You are on page 1of 512

The Physics of Glassy Polymers

JOIN US ON THE INTERNET VIA WWW, GOPHER, FTP OR EMAIL:


WWW: http://www.thomson.com
GOPHER: gopher.thomson.com A service of I®P~
FTP: ftp.thomson.com
EMAil: findit@kiosk.thomson.com
The Physics of
Glassy Polymers
Second edition

Edited by

R.N. Haward
Visiting Professor
Department of Chemistry
UMlsr
UK

and

R.J. Young
Royal Society Wolfson Research Professor of Materials Science
Manchester Materials Science Centre
University of Manchester/UMISr
UK

lalll SPRlNGER-SCIENCB+BUSINESS MEDIA, B.V.


First edition 1973

Second edition 1997

© 1997 Springer Science+Business Media Dorru:echt


Originally published by Olapman & Hall in 1997
Softcover reprint of the hardcover 2nd edition 1997
Typeset in 10/12pt Times in The Republic of Ireland by Doyle Graphics

ISBN 978-94-010-6472-9 ISBN 978-94-011-5850-3 (eBook)


DOI 10.1007/978-94-011-5850-3
Apart from any fair dealing for the purposes of research or private study,
or criticism or review, as permitted under the UK Copyright Designs and
Patents Act, 1988, this publication may not be reproduced, stored,
or transmitted, in any form or by any means, without the prior permission
in writing of the publishers, or in the case of reprographic reproduction
only in accordance with the terms of the licences issued by the
Copyright Licensing Agency in the UK, or in accordance with the terms
of licences issued by the appropriate Reproduction Rights Organization
outside the UK. Enquiries conceming reproduction outside the terms
stated here should be sent to the publishers at the London address
printed on this page.
The publisher makes no representation, express or implied, with regard
to the accuracy of the information contained in this book and cannot
accept any legal responsibility or liability for any errors or omissions that
may be made.
A catalogue record for this book is available from the British Library
Library of Congress Catalog Card Number: 97-67483

@) Printed on permanent acid-free text paper, manufactured in accordance with


ANSI/NISO Z39.48-1992 and ANSI/NISO Z39.48-1984 (Permanence of Paper).
Contents

List of contributors ix
Preface xi

1 Introduction 1
R.N. Haward and R.J. Young
1.1 Introduction 1
1.2 The glassy state 1
1.3 Stiffness and strength 9
1.4 Entanglements 17
References 29

2 Molecular dynamics modelling of amorphous polymers 33


J.H.R. Clarke
2.1 Introduction 33
2.2 Ingredients of a computer simulation 35
2.3 Preparation of model polymer melt samples 42
2.4 Characterization of chain dynamics in dense polymers 48
2.5 Studies of the glass transformation 52
2.6 Stress-strain properties 60
2.7 Penetrant diffusion 73
2.8 Conclusions and forward look 80
References 82
3 Relaxation processes and physical aging 85
J.M. Hutchinson
3.1 Introduction 85
3.2 Structural relaxation in the glass transition region 89
3.3 Secondary relaxations 128
3.4 Physical aging and mechanical properties 138
References 146
4 Yield processes in glassy polymers 155
B. Crist
4.1 Introduction 155
4.2 Mechanical testing and definitions 159
vi Contents

4.3 Yield phenomena in glassy polymers 168


4.4 Related studies of yielding 176
4.5 The nature of yielding in glassy polymers 185
4.6 Constitutive analyses 189
4.7 Molecular models 196
4.8 Molecular simulations 202
4.9 Conclusions 208
References 210
5 The post-yield deformation of glassy polymers 213
M.e. Boyce and R.N. Haward
5.1 General features of post-yield deformation 213
5.2 Developments in the measurement of true stresses
and strains 223
5.3 Physical aging and large deformations 232
5.4 Thermal effects during the deformation of glassy
polymers 242
5.5 Models for large strains 254
5.6 Application of the Gaussian theory to the study of
large deformations in thermoplastics 260
5.7 Three-dimensional modelling of large strains 268
5.8 Three-dimensional kinematics of deformation 278
5.9 Numerical simulation of inhomogeneous deformation 280
References 289
6 Crazing 295
A.M. Donald
6.1 Introduction 295
6.2 Craze morphology 297
6.3 Initiation and growth 301
6.4 Craze micromechanics 307
6.5 Molecular mechanisms 310
6.6 Effect of external parameters 321
6.7 Crazing in the presence of small molecules 330
6.8 Crosslinking 331
6.9 Craze failure 334
6.10 Conclusions 337
References 339
7 Fracture mechanics 343
].G. Williams
7.1 Introduction 343
7.2 Elastic fracture mechanics 344
7.3 Standard for linear elastic fracture tests 347
7.4 'J' testing 351
Contents VII

7.5 Essential work tests 354


7.6 Examples of fracture data 356
7.7 Conclusions 361
References 361
8 Rubber toughening 363
c.B. Bucknall
8.1 Introduction 363
8.2 Characterization 365
8.3 Toughening mechanisms - principles 369
8.4 Cavitation diagrams 387
8.5 Factors affecting deformation of toughened plastics 391
8.6 Overview 408
References 409
9 Interfaces 413
R.A.L./ones
9.1 Introduction 413
9.2 Interfaces between incompatible polymers 416
9.3 Reinforcement of polymer-polymer interfaces with
block copolymers 422
9.4 Grafted chains at polymer-solid interfaces 434
9.5 Chain conformation and dynamics in glassy polymers
near interfaces 439
References 448
10 Morphology of block copolymers 451
A.I. Ryan and 1. W. Hamley
10.1 Introduction 451
10.2 Microphase separation theory 453
10.3 Techniques used to study morphology 456
10.4 Morphology 458
10.5 Summary and conclusions 494
References 494
Index 499
Contributors

M.e. Boyce A.M. Donald


Department of Mechanical Department of Physics,
Engineering, Cavendish Laboratory,
Massachusetts Institute University of Cambridge,
of Technology Madingley Road,
Cambridge, Cambridge CB3 OHE,
Massachusetts 02139-4507, UK
USA
I.w. Hamley
School of Chemistry,
C.B. Bucknall University of Leeds,
Advanced Materials Leeds LS2 9JT,
Department UK
Cranfield University
R.N. Haward
Cranfield,
1A Gaddum Road"
Bedford,
Bowdon,
MK430AL,
Cheshire WA14 3PD,
UK
UK
J.M. Hutchinson
J.H.R. Clarke Department of Engineering,
Department of Chemistry, Fraser Noble Building,
UMIST, University of Aberdeen,
Manchester M60 1QD, King's College,
UK Aberdeen AB24 3UE,
UK

B. Crist R.A.L. Jones


Department of Materials Science Department of Physics,
and Engineering, Cavendish Laboratory,
Northwestern University, University of Cambridge,
Evanston, Madingley Road,
Illinois 60208-3108, Cambridge CB3 OHE,
USA UK
x Contributors

A.J. Ryan R.I. Young


Materials Science Centre, Materials Science Centre,
University of Manchester /UMIST, University of Manchester/UMIST,
Grosvenor Street, Grosvenor Street,
Manchester Ml 7HS, Manchester Ml 7HS
UK UK
J.G. Williams
Department of Mechanical
Engineering,
Imperial College,
Exhibition Road,
London SW7 2BX,
UK
Preface

Since the original book entitled The Physics of Glassy Polymers was
published in 1973 there have been very substantial developments in both the
theory and application of polymer physics, and new materials have been
introduced. Further, in this large and growing field of knowledge, glassy
polymers are of particular interest as they have a homogeneous structure
that is fundamentally simpler than that of crystalline or reinforced materials.
The understanding of their properties, in many ways, forms the basis of our
knowledge of structure/property relationships in all other polymeric ma-
terials.
As might be expected, progress since 1973 has been uneven. In some cases
it has been slow, as in the field of thermodynamics, where the earlier
contribution of Rehage and Borchard may still be read with advantage. In
other areas, such as molecular dynamics, crazing and interfacial properties,
completely new ideas have been introduced that have profoundly enhanced
our knowledge of structure/property relationships. Indeed, if there is one
unifying theme that characterizes the present book, it is the emergence of
the polymer molecule with its multiplicity of structures and conformations
as the major factor controlling the properties of glassy polymers.
As before, the editors have offered an introduction which is intended to
cover some of the relevant established science in the field and to summarize
concepts generally assumed in the later chapters. We would also like to
thank the distinguished contributors of specialized chapters and express the
hope that our work together will playa significant part in the development
of the subject in the future.

R.N. Haward and R.J. Young


Manchester
Introd uction 1
R.N. Haward and R.J. Young

1.1 INTRODUCTION
The aim of this introductory chapter is twofold. The first is to give some
background on the glassy state and the general mechanical properties of
glassy polymers, with which the reader may not be familiar and which are
not covered in the subsequent chapters. The concept of the glassy state is
an underlying feature of the whole of the book. It is necessary to review the
present level of understanding of the structure of glassy polymers and how
the structure controls the physical properties, as this is an important feature
covered in several of the subsequent chapters. Readers who need to obtain
more information upon the synthesis, structure and properties of polymers
in general (including glassy polymers) are directed to~ards more basic
textbooks that cover this in detail (e.g. Young and Lovell, 1991).
The second aim concerns a review of the concept of entanglements and
their influence upon the physical properties of glassy polymers. The recog-
nition of the importance of entanglements in polymer science is a major
development since the original book was published in 1973, and this is a
recurring feature of the later chapters. Because of this, a simplified account
is given of the nature of entanglements and how their density is estimated.
An indication is then given of how they affect the properties of glassy
polymers and, where appropriate, how they relate to the work presented in
the subsequent chapters.

