
oscillator elasticity 405
partial 62
Poisson 427
segment friction 385
stress-optical 426, 427
volumetric expansion 59, 60, 63
weight 460
Conformation of macromolecules 88, 497
Connection cylinder(s) 524
Connective globules 524
Constant,
Boltzmann 88
equilibrium 103
Morse potential 457
reaction rate 477
universal gas 89
Constant(s) (increments),
-, fragments 567 568
-, linear chains 342
-, network cross-linked point 342
-, self (Van-der-Waals) volume 17
-, strong intermolecular interaction 63
-, weak intermolecular interaction 63
Contraction 47
Coordination bonds 10
Copolymer
alternating 12

690

block-copolymer 12
grafted 13
of n-butyl acrylate with tert-butyl
methacrylate 629
statistical
of styrene with acrylonitrile 626
of styrene with divinylbenzene 14
of styrene with hexafluorine-2propylstyrene 628
of styrene with 4-hydroxystyrene 628
of styrene with methyl methacrylate
629
of styrene with 4-vinylphenyl
dimethylsilanol 630
of styrene with vinylphenylhexafluoromethylcarbiol 625
Compatibilizer 632
Creep
-, mechanical 441
-, optical 441
of polymer 475
Criterion
miscibility 547
solubility 513, 514
Cross-linked point
-, cross-linked chains 14, 343
-, four-functional 345
-, isocyanurate 469
-, network 346
-, trifunctional 337, 345
Curve
dilatometric 58
recovery 56
stress relaxation 56, 496
thermogravimetric 78, 79
with saturation 555
Cyclohexane 454
Cyclohexanol 454
Cyclohexanon 454

D
Debye pattern 589
n-Decane 454
Defects in the form of
branchings 375
hanged chains 375
isolated loop 375
Degradation,
-, thermal 78
-, thermooxidative 78
Degree
-, cross-linking 363
-, polymerization 520
Dendrimer(s) 333
Density
-, real 45
-, seeming 45

of cohesive energy 504, 512


of polymer 18
Diamond 353
1,2-Dibromoethane 454
Diamyl ether 454
Dibutyl ether 454
Dielectric constant 418, 447
Dielectric(s) 447
Diethyl ether 454
Diethylamine 632
Diethylketone 454
Diisoamyl ether 454
Diisocyanate 467
Diisopropyl ether 454
Difference
of main normal stresses 427
of run, optical 427
Diffusion,
-, Ficks 479
-, kinetic units 475, 479
-, relaxants 475, 479
Dimer 560
1,3-Dioxolane 629
Dipropyl ether 454
Domain(s) 471
E
Elastomer networks 456
Energy
-, anharmonic oscillator, free 405
-, chemical bonds dissociation 409
-, cohesive 504, 538, 555
-, dipole-dipole interaction 62
-, dispersion interaction 595
-, free 61
-, heat 89
-, hydrogen bonds 61
-, interaction activation 494
-, intermolecular interaction 61
-, internal 88
-, low-temperature gamma-transition,
activation 595
-, particle 89
-, relaxants interaction 495
-, rotary isomers, difference 459
-, rotation 88
-, surface 469
-, transition 88
Entropy 89, 475
Epoxy resin 14
Equation
-, Arrhenius 103
-, Boltzmann 88
-, Boltzmann-Volterra 475
-, chemical kinetics 103
-, Clausius-Mossotti 418, 446
-, Fick 503
-, Flory 72
-, Fulcher-Tamman 75

691

-, Kargin-Slonymskii 74
-, Lorenz-Lorentz 418
Ethanale 454
Ethanol 454
Ether
-, butyethyl 454
-, diamyl 454
-, dibutyl 454
-, diethyl 454
-, diglycidyl 358
-, diisoamyl 454
-, diisopropyl 454
-, polyvinylethyl 69
-, polyvinylisobutyl 405
-, polyvinylmethyl 69, 628
Ethyl acetate 454
Ethyl cellulose 449
Ethylbenzene 454
Ethylbromide 454
Ethylformate 454
2-Ethylhexyl acrylate 621
Ethyliodide 454
Ethylpropionate 454
F
Fibrils 524
Fluorine 17, 61
Formamide 534
Force constant 409
Fragments
-, basic 602
-, intercrosslinked, mean 347
-, linear 343
-, network, repeating 345
Frequency
-, D-line of sodium 446
-, oscillation, natural 89
Front-factor 456
Function
-, memory 476
-, response 613
G
Galactouronane 598
Gibbs triangle 383
Glass transition
mechanical 85
structural 85
temperature interval 91
Globule 509
Glucomannan 599
H
Halogen-containing compound(s) 530
Heat capacity 562
Height of the sphere segment 17
Hemicellulose 589