1.2 THE GLASSY STATE

1.2.1 GLASSY POLYMERS

Glassy polymers as we know them today were developed in the 1950s and
were considered originally as a replacement for cellulosic esters that had

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
2 Introd uction

Table 1.1 Chemical formulae, abbreviations and approximate glass transition tem-
perature of some common glassy polymers

-("'6
Polymer Chemical repeat unit Abbreviation Tg(OC)

Polystyrene PS 100

Poly(vinyl chloride) -CH-CH- PVC 80


2 I
CI
Poly(methyl -CH -CH- PMMA 105
methacrylate) 2 I
COOCH 3
CH 3

Polyca rbonate ~~~COO- PC 145


CH 3
0

Polyethersulphone
~~~O- PES 220
0

been introduced as the first synthetic thermoplastics earlier in the century


(Haward, 1975). Perhaps the most important glassy polymer that was
introduced over this period was polystyrene (PS). Although it was consider-
ably more brittle than the cellulosic materials, it soon became used widely
because it could be produced relatively cheaply and had good processing
properties. Moreover the problem of brittleness was relieved by the intro-
duction of rubber toughening (Chapter 8). Polystyrene soon became both
widely used in practical applications and the preferred material for academic
studies of the physical properties of glassy polymers.
In the past 40 years the application of glassy polymers has developed
considerably. The use of poly(methyl methacrylate) (PMMA), with its
excellent optical properties, has been extended to general thermoforming
and moulding applications. Polycarbonate has also been introduced and its
clarity and toughness have been exploited widely. The chemical formulae of
a number of commonly studied glassy polymers are given in Table 1.1 along
with the abbreviations that are used for them in this and subsequent
chapters of the book.
The glassy state 3
1.2.2 THE GLASS TRANSITION
Many materials, not only polymers, may be found in the glassy state. The
simplest way of considering the transition to the glassy state is through the
rapid increase in viscosity that occurs when liquids are cooled and when
crystallization does not occur. Most simple compounds such as metals will
crystallize on cooling below their melting temperature but in more complex
materials, such as polymers, the rate of crystallization may be so slow that
a glass is formed before a significant proportion of the material has
crystallized. In other cases, especially with polymers that have a low level of
structural symmetry, such as atactic vinyl polymers (Young and Lovell,
1991) the formation of a crystalline structure is not possible and the glass
forms naturally as the low temperature state of the polymer.
The glassy state can be characterized from an experimental point of view
through the changes that occur in a polymer (or any other glass-forming
material) and it is cooled at constant pressure. Once the material is cooled
below its melting temperature, a supercooled liquid is obtained. The
viscosity of this liquid increases very rapidly as the temperature is reduced
until a situation is reached where the rate of intermolecular rearrangement
is slow compared with the time scale of the experiment, i.e. the material has
turned into a glass (Gee, 1970).
The physical properties of the melt, such as the Young's modulus, specific
heat capacity and thermal expansion coefficient, change in a characteristic
manner in the vicinity of the glass transition and, in principle, any of these
can be followed to investigate the glass transition. The classical way of
following the glass transition is by measuring specific volume as a function
of temperature as shown in Figure 1.1, although similar plots could be
obtained to describe changes in the other properties.
The representation shown in Figure 1.1 assumes that. when cooled, a
substance can either crystallize or transform to a glass depending upon the
experimental conditions. If crystallization takes place there is a step change
in volume (nearly always a decrease) at the melting temperature Tm' On the
other hand, if the melt is cooled and crystallization does not take place, or
is prevented, then at a temperature T. there is a transition from a super-
cooled melt to a rigid glass. Considering the case of glass 1 the volume-
temperature trace joins the curve of the glass at ~l where the solidification
process is terminated. The glass transition temperature is defined as the
point of intersection of the extrapolated curves for the glass and melt. Figure
1.1 shows also that the behaviour depends upon the rate of cooling, with
the value of ~ being lower for a more slowly cooled glass. Thus, it also
follows that a glass is not in a state of internal thermodynamic equilibrium
and its properties are not fully defined when the temperature and pressure
are fixed.
4 Introduction

I I

Figure 1.1 The volume-temperature relations for a polymer glass. Glass 1: fast
cooling; glass 2: slow cooling (Tm = melting point, Te and Tgi are temperatures at
the beginning and the end of the vitrification process, Tg = glass transition tempera-
ture, Te = freezing-in temperature). (Source: Rehage and Borchard, 1973.)

Although there is a large body of literature dealing with the process of the
glass transition, there are still many unanswered questions concerning the
physics of the glass formation process and the nature of the glassy state
(McKenna, 1989). The most fundamental question that is still unanswered
concerns whether or not the glass transition is a true second-order ther-
modynamic transition or a kinetic phenomenon. At this point it is best to
consider the definitions of thermodynamic transitions.
A first-order transition is one for which the free energy as a function of
any given state variable (V, P, T) is continuous, but where the first partial
derivatives of the free energy with respect to the relevant state variables are
discontinuous. Hence, for a first-order transition the Gibbs free energy G at
the transition temperature is continuous but (aG/aT)p and (aG/aph are
discontinuous. For a first-order transition such as melting or vaporization
there are discontinuities in entropy S, volume V and enthalpy H since
(McKenna, 1989)
aGJ
[aT p
=-S (1.1)

[~~l = V (1.2)
The glassy state 5

[ 8(G/T)] (1.3)
8(1/T) p = H

For example, there is a step change in Vat the Tm as shown in Figure 1.1.
In the case of a second-order transition, there is a discontinuity in the
second partial derivatives of the free energy function but continuity of both
the free energy and its first partial derivatives. Hence there are no discon-
tinuities in S, V or H at the temperature of the transition but there are
discontinuities in the variations of heat capacity Cp ' compressibility" and
thermal expansion coefficient IX with temperature since (McKenna, 1989)

(1.4)

[88pG] = [8V]
2
2 8P
T T = - "v (1.5)

~[[8(G/T)] ] _ [8H] _ C (1.6)


8T 8(1/T) p p - 8T p - p

(1.7)

There are a variety of phenomena in materials science that may be termed


second-order transitions. They include order-disorder transitions found in
some metal alloys and the onset of ferromagnetism.
It is now possible to consider whether the glass transition can be
considered as a thermodynamic transition. Figure 1.2 shows the temperature
dependence of the thermodynamic quantities G, V, Hand S and the
derivatives IX, Cp and " for first and second order transitions and for the
glass transition (Rehage and Borchard, 1973). It can be seen that the glass
transition appears to have similarities with a second-order transition. There
are, however, some significant differences. In contrast to the second-order
transition the values of IX, Cp and " are smaller below ~ than above it.
Another difference is found at different cooling rates when ~ shifts to higher
temperatures as the cooling rate increases. This could not happen for a true
second-order transition (Haward, 1975).
The features described above indicate that the process of glass formation
cannot be regarded as a true transition in the thermodynamic sense, but
rather as an inhibition of kinetic processes. The fundamental reason for this
is that there does not exist an internal thermodynamic equilibrium on both
sides of ~, as would be the case for a second-order transition. Because of
this it has been suggested (Rehage and Borchard, 1973) that it might be
better to use the term 'glassy solidification' to describe the process.
6 Introduction

(0) (b) (c)

G G G

Melt
7;, r,

/ -I
~
II)
I
:t I
II)
:to
II)
:to
~ I 'I
",-0 ~ /1I
~ ,
I

T" 7;, r,

~
I
': ': r--
" I
,,"
0
r--
I
\)"
.;
\J"
0
I
I
I I
I ----'I
I

T" T T" T T, T

Figure 1.2 Schematic representation of the changes with temperature of the free
energy and its first and second derivatives for (a) first order, (b) second order and (c)
glass transitions. (Source: Rehage and Borchard, 1973.)

Although there is much to commend this suggestion, such terminology is


not in general use and the term 'glass transition' will be used throughout
this book.

1.2.3 STRUCTURE IN GLASSY POLYMERS

(a) Wide Angle X-ray Scattering


It is possible to describe in considerable detail the exact arrangements of the
atoms in polymer crystals in three-dimensional space (Young and Lovell,
1991). In amorphous polymers, however, because of the 'random' nature of
their structure it is impossible to be so precise in defining the atomic
arrangements. The technique of choice in determining the structure of
materials on the atomic level is wide angle X-ray scattering (WAXS), which
is used widely to determine both crystal structures and the degree of
The glassy state 7

crystallinity for semicrystalline polymers. The typical W AXS pattern from a


semicrystalline polymer consists of a series of crystalline peaks on a diffuse
amorphous background. The position of the crystalline peaks gives the
Bragg spacings of the lattice planes in the polymer crystals. The degree of
crystallinity may be estimated by dividing the area under the crystalline
peaks by the total area under the scattering curve (Young and Lovell, 1991).
It is also possible, however, to obtain useful information upon the structure
of amorphous polymers using W AXS from the shape of the diffuse scattering
curve (Fitzpatrick and Ellis, 1973; Mitchell, 1987).
The W AXS from amorphous polymers is rather diffuse in nature. This
leads to the requirement of more demanding experimental techniques than,
for example, are required to obtain scattering patterns from crystalline
inorganic materials, where only the spacing of the Bragg peaks is normally
of interest. It is necessary to extract the elastic single scattering intensity /(s)
from the experimentally measured scattering function which contains the
effects of polarization and absorption as well as incoherent and multiple
scattering. [The parameter s = 4 n sin 8/A where 28 is the angle between the
incident and scattering beam paths and A is the wavelength of the radiation.]
Procedures for doing this have been described by Mitchell (1987). Figure 1.3

PNBMA
Corrected & normalised
intenSity function

Independent coherent
8. incoherent scutterif'g

o 2 3.
s(A-')
4 5 6

Figure 1.3 A fully corrected and normalized (scaled to electron units) X-ray
scattering intensity function 1(5) for poly(n-butyl methacrylate). The dashed curve
indicates the independent component of the scattering which includes the incoher-
ent scattering. (Source: Mitchell. 1987.)
8 Introduction

shows a fully corrected and normalized scattering function for poly(n-butyl


methacrylate). The dashed line in the figure represents the scattering that
would be obtained in the absence of any structural correlations. This can be
subtracted from the solid curve to give a quantitative scattering function free
from any experimental aberrations.
The conventional way of interpreting the scattering functions for glassy
polymers is to compare them with those derived from sophisticated models
(Mitchell, 1987). There are, however, some simple principles that emerge and
the features that determine the form of the scattering functions can be
conveniently separated into inter- and intrachain structures. For example,
there is an intense peak at around s '" 1.5 A- 1 in the scattering function for
natural rubber, which is normally interpreted as being due to ordered arrays
of parallel chain segments that may occur in the structure (Mitchell, 1987).
Because of the repeating nature of the polymer backbone, there will also be
regular aspects of the intrachain structure along the backbone of the
polymer molecule but they will be disturbed by factors such as irregular
tacticity and random trans and gauche conformations. The complete form
of the scattering function for an amorphous polymer will be determined by
a combination of these inter- and intrachain structural correlations which
can only be considered properly using the structural models (Mitchell,
1987).
(b) Small Angle X-ray Scattering
Small angle X-ray scattering (SAXS) can be used to characterize the
microstructure of materials on the 100-1000 A level and has been proven to
be a very useful technique of structural characterization in glassy polymers
(Fitzpatrick and Ellis, 1973). Early attempts to obtain SAXS patterns from
glassy polymers appeared to show evidence for structural variations in the
range 50-2000A. However, this was subsequently shown to be an artefact
due to the presence of voids, additives impurities, etc., and when SAXS is
undertaken upon specially prepared samples there is very little evidence of
particle scattering on this scale (Wendorff, 1987).
One area where SAXS has proved to be extremely useful is in the
characterization of the microstructure of block copolymers (Folkes and
Keller, 1973). These materials often have a microstructure on the 100-1000A
level that is sufficiently well defined that it can be considered to be
pseudocrystalline (Chapter 10). The technique has also been found to be
useful for the investigation of the formation of crazes (Chapter 6) in both
untoughened and toughened (Chapter 8) glassy polymers. Recent studies
have shown that, in conjunction with synchrotron radiation, SAXS can be
used to distinguish between the formation of either voids or crazes during
the deformation of rubber-toughened PMMA (Lovell, Sherratt and Young,
1996).
Stiffness and strength 9

(c) Microstructure

There has been some debate over the possibility of glassy polymers having
a microstructure on the 10 nm level from the appearance of the nodular
features that have been observed in non-crystallizable polymers, such as
atactic polystyrene, using electron microscopy (Yeh, 1972). Careful work by
Thomas and Roche (1979,1981) has demonstrated clearly that such features
are artefacts caused by the focusing conditions in the electron microscope.
The nodular features are observed only when the microscope is out of focus
and are not present when it is focused correctly. This finding is clearly
consistent with the lack of a well defined microstructure demonstrated by
SAXS measurements. What is not so clear is if similar nodular or granular
features observed in amorphous specimens of crystallizable polymers such
as poly(ethylene terephthalate) are similar artefacts or evidence of precur-
sors of crystallization (Geil, 1987).
There is strong evidence that the microstructure of amorphous polymers
in the glassy state is essentially statistical in nature. Many years ago Flory
(1949) postulated that molecules in a bulk amorphous polymers would be
in the form of random coils and have a conformation equivalent to that in
a theta solvent. For many years this was a working hypothesis for re-
searchers interested in glassy polymers and has served them well. Indeed,
the statistical theory of rubber elasticity (Treloar, 1975) is based upon this
premise and it has been found to be very successful. It is only relatively
recently that the suggestion of Flory has been proved experimentally from
small angle neutron scattering studies upon mixtures of protonated and
deuterated polymer molecules (Wignall, Ballard and Schelten, 1976). Never-
theless, it is quite difficult to prove that a glassy polymer does have a
structure with some degree of order. Boyer (1987) has examined the
possibility that there might be chain folding in amorphous polymers, albeit
less perfect than that in crystalline polymers. He demonstrated that the
experimental evidence makes it difficult to distinguish unambiguously be-
tween molecules in either random coil or chain-folded conformations in
amorphous polymers.

1.3 STIFFNESS AND STRENGTH

1.3.1 ELASTIC DEFORMATION


Glassy polymers can exhibit a wide range of elastic properties depending
upon the testing conditions. The variation of Young's modulus E with
testing temperature T for a typical amorphous polymer is shown in Figure
1.4 (Young and Lovell, 1991). The polymer is glassy at low temperatures
with a modulus ('" 3 X 109 Pa). As the test temperature is increased the
10 Introduction

10

9 Glass

-;;; Rubber
~ 7
w
Ol
Cross-linked
.36

5
Linear

Temperature

Figure 1.4 Typical variation of Young's modulus E with temperature for a polymer,
showing the effect of crosslinking upon E in the rubbery state. (Source: Kinloch and
Young, 1983.)

modulus falls rapidly through the region of ~ where the polymer is


viscoelastic and the modulus is very rate and temperature dependent. At a
sufficiently high temperature the polymer becomes rubbery. If the polymer
is not crosslinked the modulus decreases rapidly and it flows like a viscous
liquid. Otherwise, if the polymer is crosslinked the modulus actually
increases with increasing temperature (Treloar, 1975), although over a
narrow range of temperature the modulus remains approximately constant
at '" 106 Pa.
The elastic behaviour of polymers reflects the deformation of the structure
on the molecular level. For example, in the case of high-modulus polymer
fibres such as aromatic polyamides (e.g. Kevlar or Twaron) deformed
parallel to their fibre axes, elastic deformation takes place by a combination
of backbone covalent bond stretching and bond angle opening (Young,
1995). This requires very high molecular forces and leads to values of
Young's modulus well in excess of 1011 Pa which are comparable with
metals such as steel.
In isotropic glassy polymers the molecules are coiled in a randomly
oriented frozen microstructure as explained in the previous section. Hence
elastic deformation induces relatively easier forms of intramolecular defor-
mation such as bond rotation (Bowden, 1968) and is dominated by weak
intermolecular van der Waals interaction. This leads to glassy polymers
Stiffness and strength 11

having significantly lower Young's moduli ('" 3 X 109 Pa) than polymer
fibres.
The elastic deformation of amorphous polymers in the rubbery state has
received a great deal of interest over the years and is now well understood
on the molecular level using statistical thermodynamics (Treloar, 1975). The
application of stress causes the polymer chains to uncoil and since a
stretched chain has fewer conformations available to it, its entropy is
reduced. Removal of the applied stress allows the chains to increase their
entropy by readopting their randomly coiled conformations, leading to the
well known phenomenon of rubber elasticity controlled essentially by an
'entropy spring'.
In many ways it is remarkable that most glassy polymers tend to have
similar values of Young's modulus, typically in the range 2-4 x 109 Pa, and
even with highly crosslinked polymer resins it is difficult to achieve modulus
values in excess of 6 x 109 Pa. Haward and coworkers have investigated this
phenomenon through the preparation of crosslinked polymers with in-
creased modulus through the use of high pressure polymerization (Rackley
et al., 1974; Price, Haward and Parsons, 1979). They polymerized both
divinyl and trivinyl benzene at high pressures, and found that the extent
of conversion of the vinyl groups, and the density and shear modulus
of the polymers all increased with the temperature and pressure of polymer-
ization. In fact an approximately linear correlation is found between
Young's modulus and density of the polymers as shown in Figure 1.5.

0"

o
J
~;----~::::----~---1·20 J
Density (Mg/m 3 )
Figure 1.5 Variation of shear modulus with density for a series of divinyl benzene
polymers polymerized at high pressures. (Source: Rackley et ai., 1974.)
12 Introduction

Rackley et al. (1974) concluded that this behaviour is an indication that the
elastic properties of the glassy polymers are determined largely by the