n-Heptane 454
n-Heptylacrylate 621
Hexagonal packing 528
Hexamethylene tetramine 629
n-Hexane 454
n-Hexanol 454
Hydrocarbons
-, monosubstituted, unsaturated 11
-, substituted, saturated 11
Hydrogen 17, 63
I
Interaction cylinder 457
Interaction,
-, chemical 16
-, dipole dipole 61, 101
-, dispersion self 61
-, dispersion weak 61
-, intermolecular (physical) 16, 607
-, intermolecular strong 61
-, intramolecular 607
Iodine 17
Isobutanol 454
cis-Isomer 11
Isomer(s) 11, 320
Isopentane 453
Isopropylbenzene 454
J
Joining of units
-, anomalous (head-to-head, tailto-tail) 10
normal (head-to-tail) 10
K
Ketone(s) 533
Kinetic unit 87, 475
L
Law (rule),
-, additivity 405
-, Ilushin 494
-, Laplace 511, 524
-, mass action 103
-, Maxwell 447
-, molar parts, Zhurkovs 105
-, Wertheim 426
Length,
-, chemical bond 17
-, Cune segment 385
-, intermolecular radius 17
-, wave 418
Lifetime,
-, intermediate component 52
of atom, settled 89
of long living component 51, 52

692

-, positron 51
Lifetime spectrum of positron(s) 51
Long-living component intensity 52
Lorentz correction 418
M
Macrodiisocyanates 468
Macromolecule
branched 13
flexible 73
linear 13
rigid 73
Main
principal stress(es) 427
Mechanism
glass transition 87
of relaxation process 87
of sorption, relaxation 497
relaxation 54, 55
swelling, relaxation 497
Melting
enthalpy 398
entropy 398
point 398
Memory function 477
Metha-, para-phenylene diamine 109
Methanol 454
Method
absorption, equilibrium 43
calorimetric 83, 85
capillary condensation 43
dilatometric 45, 83, 85
DSC 626
electron microscopy 469
electron spectroscopy 632
extrapolation 398
Foukes 535
Fourier-transformation 630
IR-spectroscopy 626, 630
neutron scattering 509
NMR, 13C NMR, 31P NMR 630, 631
of atomic approaches 76
of atomic constants 584
of deformations freezing 439
of dynamic mechanical analysis 86,
469
of experiment design 612
of gradient tubes 45
of dynamic tests 83
of least squares 108
of local Lorentz field 417, 418
of mercury porometry 43
of miscibility diagrams 587
of polymerization in solution 45
of registration of positron lifetime
spectra 51
of small-angle X-ray scattering 509
of X-ray photoelectronic spectroscopy
469

optical 83
photocreep 443
photoelasticity 439
polarization-optical 439
polycondensation 45
positrons annihilation 43
pycnometric 45
reverse variable step 496
Runge-Cutt, numerical 482
Simpson 496
thermogravimetric 79
thermomechanical 66, 78
Methyl acetate 454
Methyl formate 454
Methyl methacrylate 628
Methylethylketone 454
4-O-Methylglucuronoxylane
Methyliodide 454
Microdomain(s) 471
Microphase 471, 547
Miscibility of polymers 547, 625
Mobility
translation 102
vibration 102
Modulus
elasticity 79, 86, 440, 460, 474
equilibrium 456
of cross-linked points 460
of linear fragments 460
relaxation 494
rubbery 79
storage 473
Molecular mass
of chain 385
of cross-linking bridges 350
of cross-linking point 347, 350
of intercross-linking point fragment
456
of linear chain between cross-linking
points 348
of polymer 555
of segment 71
of the repeating unit 35
of unit neighboring to the crosslinking point 347
Moment
dipole 595
magnet 445
Monomers
bifunctional 14
disubstituted 11
polyfunctional 14
N
N,N-dimethylacrylamide 630
N,N-dimethyl-substituted Nylon-6,12 626
Naphthalene 454
Network(s)
high cross-linked 343