intermolecular forces such as van der Waals interactions. Moreover they
showed that non-crosslinked polystyrene displayed similar behaviour when
it underwent a significant increase in modulus after being densified by
cooling under high pressure.
One aspect of the mechanical behaviour of polymers is the way in which
their response to applied stress or strain depends upon the time-period of
loading (Young and Lovell, 1991). Elastic materials obey Hooke's law

Str~ss syst~m strain syst~m

Figure 1.6 Schematic representation of the variation of stress and strain with time
indicating the input (I) and response (R) for different types of loading. (a) Creep; (b)
relaxation; (c) constant stress rate and (d) constant strain rate. The dashed lines are
for elastic materials and the solid lines for viscoelastic ones. (After Williams, 1973.)
Stiffness anci strength 13

whereby the stress is proportional to strain and independent of strain rate.


Viscous materials tend to obey Newton's law whereby the stress is propor-
tional to the strain rate and independent of the strain. The behaviour of
many polymers can be thought of as being somewhere between that of
elastic and viscous materials and is often termed 'viscoelastic'. The phenom-
enon of viscoelasticity is most obvious for amorphous polymers at tempera-
tures around ~. For glassy polymers that are well below their ~ it is often
possible to investigate their mechanical properties at a low or moderate
stress without having to consider time-dependent behaviour to any great
extent. However, for a complete understanding of the physical properties of
glassy polymers it is necessary to consider viscoelastic properties properly.
Examples of the mechanical behaviour for a viscoelectric polymer are
given in Figure 1.6 (Williams, 1973). This shows the variation of stress (j and
strain e with time t for a polymer specimen subjected to four different
deformation histories. Specific examples are given for creep, stress relaxation
and constant stress rate and constant strain rate deformation. The response
of the stress or strain is indicated in each case for the applied strain or stress
and the behaviour of an elastic material is indicated by dashed lines. It is
often possible to analyse the behaviour of viscoelastic materials using simple
mechanical models (Young and Lovell, 1991) but the reader is directed
towards more specialized texts (McCrum, Read and Williams, 1967; Ferry,
1980) for a full account of viscoelasticity.

1.3.2 PLASTIC DEFORMATION


Glassy polymers are generally only linearly elastic and obey Hooke's law at
vanishingly small strains. At sufficiently high strains, if they do not first
suffer brittle fracture, they can undergo bulk plastic deformation (Young
and Lovell, 1991). Two principal mechanisms of plastic deformation have
been identified for glassy polymers - yielding and crazing -- both of which
are described in detail in Chapters 4-6 and so will not be considered in any
depth here. The onset of plastic deformation by shear yielding is character-
ized by a permanent change in specimen shape taking place essentially at
constant volume (Bowden, 1973). This should be contrasted with crazing
which leads to an increase in the measured specimen volume as crazes,
which can be thought of simply as microcracks bridged by fibrils, are
nucleated (Andrews, 1973). It is known that there is an intimate relationship
between crazing and fracture in some glassy polymers, and the mechanisms
of shear yielding and crazing are both important in controlling the strength
and toughness of glassy polymers (Kinloch and Young, 1983) as described
in the next section.
14 Introduction

1.3.3 STRENGTH AND TOUGHNESS


The fracture of a body involves the creation of new surfaces and in a glassy
polymer involves the breaking of both primary (covalent) and secondary
(e.g. van der Waals or hydrogen) bonds. It is possible to estimate the
theoretical strength O'lh of a glassy polymer using the relatively simple
analysis of Kelly and MacMillan (1986). It can be shown (Young, 1989) that
the theoretical strength of most solids would be expected to be of the order
of one-tenth of the Young's modulus, i.e. E/I0. Since most glassy polymers
have values of modulus of 2-4 GPa, O'lh would be expected to be in the range
200-400 MPa. It turns out that this is a reasonable estimate of the upper
limit of the strength of glassy polymers, which often have measured fracture
strength values of up to about 150 MPa (Seitz, 1993). This should be
contrasted with many other materials for which their measured strength
values are orders of magnitude less than E/I0 due to the presence of defects.
This implies that glassy polymers are somewhat tolerant of the presence of
defects (Kinloch and Young, 1983) - one reason why they have found such
widespread use.
The analysis described above does not take into account the nature of the
chemical bonding in the material. The covalent C-C bond found in the
backbone of the molecules in glassy polymers (Table 1.1) is known to be
very strong (diamond is essentially a three-dimensional network of C-C
bonds). Conservative estimates of the strength of the C-C bond give values
of least 3 x 10 - 9 N (Young, 1989). The measured strength of isotropic
glassy or crystalline polymers can be maximized by raising the contribution
of C-C bonds as compared to van der Waals forces. This may be done by
arranging the C-C bonds either in a crystal or an oriented fibre. The
cross-sectional area occupied by each chain in a polyethylene crystal is
about 0.18 nm 2 and this implies that the theoretical strength of poly-
ethylene crystals deformed in a direction parallel to the chain direction
is at least of the order of 17 GPa. Such values of strength are not reached
even for the highest modulus polyethylene fibres, although fibres with
a tensile strength value of the order of 5 GPa have now been prepared
(Ward, 1987). In a glassy polymer the packing of the polymer chains will not
be as efficient as in polyethylene crystals and the orientation of the
molecules will be random. Nevertheless, it is clear that the measured
strength values of glassy polymers of the order of 150 MPa imply that
fracture takes place without the widespread breaking of covalent bonds, and
is more concerned with the failure of the secondary bonding. It should be
noted that this is consistent with the elastic deformation of glassy polymers
being dominated by the deformation of secondary bonds, as described in
section 1.3.1.
Stiffness and strength 15

250

0
200

a;-
a..
e.
~
150
0 0
00

c,
C
~
Ui 100 0
Ql
B
~ 8 0
50
0

0
0

Number of Bonds/m 2 x1 0- 18

Figure 1.7 Variation of the brittle strength with the number of backbone bonds per
unit area n B for a number of polymers. (Source: Seitz, 1993.)

There have been several attempts to quantify the importance of covalent


bonding upon the fracture of glassy polymers (Vincent, 1972; Seitz, 1993).
Vincent (1972) related the brittle strength of a series of polymers to the
number of backbone bonds per unit area in the polymer. nB , calculated
through the following equation:

(1.8)

where N A is the Avogadro constant, 1m the length of the repeat unit and V
the molar volume. Figure 1.7 shows a plot of the measured brittle strength
of a series of polymers against nB . The slope of the line -0.04 x 1O- 9 N
then represents the strength of the bonds (Seitz, 1993). Since this is only
about 1% of the strength of a covalent C-C bond (see above), it implies
that the fracture of glassy polymers involves the breaking of no more than
1% of the covalent bonds in the material (Seitz, 1993).
The fact that the level of strength of many materials is well below (Jlh due
to the presence of defects was recognized many years ago (Griffith, 1921).
This led to the development of the Griffith equation (Kinloch and Young,
1983) which relates the strength of a material (Jf to the size of defects or
cracks a through

(Jf -
_ (2Ey)1/2 (1.9)
na
16 Introduction

PMMA

20

o 2
a(mm)

\
\ PS

40

~ ao
Q.
L

b
20

o 5 10
a(mm)

Figure 1.8 Dependence of the fracture stress O"f upon crack length a for PMMA and
PS. The inherent flaw size ao is indicated for PS. (After Berry, 1971.)

where y is the surface energy of the material. In the early 1960s Berry used
this equation to examine the dependence of {Jf upon crack length for
tensile specimens of glassy polymers such as PMMA and PS (Berry,
1971). The behaviour is shown in Figure 1.8 and it can be seen that the
form of equation (1.9) is obeyed in that the strength of the polymer increases
as the crack length is reduced. However, the increase in strength with
decreasing crack length does not go on indefinitely, since when the flaw size
is reduced below a critical level (about 1 mm for PS and 0.07 mm for
PMMA at room temperature) {Jf becomes independent of flaw size. This is
Entanglements 17

because the polymers behave as though they contain natural flaws of these
critical sizes that are termed 'inherent' or 'intrinsic' flaws. Such flaws,
however, are not present in the material before deformation but appear to
be formed on loading. For PS and PMMA, inherent flaws may be related
to crazes (Chapter 6) that have nucleated under stress and then broken
down to form cracks during deformation (Kinloch and Young, 1983).
Another point to note is that values of y of the order of kJ m - 2 are
obtained for PMMA and PS, which is well above the surface energy of the
glassy polymers. It is now recognized that the term 2y in equation (1.9)
should be replaced by a fracture energy Gc which takes into account the
energy required both for the creation of new surface and for any other
energy dissipation processes that occur in the vicinity of the crack tip (e.g.
shear yielding or crazing; Kinloch and Young, 1983). Hence equation 1.9
becomes

_(EG )1 /2
l1f -
c
(1.10)
na

This equation is the basis of fracture mechanics and has wide-ranging


applications to the analysis of the fracture of glassy polymers (Chapter 7).
An alternative approach to the fracture of brittle solids (Kinloch and Young,
1983) considers the stress field around a sharp crack and shows that fracture
can be characterized by a critical value of the stress intensity factor Kc. In
fact for linear elastic material the parameters Kc and Gc are related through
(1.11)
This fracture of many glassy polymers can be analysed using linear elastic
fracture mechanics (Kinloch and Young, 1983). This can be done as long as
the mechanical behaviour of the polymer is linearly elastic even though
small scale inelastic deformation or yielding may take place. With materials
with higher levels of toughness, such as rubber-toughened polymers (Chap-
ter 8), more sophisticated approaches need to be employed, as outlined in
Chapter 8.

1.4 ENTANGLEMENTS

1.4.1 THE CONCEPT OF ENTANGLEMENTS


It has been known for a long time that when an isotropic glassy polymer
was deformed at normal temperatures it retained its new shape and reverted
to its original form only if it was heated above ~. These observations
naturally led on to the idea that there were permanent entanglements
18 Introduction

between the polymer chains that were responsible for preserving the
memory of shape (Hoff, 1952). At the time it was also observed that the
same materials also had a limited capacity for extension in a tensile test
which was assumed to be related to the straightening out of the polymer
chains (Haward, 1949).
However experimental evidence for these ideas remained rather limited
until the late 1960s. At that time Pinnock and Ward (1966) and Allison and
Ward (1967) published their work on the stretching of poly(ethylene
terephthalate) (PET) filaments and developed the concept of an underlying
macromolecular network, which determined the ultimate extensibility and
ensured a return to the original form when the temperature was raised. In
the same period Fletcher, Haward and Mann (1965) proposed that the
stresses required for large deformation of a thermoplastic were composed of
a viscous process and an elastic component similar to that of a rubber. This
concept was formalized and quantitatively related to known macromolecu-
lar dimensions by Haward and Thackray (1968). Later Argon (1973)
pointed out that the use of a rubber elasticity equation was simply a way of
representing changes in entropy during the extension of a long chain
polymer and its consequent transformation from a random to an oriented
conformation.
In the period during which these proposals have been developed there has
also grown up a body of knowledge surrounding the idea of entanglements
and their influence on mechanical properties at temperatures above ~.
These will now be briefly presented. At the same time it seems appropriate
to register a certain reservation concerning the actual concept of entangle-
ments as a constant topological feature extending above and below I'g.
When a polymer glass consisting of long chain molecules is deformed so
as to orient them in a state of reduced entropy, they are prevented only by
frictional forces from returning to conformational equilibrium. When the
viscosity is reduced by a rise in temperature, they resume their random
configurations. The minimal requirement for the specimen comprising them
to return to its original shape is that the distribution of the centres of gravity
of the polymer molecules after retraction should be statistically similar to
what it was before the deformation took place. The presence of chemical
crosslinks or of physical entanglements should be a sufficient though not a
necessary condition for this result to occur. A situation has therefore arisen
in which the interaction of polymer chains, their structure and entangle-
ments play an increasing part in the understanding of mechanical properties.
It therefore seems appropriate at this stage to give a brief account of the two
classical methods by which the entanglements in glassy polymers have been
estimated.
Entanglements 19

1.4.2 MEASUREMENT OF ENTANGLEMENTS ABOVE THE CLASS


TRANSITION TEMPERATURE

Most of the established techniques for estimating the concentration of


entanglements between polymer molecules have been carried out above ~.
Two distinct methods have been employed. One of these takes advantage of
the occurrence of a rubbery state at temperatures not too far above ~. The
other method depends on the determination of the melt viscosity (at low
rates of shear) as a function of molecular weight. A fuller account of these
techniques is given by Ferry (1980), to whom the reader is referred. Only a
very short summary of the principles assumed is given here, which is
intended to clarify the nature of the measurements under discussion.

(a) Entanglement Molecular Weight from the Plateau Modulus


When a polymer glass is heated to a temperature somewhat above ~, its
properties enter a rubbery stage where the amount of flow, especially at low
stresses and short time is small. This makes it possible to measure a 'plateau
modulus' G~ as shown in Figure 1.9, where it can be seen that the value
obtained is slightly affected by the temperature and frequency of measure-
ment. Ferry (1980) therefore recommends an alternative method in which
G~ is obtained by integrating the loss modulus Gil against the logarithm of
the frequency of measurement, and this method is now commonly applied.
Although results for G~ have been widely quoted, doubts have been
expressed about the accuracy of some of the work (Aharoni, 1986).
In any case the values of G~ may be used to estimate the entanglement
molecular weight Me from the Gaussian equation (Treloar, 1975)
Me = pRT/G~ (1.12)

where p is the density.


From Me' the entanglement density per unit volume n [note that 'v' is
also used for this quantity] may be derived using the relation
(1.13)

where N A is the Avogadro constant.


In principle it is possible to compare the values of n obtained in this way
with those provided by the Langevin treatment of large strains at normal
temperatures (sections 5.7-5.9). Similarly G~ could be compared with
measurements of the Gaussian strain hardening modulus Gp from the
application of the Gaussian equation to strain hardening in tension below
~ (section 5.25), but serious problems arise from the variation of Gp with
20 Introduction

7 PS 160°C

C?
e:. 4
b
Cl
.2
3

°_6 -5 -4 -3 -2 -1 o 2 3 4 5

log waT (5- 1)

Figure 1.9 The storage modulus of narrow distribution polystyrenes plotted logar-
ithmically against frequency reduced to 160°C. Viscosity average molecular weights
from left to right, x 104 : 58, 51, 35, 27.5, 21.5, 16.7, 11.3, 5.9,4.6. (Source: Onogi,
Masuda and Kitagawa, 1970.)

temperature (Haward, 1993) so that only large differences in Gp are likely to


be significant. Similar problems of temperature variability arise if the strain
hardening constants are determined using the Langevin equations (Arruda,
Boyce and Jayachandran, 1995).

(b) Measurements from Viscosity


When the melt viscosity of a glassy polymer above 1'g is plotted against
molecular weight using logarithmic scales, a graph is obtained which
consists of two intersecting lines as illustrated in Figure 1.10. At low
molecular weights the viscosity (rf) is found to be simply proportional to
the molecular weight, but as the molecular weight is increased the results
suddenly turn upwards and the line follows a dependency of the type
11 = M 3 . 4 (Ferry, 1980). The point at which this occurs is easily determined
Entanglements 21

6.0 r----r------r----~~

4.0
~

5:
0
Cl
>-
0
:;;:
> 2.0
0>
2

2.0 3.0 4.0


log 2Pw
Figure 1.10 Dependence of steady-flow viscosity on number of chain atoms per
molecule for polystyrene fractions at 217°( (1 poise = 0.1 Nsm- 2). (Source: Fox,
Gratch and Loshaek, 1956.)

on a double logarithmic plot and the value obtained is denoted as Me'


which is higher than Me. Generally it is found that Me ~ 2Me, though
factors of up to 2.5 have been reported (Ferry, 1980). Me is often considered
to define the molecular weight below which glassy polymers become very
brittle (section 5.12). Generally measurements of melt viscosity have to be
made at temperatures above those appropriate to the estimation of the
plateau modulus (e.g. for polystyrene Ferry reports measurement on the
former at 217°C and the latter at 160°C).
In applying entanglement theory to the study of mechanical properties, it
is also necessary to consider the conformation and dimensions of the
polymer chain, which may be determined either in solution in a theta solvent
or in the glassy state. Such measurements generally result in the estimation
of the quantity R~/L where R~ is the mean square end-to-end distance of the
unperturbed polymer chain with an extended or contour length L. Then,
according to Kuhn, the statistical chain may be divided up into N freely
jointed statistical lengths lk (Kuhn length), such that Nlk = L and Nl~ = R~.
This finally leads to the relation
L/Ro = N 1 / 2 (1.14)
For an entangled polymer with Ne Kuhn lengths between points of
entanglement and a contour length Le = Nelk we may write R; = Nel~,
22 Introduction

Table 1.2 Entanglement data for a number of glassy polymers (molecular weights
ingmol-')

Polymer M ea Mb
e lk (nm)b Le(nm)b Me
e
Polystyrene 17851 35000 1.69 103 32000
Poly(methyl methacrylate) 9200 31530 1.70 97 18000
Poly(vinyl acetate) 8667 24540 1.60 87.8 25000
Poly(vinyl chloride)d 6250 1.19 30.8 10700
Poly(phenylene oxide) 3620 3360 2.46 15.1
Polycarbonate 2495 4875 2.94 28.8 4300
Poly(ethylene terephthalate) 1450 3270 1.11 21.3
Cellulose acetated 1800 16 5
Poly(butyl isocyanate)b.d 7425 100 19.8

·Seitz (1993). These results were used in a preliminary form by Donald and Kramer (1982).
bAharoni (1983, 1986).
'Calculated by Wool (1993).
dSome crystallinity present in the solid state.

so that the maximum possible extension between fixed entanglements is


given by
(1.15)
It should be noted that the quantity nNlk represents the total length of
the polymer chains in the system, so that when lk is constant nN is constant
as assumed by Boyce (Chapter 5). In the application of the rubber elasticity
theory described in this chapter, nand N (and therefore Amax) are derived
directly from the Langevin model and ought in principle to be similar to
that derived from the entanglement molecular weight Me as measured by
the plateau modulus. There is, however, a potential problem as the measure-
ments are normally undertaken at different temperatures.
Table 1.2 lists some of the quantities associated with entanglements