693

less cross-linked 343


Nitrile
acetate 455
butyrate 455
isoheptanate 455
pentanate 455
propionate 455
Nitrobenzene 518 520
Nitrogen 17, 63, 505, 563
Nitroethane 454
Nitromethane 454
1-Nitropropane 454
2-Nitropropane 454
n-Nonane 454
n-Nonyl acrylate 454
Number
-, coordination 64
-, repeating units in a segment 385
Nylon-6 626
O
n-Octane 454
n-Octanol 454
Oscillator 446
Osmotic pressure 497
Oxygen 17, 63, 399

P
Parachor
of atom 527
of frequency 527
n-Pentane 453
n-Pentanol 454
Perfluorocompound 532
Permeability 595
Phase transition
-, the first kind 87
-, the secondary kind 87
Phenolic resins 626
Phenoloformaldehyde resin 10, 607, 631
Phenolphthalein 70
Physical state
-, glassy 29, 41, 67
-, rubbery 41, 67
-, viscous flow 67
Polarizability 446, 595
Polarization,
atomic 418
deformational 447
electronic 418, 446
microscopic 446
orientational 418, 447
1,2-Polybutadiene 37
1,4-Polybutadiene 68, 449
1,4-Polyisocyanurate(s) 439
1,4-Polyisoprene 68, 449

1,4-cis-Polybutadiene 37
1,4-cis-Polyisoprene 37
1,4-trans-Polybutadiene 37
Poly(,,,-tetrafluorine-p-xylylene) 449
Poly(-methyl styrene) 449
Poly(-vinyl naphthalene) 449
Poly(-vinyl naphthalene) 449
Poly(1,1,2-trichlorobutadiene) 632
Poly[1,1-cyclohexane bis(4-phenyl)carbonate]
449
Poly[1,1-ethane bis(4-phenyl)carbonate] 449
Poly(1,4-cyclohexylidene dimethylene terephthalate) 449
Poly(1-butene) 449
Poly(1-hydroxy-2,6-methylphenylene) 629
Poly(1-hydroxyl-2,6-methylenephenylene)
629
Poly[2,2-(m-phenylene)-5,5bibenzimidazole] 450
Poly(2,6-dimethyl-1,4-phenylene oxide) 449,
625
Poly(3,4-dichlorostyrene) 449
Poly[4,4-diphenoxy di(4-phenylene)sulfone]
450
Poly[4,4-isopropylidene diphenoxy di(4phenylene)sulfone] 450
Poly[4,4-sulfone diphenoxy di(4-phenylene)
sulfone] 450
Poly(4-hydroxy styrene) 629
Poly(4-methyl-1-pentene) 449
Poly-4-methylpentene-1 38, 404, 424
Poly(4-vinyl pyridine) 631, 642
Polyacetylene 9
Polyacrylamide 72
Polyacrylonitrile 72, 450
Polyamide 9, 38, 75, 109, 399, 405, 537
Polyamide 6,6--isomer 38, 537
Polyamide 6,6--isomer 38
Polyamide based on sebacic acid and anilinephthalein 70
Polyamidophenylquinoxaline(s) 320
Polyarylate F-1 74
Polybenzimidazole(s) 582
Polybenzoxazole 484
Polybutadiene 449
Poly(butyl acrylate) 69
Poly(butyl methacrylate) 625, 449
Polycarbonate 582
Polycarbonate based on bisphenol A 70, 87,
449, 625
Polychloroprene 37, 41
Poly(chloro-p-xylylene) 449
Poly(chlorotrifluoroethylene) 449
Poly(cyclohexyl methacrylate) 449
Poly(dimethyl siloxane) 68, 449
Polyester(s) based on
secacic acid and phenolphthalein 70
isophthalic acid and phenolphthalein
70
Polyether(s) 9, 468