determined for a number of common glassy polymers.

(c) Calculation of Entanglements from Chain Structure


Further evidence on the relation between chain conformation and entangle-
ment comes from the model calculations of Wu (1989) and Wool (1993) on
typical glassy polymers. Both workers employed the chemical chain struc-
ture to predict the number of entanglements estimated as the proportion of
hook- or loop-like configurations. Although there were significant differen-
ces in the two calculations, both found that the number of entanglements
increased with a quantity represented by Coo, a measure of the extension of
the polymer chain defined as follows (Kurata and Tsunashima, 1975):
(1.16)
Entanglements 23

where there are na actual chain units separated by rotations. In a vinyl


polymer with a linear carbon chain 1= O.l54nm. These calculations were
based on chain units derived from the chemical formula and not from the
statistical units defined by Kuhn. Nevertheless, both COO and R~ increase
when I or lk are large. Both workers found that the higher the value of COO
and the more extended the chain, the greater the density of entanglements
and the smaller the value of Me' a conclusion shown to be in line with
experimental results (Table 1.2).
Although these models are supported by the measurements reported in
Table 1.2, many difficulties are encountered when measurements on a wider
range of polymers are considered. As shown in the table, the measurements
of Seitz (1993), the calculations of Wool and the compilations of Aharoni
(1983, 1986) for characteristic glassy polymers follow a similar trend.
However, in his papers Aharoni brought to light a number of unexpected
difficulties when he reviewed literature estimates of chain conformation in
solution and values of G~, Me and Me measured in bulk on a large number
of polymers. He found that flexible polymers (containing many rotations
which affect chain conformation) can be grouped together when different
types of correlation are evaluated. This group includes all the commercially
important glassy polymers and includes those studied by Wu and Wool. On
the other hand there are small groups of semirigid (typically cellulosics) and
rigid (polybenzamide and the polyisocyanates) polymers for which it is
found that the Kuhn length lk exceeds the estimated entanglement length
(Table 1.2). This means that the points along the chain at which stress is
transmitted from one chain to another, imposing a constraint on free chain
mobility, are closer together than the length of a statistical link which
defines the end-to-end separation. It is difficult to reconcile these con-
clusions with the conventional idea of a localized topological entanglement,
which may be expected to mimic the behaviour of a chemical crosslink.

1.4.3 ALTERNATIVE MODELS FOR CONDENSED POLYMER SYSTEMS


A number of theories have been proposed which describe a dense polymer
system without introducing the idea of a topological entanglement. For
example, Graessley and Edwards (1981) considered the characteristics of a
concentrated polymer liquid, such as the plateau modulus G~, which had
formerly been attributed to entanglements. Instead, they evaluated the
limitations on the motion of a polymer chain caused by other macro-
molecules and proposed that G~ was determined more generally by the
interaction between different polymer chains. When estimated in this way,
its value depended on the length of uncrossable volumeless chain contours
and on their Kuhn step length lk. Their conclusions were shown to be in
line with measurements quoted by Ferry (1980) for flexible polymers.
24 Introduction

Another method of describing a dense polymer system was published by


De Gennes (1971). He proposed that the polymer molecule in the melt
above Tg is constrained within a tube by restrictions due to the presence of
other polymer molecules. The role of the neighbouring but separate polymer
chains is then to restrict movement to that of a wriggling or 'reptation'
process within the tube. This model has subsequently been elaborated to
describe many aspects of high polymer properties above the glass transition
temperature. Although there have not been many studies at lower tempera-
tures, Edwards and Viglis (1987) have extended the tube theory to the glassy
state. They postulated that deformation occurred by a process of straighten-
ing out planar meanders in the polymer chain between 'slip links' until it
attained the shortest (primitive) length along the tube. A further account of
the tube theory has been given by Muthukumar and Edwards (1989).
Although these theories offer promise for the future, they have not so far
been used to estimate entropy changes during the orientation of separate
polymer chains and do not yet have more than a limited application to the
problems outlined below.
The relation between molecular structure and mechanical properties may
also be studied by means of molecular dynamics modelling as described in
Chapter 2. In these studies McKechnie et al. (1993) modelled polymethylene
in the glassy state and showed that typical stress-strain curves could be
obtained. These showed marked strain hardening, the magnitude of which
could be increased if the glass was prepared in such a way that the
proportion of trans rotational isomers was increased. Under these condi-
tions the polymer chains had a greater persistence length (higher lk). Later
work showed that the deformed polymer would retract towards its original
shape if heated above the effective glass transition temperature. This work
is important as it shows that the key properties which led to the concept of
entanglement can be modelled successfully without recourse to a specific
assumption of this type. However a wide distribution of related motions and
topological environments for different monmeric groups are included so that
the model should also prove itself congruent with the assumption that
'entanglements' can dissociate at higher temperatures (Chapter 5). On these
lines dynamic modelling offers the possibility of a further clarifcation of the
mechanism of strain hardening in the future.

1.4.4 MECHANICAL PROPERTIES BELOW THE GLASS TRANSITION


TEMPERATURE

The hypothesis of entanglement has been introduced into the theory of large
deformation in two ways. As already noted, it was proposed by Haward and
Thackray (1968) and by Argon (1973) that the stress required to cause large
Entanglements 25

deformations contains a very significant strain hardening component which


is assumed to arise from conformational and consequent entropy changes in
the polymer chains. This stress can then be modelled using established
theories of rubber elasticity (Treloar, 1975). When this is done, however, it
is necessary to introduce the concept of entanglements as a substitute for the
chemical crosslinks which are present in a rubber. An account of the
development of this theory forms a substantial part of Chapter 5.
A distinct but related concept has been employed by Donald and Kramer
(1982) in the interpretation of crazing as described in Chapter 6. Briefly it
has been found that a craze consists of a thin sheet of locally extended
polymer fibrils (Figure 6.1) which can be widened only by drawing in
fresh material from the matrix. After reaching a characteristic extension
ratio, the material in the craze does not extend further. Donald and Kramer
then went on to show that this extension ratio approximated to the value
of A. max as defined above (equation 1.15) and so introduced the idea of
entanglement into the theory of crazing.
Our understanding of both large deformations in bulk and in crazing is
also influenced strongly by mechanical processes generally associated with
what is called 'plastic instability'. This describes a situation in which a
plastic deformation is accompanied by a faU in the remotely applied stress
and leads to a localization of plastic strain. This treatment, which was first
recognized by Considere and is described in Chapters 4 and 5, makes it
necessary to draw a distinction between the true strains and stresses present
in the material and the nominal stress which may be measured by a
machine. When nominal stresses are recorded, an apparent softening of a
material may be observed, as when necking takes place in a tensile test,
while the true stress-strain relation exhibits only continuous strain harden-
ing. Under these conditions deformation becomes localized in a neck until
a deformation is reached where the material has strain hardened sufficiently
for the undeformed polymer to be drawn into the neck ('natural draw ratio').
After passing through the neck, further deformation becomes very slow and
extension of the whole test piece occurs almost entirely by means of neck
propagation. Thus the extension of a craze as described in Chapter 6 has a
certain analogy with the tensile test and, arguably, is also subject to the
Considere criterion.

(a) Entanglement Theory and Large Deformations


The model for the large deformation process as outlined above is illustrated
in Figure 5.25. In the present context we are primarily concerned with the
use of a rubber elasticity equation to represent the changes in entropy
associated with the straightening of a polymer chain and the associated
26 Introduction

strain hardening process. To this end the concept of entanglement density,


represented by n, is used to replace the chemical crosslinks in a true rubber.
The latter are, of course, invariant with temperature and this leads to a
direct dependence of the elastic stress on the absolute temperature. If the
number of entanglements was also independent of temperature, then the
amount of strain hardening would be expected to increase with temperature
in the same way. However, the viscous component of the model, generally
representing the yield stress, invariably falls as the temperature is raised, as
illustrated in Figures 4.12 and 4.13 for polycarbonate and in Figure 5.33 for
PMMA. Such a combination of a falling yield stress and increasing strain
hardening would inevitably also require a steady fall in the natural draw
ratio with increasing temperature, a type of behaviour which is seldom if
ever observed. On the contrary, several workers have reported a relative
constancy in this quantity over a wide range of temperature. For example,
Andrews and Ward (1970) and Coates and Ward (1980) with linear
polyethylene, Allison and Ward with PET, and Zhou et al. (1995) have all
reported little change in the natural draw ratio at different temperatures, as
shown in Figure 1.11 for polycarbonate.

2.00 . . , - - - - - - - - - - - - - - - - - - ,

1.75
o
;:
IV
IX:
~
~ 1.50

~
...
:::J
IV 10 I
z 1.25
o Thickness =0.7 mm
IJThickness =1.3 mm 0riIT, ~ I rom
1.00 ;--.....--,----.--...----r---.--..---i
o 40 80 120 160

Temperature (OC)
Figure 1.11 The natural draw ratio of polycarbonate at different temperatures.
(Source: Zhou et al., 1995.)
Entanglements 27

A relative constancy of natural draw ratio with temperature requires that


the viscous and entropic components of stress change in a similar way. So
far, no direct information on this point is available for glassy polymers but
with linear polyethylene the ratio of the yield stress to the Gaussian strain
hardening constant Gp has been shown to be constant over the temperature
range 0-60°C (Haward, 1993). Similarly it has been found that for an
amorphous polyimide the strain hardening constant falls at higher tempera-
tures. Other, and more accurate work has been carried out on PMMA by
Arruda, Boyce and Jayachandran (1995) as described in section 5.7. Using
the Langevin model for rubber elasticity, they obtained good agreement
with experiment by applying a concept proposed previously by Raha and
Bowden (1972), in the course of studies on the deformation and birefrin-
gence of PMMA. The latter proposed that some of the entanglements
present could dissociate at higher temperatures, with the result that the
value of 'n' would decrease and the entanglement molecular weight would
increase at higher temperatures.

(b) Entanglements and Crazing


When Donald and Kramer (1982) introduced the idea of entanglements into
the study of crazing, they related the deformation in a craze formed at
normal temperatures to the value of Amax calculated from G~ measured
above ~. For a series of different glassy polymers they demonstrated not
just a good correlation between the two quantities but also a near identity
in the actual figures. However, when the extension ratio in a polystyrene
craze was measured as a function of temperature it remained constant at the
predicted level between 20 and 50°C, but above 50°C, especially with low
molecular weight materials, the extension ratio increased (Figure 6.19).
These results generally accord with those reported by Arruda, Boyce and
Jayachandran (1995) for PMMA. In later work Plummer and Donald
(1990) also went on to ascribe their increase in A to a loss of entanglement
but envisaged this as a process in which polymer chains escaped from the
constraint of surrounding molecules by a process of reptation according to
the tube model. This took place at the head of the fibrils as they separated
from the matrix.
These results offer an apparent conflict. The low temperature value of Amax
agrees with that predicted from G~ but diverges from it at temperatures
approaching that at which G~ was measured. However, a possible explana-
tion can be offered if we note that G~ is measured at low stresses and short
times which may reasonably be assumed to be insufficient for the reptation
process even at the higher temperature.
28 Introduction

In considering the relation of crazing to large deformations, it should be


appreciated that the former relates directly to A whereas the model for large
deformations operates in terms of stress. However, as already pointed out,
deformation in tension also provides a characteristic strain known as the
natural draw ratio, which is formed by a process which can be compared
with that of a craze fibril. Unfortunately most of the polymers which exhibit
crazing are also brittle and do not show bulk deformation in tension. An
exception is polycarbonate where Donald and Kramer (1982) derived a
value of Amax = 2.5 from G~ and measured a value of 2.0 for a craze. These
may be compared with values around 1.7 from Figure 1.11. It will be
appreciated that Amax is the ultimate limit of possible extension and so will
tend to exceed the other quantities. Thus the level of agreement between the
three measurements may be considered satisfactory.

1.4.5 POLYMER STRUCTURE AND LARGE STRAIN PROPERTIES

It is believed that the discussion above provides some diffuse but persuasive
indications of the relation between the structure of the polymer chain and
properties associated with large strains. This is the case in spite of real
doubts as to whether or not any actual structural features which may
properly be called entanglements really exist, or whether the effects ascribed
to them are only a way of summarizing the consequences of a more
dispersed interaction between polymer chains.
We find, for example, that crazing is characteristic of polymers with a
fairly low entanglement density. As this quantity increases the deformation
of craze fibrils is reduced and eventually shear yielding takes place (Chapter
6). Similarly, from the study of large deformations we find that a high
entanglement concentration increases strain hardening, which tends to
reduce the natural draw ratio after necking and in the limit might be
expected to eliminate neck formation (rubbers do not neck). If we now go
on to ask which structures determine entanglement density we find that
different theories tend to give the same answer. We find that the models of
Wu (1989), of Wool (1993), the conclusions of Graessley and Edwards
(1980) and the molecular dynamics of McKechnie et al. (1993), together
with much early speculation, all lead to the conclusion that a high value of
lk (or of the persistence length) is associated with more entanglement and
higher strain hardening. Further it is found that the cellulosics and poly-
isocyanates at the foot of Table 1.2, with the highest Kuhn length, also
exhibit uniform deformation with no necking (Haward, 1993). This raises
the further question as to whether it is possible to modify a polymer so as
to increase lk and to measure changes in properties. Attempts to do this have
been made by introducing a terephthalate unit into the polycarbonate chain.
References 29

This behaves as a rigid link (Birstein, 1977) and so should increase lk (Table
1.2). It has also been shown to increase strain hardening as indicated by a
reduction in the post-yield stress drop (Bubeck, Smith and Bales, 1987), and
a trend towards uniform deformation (Bosnyak et al., 1980). Other work on
these copolymers by Prevorsek and De Bona (1981,1986) has demonstrated
that the replacement of a carbonate by a terephthalate group does in fact
reduce Me and increase lk' Thus the proposition that an increase in lk leads
both to an increase in strain hardening and to the density of entanglement,
as currently estimated, is supported by synthetic studies.

1.4.6 CONCLUSIONS FROM THE DISCUSSION ON ENTANGLEMENTS

Our understanding of the nature and role of entanglements in an organic


glass is still rather limited. In particular, it is not yet clear whether the
measured results relate to localized or general features of the interaction
between polymer chains. In recent studies of crazing and large deformation
processes a role for these interactions has been proposed but there are many
gaps in the available experimental evidence, especially regarding the effect
of temperature changes.

REFERENCES
Aharoni, S.M. (1983) Macromolecules, 16, 1722.
Aharoni, S.M. (1986) Macromolecules, 19, 426.
Allison, S.W. and Ward, I.M. (1967) J. Appl. Phys., 18, 1151.
Andrews, E.H. (1973) in The Physics of Glassy Polymers (ed. R. N. Haward), Applied
Science Publishers, London, Ch. 7.
Andrews, J.M. and Ward, I. M. (1970) J. Mater. Sci., 5, 411.
Argon, A.S. (1973) Phil. Mag., 28, 39.
Arruda, E.M., Boyce, M.e. and Jayachandran, R. (1995) Mech. }\Jaterials, 19, 193.
Berry, J.P. (1971) in Fracture VII (ed. H. Liebowitz), Academic Press, New York.
Birstein, T.M. (1977), Polym. Sci. USSR, A19, 54.
Bosnyak, e., Haward, R.N., Hay, J.N. and Parsons, I.W. (1980) Polymer, 21, 1449.
Bowden, P.B. (1968) Polymer, 9, 449.
Bowden, P.B. (1973) in The Physics of Glassy Polymers (ed. R.N. Haward), Applied
Science Publishers, London, Ch. 5.
Boyer, R.F. (1987) in Order in the Amorphous State of Glassy Polymers (eds S.E.
Keinath, R.L. Miller and J.K. Rieke), Plenum Press, New York, p. 135.
Bubeck, R.A., Smith, P.B. and Bales, S.E. (1987) in Order in the Amorphous State of
Glassy Polymers (eds S.E. Keinath, R.L. Miller and J.K. Rieke), Plenum Press,
New York, p. 347.
Coates, P.D. and Ward, I.M. (1980) J. Mater. Sci., 15,2897.
De Gennes, P.G. (1971) J. Chern. Phys., 55, 572.
Donald, A.M. and Kramer, E.J. (1982) J. Polym. Sci., Polym. Phys. Edn, 20, 899.
Edwards, S.F. and Viglis, Th. (1987) Polymer, 28, 375.
Ferry, J.D. (1980) Viscoelastic Properties of Polymers, 3rd edn, John Wiley, New
York.
30 Introduction

Fitzpatrick, J.R. and Ellis, B. (1973) in The Physics of Glassy Polymers (ed. R.N.
Haward), Applied Science Publishers, London, Ch. 2.
Fletcher, K., Haward, R.N. and Mann, J. (1965) Chern. and Ind., 1854.
Flory, P.J. (1949) J. Chern. Phys., 17, 303.
Folkes, M.J. and Keller, A. (1973) in The Physics of Glassy Polymers (ed. R.N.
Haward), Applied Science Publishers, London, Ch. 10.
Fox, T.G., Gratch, S. and Loshaek, S. (1956) Rheology, Vol. 1, Academic Press, New
York, p. 446.
Gee, G. (1970) Contemp. Phys., 11, 353.
Geil, P.H. (1987) in Order in the Amorphous State of Glassy Polymers (eds S.E.
Keinath, R.L. Miller and 1.K. Reike), Plenum Press, New York, p. 83.
Graessley, w.w. and Edwards, S.F. (1981) Polymer, 22, 1229.
Griffith, A.A. (1921) Phil. Trans. R. Soc. London A, 221, 163.
Haward, R.N. (1949) The Strength of Plastics and Glass, Cleaver-Hume Press,
London, and Interscience, New York, p. 103.
Haward R.N. (1975), in Molecular Behaviour and the Development of Polymeric
Materials (eds A. Ledwith and A.M. North), Chapman & Hall, London, Ch. 12.
Haward, R.N. (1993) Macromolecules, 26,5860.
Haward, R.N. and Thackray, G. (1968) Proc. R Soc. London A, 302, 453.
Hoff, E.A.W. (1952) J. Appl. Chern., 2, 44l.