694

Poly(ether ether ketone) 449


Poly(ether ketone(s)) 449
Poly(ether sulfone(s)) 449
Poly(ethyl -chloroacrylate) 449
Poly(ethyl methacrylate) 69, 449
Poly(ethyl oxazoline) 627
Polyethylene 9, 449
Poly(ethylene adipate) 38
Poly(ethylene isophthalate) 38
Poly(ethylene oxazoline) 628
Poly(ethylene oxide) 9, 537, 625
Poly(ethylene terephthalate) 70, 450
Polyformaldehyde 537
Poly(hexamethylene adipamide) 450
Poly(hexamethylene sebacamide) 449
Poly(hexamethylene terephthalate) 37
Poly(hexyl methacrylate) 69
Polyimide 69
Poly(isobutyl methacrylate) 449
Polyisobutylene 69, 449
Polyisoprene 449
Polymer chains
-, branched 12, 13
-, cross-linked 14
-, linear 13
-, main 13
Polymer structure
-, fibrillar 524
-, globular 523
-, macroporous 44
-, microporous 43
-, nodular 509
-, porous 43
-, supermolecular 509
Polymeric unit,
-, elementary 10
-, repeating 10
Polymers
amorphous 34, 46, 84
atactic, irregular 11
branched 13
carbochain 9, 10
cardo-type structure 514
chelate 10
crystalline 46, 84
dendrite 13, 335
elementorganic 9, 10
erythro-diisotactic 12
glassy-like 68
globular, amorphous 510, 520
gradient 465, 472
heat-resistant 68, 514
heterochain 10
heteromodular 465
homo- 35
inter- 15
isotactic 11
isotropic, amorphous 520
ladder 15
network 10, 608

non-globular, amorphous 510


non-polar 538
optically sensitive 426
oriented 524
partially crystalline 34, 84
polar 538
rigid-chain 73
rubber-like 68
semi-crystalline 45, 46, 84
star-like 335
stereoregular 11
syndiotactic 11
treo-diisotactic 12
Poly(m-chlorostyrene) 449
Poly(methyl -chloroacrylate) 450
Poly(methyl acrylate) 69
Poly(methyl methacrylate) 69, 449, 626, 629,
630
Polymethylidene phthalide 46
Poly(N,N-dimethyl acrylamide) 627
Poly[N,N-(n,noxydiphenylene)pyromellitimide] 450
Poly(naphtoylene benzimidazole) 68
Poly(n-butyl acrylate) 68, 629
Poly(n-butyl methacrylate) 36, 626, 640
Poly(n-butyraldehyde) 41
Poly(n-propyl methacrylate) 36
Poly(N-vinyl carbazole) 449
Poly(o-methyl styrene) 449
Poly(oxy-2,2-dichloromethyl trimethylene)449
Poly[oxy(2,6-diphenyl-1,4-phenylene)] 449
Polyolefins 582
Polyoxadiazole 9, 484
Polyoxymethylene 39
Poly(p-chlorostyrene) 449
Poly(phenyl quinoxaline) 514
Poly(phenyl sylsesquioxane) 15
Polyphenylene 9
Poly(phosphonitrile chloride) 9
Poly(p-hydroxybenzoate) 450
Poly(p-hydroxystyrene) 628
Polypropylene 9, 37, 449
Poly(propylene oxide) 40
Poly(p-xylylene) 449
Polystyrene 69, 449
Polysulfide(s) 582
Polysulfone(s) 582
Poly(t-butyl methacrylate) 629
Poly(tert-butyl acrylate) 629
Poly(tetramethylene adipate) 405
Poly(tetramethylene oxide) 405
Poly(tetramethylene terephthalate) 449
Polytetrafluoroethylene 39, 50
Poly[thio(p-phenylene)] 449
Poly(trichlorobutadiene) 15
Poly(trimethylene adipate) 405
Polyurethane(s) 9, 582
Poly(vinyl acetate) 69, 450
Poly(vinyl alcohol) 70
Poly(vinyl butyral) 449

695

Poly(vinyl chloride) (PVC) 39, 69, 449, 537


Poly(vinyl cyclohexane) 449
Poly(vinyl fluoride) 39, 537
Polyvinylmethyl ether 69
Polyvinylethyl ether 69
Poly(vinyl pyrrolidone) 426, 628
Poly(vinyliden chloride) 449
Pore(s)
distribution by sizes 43
specific surface 43
total volume 43
Porosity factor 45
Potential
chemical interaction 409
Lennard-Jones 406
Morse 409
Principal strains 427
Process
of adhesive wetting 511
of copolymerization 12
of creep 475
of desorption 500
of microphase separation 470
of polymerization 46
of stress relaxation 475
of swelling 497
of syneresis 403
Propanal 454
n-Propanol 454
n-Propyl acetate 454
Pyromellitic dianhydride 70
R
Rate
of diffusing particle 497
of system cooling 91
of system heating 91

-, of the method of constants 585


Segment
-, molecular mass 72, 73
Cune 385
mechanical 73
of macromolecule 51
Silicon 17, 63
Simplex lattice 613
Size
of bond, typical 459
of crystallites 398
of solvent, typical 513
typical 385
Solubility
criterion 513
-, Hildebrand parameter 504
Sorbate 44, 45
Sorbent 44, 45
Spherical segment height 17
Square (surface)
of interaction cylinder 457
Van-der-Waals, of atom 459
Steric factor 103, 588
Structural element 89
Styrene 454
Sub-micropore(s) 50
Substituent, side 73
Substitution,
-, metha- 63
-, ortho- 63
-, para- 63
Sulfonated
polyamide 631
poly(p-phenylene
terephthalamide)
631
Sulfur 17, 63
T

Reaction
of polycyclotrimerization 15, 468
of trimerization 15
of urethane formation 468
Refraction
molar 448
molecular 418
of atoms, by Eizenlor 419
of atoms, specific 418
Relaxant(s) 475
Relaxation
of microdefects 55
stress 475
Rubber,
-, butadiene 514
-, isoprene 514
-, synthetic 10
S
Scheme
-, additive 527

Teflon 51
Temperature (point)
-, critical 405
-, dissolution, lowest critical 627
-, flow 67
-, glass transition 36, 66, 589
-, intensive thermal degradation 408
-, melting 398
-,onset of intense thermal degradation
408
-, phase transition 406
Tension
interfacial 510
surface 510
Theory (concept, rule)
-, Flory- Haggins 520
-, Iliushin 494
-, Patterson and Robard 627
-, rotary-isomeric 521
-, Simha-Boyer 60
-, Volkenstein-Ptytsin 90

696

-, Zhurkov 101

Vector,

Unsaturated hydrocarbons
disubstituted 11
monosubstituted 11

-, electric field intensity 445


-, electric induction 445
-, polarization 445
Viscosity,
-, shear, Newtonian 385
-, system 75
Volume
empty 17
fluctuation 42
free 42
molar 17
of atom 17
of defect 457
of end groups 555
of expansion 42
of ideal crystal 44
of interaction cylinder 457
of the repeating unit 17
self (real) 17
specific 17
total molar 17
Van-der-Waals, free 17
Van-der-Waals, of polymer 557
Van-der-Waals, of the repeating
fragment of the network 344

Value,

Wisevesser formula 585


Work of adhesion 510

Tests,
-, thermogravimetric 78
-, thermomechanical 78
-, X-ray structural 79
with periodical loading 84
Thermodynamic
function 87
probability 88
Time,
-, observation 89
-, relaxation 51, 89
of transition from the glassy state into
the rubbery state 88, 89
Tin 17
Toluene 454
2,4-Toluylenediisocyanate 468, 472, 473
Trans-isomer 11
Triethylamine 632
Trimer 568
U

-, defect 457
-, mechanical segment 73
Van-der-Waals,
-, radius of atom 17, 457 459
-, surface of atom 459
-, volume of atom 17 29, 453
455, 457 459

X
m-Xylene 454
o-Xylene 454
p-Xylene

454

You might also like