Kelly, A. and MacMillan, N.H. (1986) Strong Solids, 3rd edn, Clarendon Press,
Oxford.
Kinloch, A.J. and Young, R.J. (1983) Fracture Behaviour of Polymers, Elsevier
Applied Science, London.
Kurata, M. and Tsunashima, T. (1975) Polymer Handbook, 2nd edn, (eds J. Brandrup
and E.H. Immergut), John Wiley, New York, VII/3.
Lovell, P.A., Sherratt, M.N. and Young, R.J. (1996) in Toughened Plastics II: Science
and Engineering (eds c.K. Riew and A.J. Kinloch), Advances in Chemistry Series
252, American Chemical Society, Washington, DC, p. 21l.
McCrum, N.G., Reed, B.E. and Williams, G. (1967) Anelastic and Dielectric Effects
in Polymeric Solids, John Wiley, London.
McKechnie, 1.1., Haward, R.N., Brown, D. and Clarke, J.H.R. (1993) Macro-
molecules, 26, 198.
McKenna, G.B. (1989) in Comprehensive Polymer Science, Vol. 2 (eds C. Booth and
C. Price), Pergamon, Oxford, Ch. 10.
Mitchell, G.R. (1987) in Order in the Amorphous State of Glassy Polymers (eds S.E.
Keinath, R.L. Miller and 1.K. Rieke), Plenum Press, New York, p. l.
Muthukumar, M. and Edwards, S.F. (1989) in Comprehensive Polymer Science, Vol.
2 (eds C. Booth and C. Price), Pergamon, Oxford, Ch. 1.
Onogi, S., Masuda, T. and Kitagawa, K. (1970) Macromolecules, 3, 109.
Pinnock, P.R. and Ward, I.M. (1966) Trans. Faraday Soc., 7, 66.
Plummer, C.J.G. and Donald, A.M. (1990) Macromolecules, 23, 3929.
Prevorsek, D.C. and De Bona, B.T. (1981) J. Macromol. Sci.: Phys., B19, 605.
Prevorsek, D.C. and De Bona, B.T. (1986) J. Macromol. Sci.: Phys., B25, 515.
Price, L., Haward, R.N., and Parsons, I.W. (1979) Polymer, 20, 162.
Rackley, F.A., Turner, H.S., Wall, W.F. and Haward, R.N. (1974) J. Polym. Sci.,
Polym. Phys. Edn., 12, 1355.
Raha, S. and Bowden, P.B. (1972) Polymer, 13, 174.
Rehage, G. and Borchard, W. (1973) in The Physics of Glassy Polymers (ed. R.N.
Haward), Applied Science Publishers, London, Ch. l.
Seitz, 1.T. (1993) J. Appl. Polym. Sci., 49, 1331.
Thomas, E.L. and Roche, E. 1. (1979) Polymer, 20, 1413.
Thomas, E.L. and Roche, E. J. (1981) Polymer, 22, 333.
References 31

Treloar, L.R.G. (1975) The Physics of Rubber Elasticity, 3rd edn, Clarendon Press,
Oxford.
Vincent, P.I. (1972) Polymer, 13, 557.
Ward, I.M. (1987) Developments in Oriented Polymers 2, Elsevier Applied Science,
London.
Wendorff, J.H., (1987) in Order in the Amorphous State of Glassy Polymers, (eds
S.E. Keinath, R.L. Miller and J.K. Rieke), Plenum Press, New York, p. 53.
Williams, J.G. (1973) Stress Analysis of Polymers, Longman, London.
Wignall, G., Ballard, D.G.H. and Schelten, J. (1976) J. Macromol Sci.: Phys., Bt2, 75.
Wool, R.P. (1993), Macromolecules, 26, 1564.
Wu, S. (1989) J. Polym. Sci., Polym. Phys. Edn, 27, 723.
Yeh, G.S.Y. (1972) Crit. Rev. Macromol. Sci., 1, 173.
Young, R.J. (1989) in Comprehensive Polymer Science, Vol. 2 (eds C. Booth and C.
Price), Pergamon, Oxford, Ch. 15.
Young, R.J. and Lovell, P.A. (1991) Introduction to Polymers, 2nd edn, Chapman &
Hall, London.
Young, R.J. (1995) J. Textile Inst., 86, 360.
Zhou, Z., Chudnovsky, A., Bosnyak, c.P. and Sehanobish, K. (1995), Polym. Eng.
Sci., 35, 304.
Molecular dynamics
modelling of
amorphous polymers 2
J.H.R. Clarke

2.1 Introduction
The past decade has seen steadily growing activity in the detailed atomistic
modelling of polymer melts and glasses. These studies have been aimed at
improving our understanding of a variety of physical properties such as
stress-strain behaviour, diffusion of small solute molecules and local chain
motions. Such simulations cannot be classed strictly either as experiment or
theory; rather they are a separate and complementary approach to the
problem, providing molecular level insight which can be used to enhance the
interpretation of experimental data and as a basis for developing new
mathematical relations between physical properties.
It must be emphasized that a great deal has already been learned about
the properties of polymers through the use of analytical theory (de Gennes,
1979; Doi and Edwards, 1986) and, more recently, from computer simula-
tions employing highly simplified (often referred to as coarse grain) models
which focus attention on just the essential features of polymers, such as
connectivity and van der Waals interactions (Binder, 1995). In these cases
the atomistic detail is ignored and a polymer chain is considered as a
connection of freely jointed 'statistical units' or Kuhn segments. The
limitation of this approach is that it does not readily provide a direct
connection between monomer level structure and bulk properties. In this
article we shall concentrate wholly on atomistic-level molecular dynamics
simulations which offer new opportunities for studying properties ab initio.
With a knowledge of interatomic forces together with the assumptions of
classical mechanics it is possible, at least in principle, to give a complete
atomic-level description of a polymer system. The emergence of these
modelling studies is particularly timely in view of very recent developments
in experimental techniques for detailed investigation of polymer dynamics,

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
34 Molecular dynamics modelling of amorphous polymers

such as multidimensional NMR (Spiess, 1991) and inelastic neutron


scattering (Richter, 1992).
Of course even atomistic polymer modelling is not new. Many years ago
Flory proposed a model for the calculation of polymer dimensions purely
from the local intramolecular interactions in a single isolated chain. The
model was founded on the assumption that all long range interactions
within chains are completely screened. In combination with the well known
rotational isomeric states (RIS) model also introduced by Flory, this local
energy approximation is widely used to predict the dimensions of polymer
chains in melts and in theta solvents (Flory, 1988). In contrast, no such
approximation is made for the new simulation techniques under discussion
here; they are applied directly to melts and glasses containing many
interacting chains, they include all 'non-bonded' interactions explicitly and
can be applied to study properties which are determined by the detailed
interplay of entropy effects with intra- and intermolecular forces.
One of the great attractions of computer simulation is the control that
one has over the parameters defining a particular system. Coupled with the
ready availability of detailed information from the simulation, this means
that systematic studies can be devised to show how molecular parameters
are related to bulk properties. Although the applications to polymers are in
their infancy, the increasing power and availability of computer technology
provide a strong driving force. We have much to learn but it will be difficult
to ignore computer-aided molecular design as an important new tool for the
development of polymer materials.
There are thus many exciting prospects for atomistic-level polymer
modeling but with such a powerful technique at our fingertips it is necessary
to be careful and critical in its application. What you get out of a simulation
exercise depends entirely on what you put into it. In this article we shall
limit the discussion to molecular dynamics simulations; the aims are to
outline some of the application areas relevant to glassy polymers where
significant progress has been made and to focus attention on some of the
important technical issues that arise in the use of this technique.
We shall limit the discussion purely to amorphous polymers. While the
phenomenon of partial crystallization is the rule rather than the exception
for laboratory polymers, very much the opposite is true in simulations.
Homogeneous nucleation rates are sufficiently slow that, with the small size
of the samples and with maximum simulation times limited to about 10 - 8 s,
the probability of spontaneous crystallization for all but the shortest
polymer chains is extremely low. Much of the discussion will be focused on
amorphous polyethylene, which as a result of its relative simplicity has
received a great deal of attention in these early years. It is perhaps ironic
that the pure amorphous material is extremely difficult to prepare in the
Ingredients of a computer simulation 35

laboratory, so close comparisons with experiment have been quite limited.


From the simulation point of view, however, it is quite clear that a sound
understanding of the simple systems must be achieved before embarking on
more ambitious studies of more complex polymers.
While the size and time scale of simulations might therefore be deemed
an advantage for the study of pure amorphous materials, these limitations
also lead to some rather fundamental questions. For instance, to what extent
are the properties obtained in simulations relevant to those observed in the
laboratory? The characteristic behaviour of polymer chains covers a wide
range of length scales from the mean square end-to-end distance and the
entanglement length down to those characteristic of a single monomer.
Similarly the range of relevant time scales is extremely broad. Monomer
motions occur on the time scale of picoseconds but the typical relaxation
times of, for instance, topological constraints in real polymers can be of the
order of seconds close to the glass transition (Ferry, 1980). Some of these
problems can be sidestepped by modelling oligomers containing perhaps
20-50 monomers, but one has to be cautious in transferring the interpreta-
tion of results so obtained to the very long polymer chains common in
experiments and most practical applications.

2.2 INGREDIENTS OF A COMPUTER SIMULATION


There are several questions to address before embarking on the modelling
of a polymer. What approach should be taken in choosing a force field for
the intrachain and interchain interactions? Which simulation technique
should be used? How should we prepare the initial state of the polymer? In
this section we shall discuss possible answers to all these questions.

2.2.1 WHAT METHODS ARE AVAILABLE?


We start by considering the choice of simulation technique. Basically, three
quite different methods are available - energy minimization, Monte Carlo
sampling and molecular dynamics simulation. Although most of the atomic
simulation results to date have utilized molecular dynamics, it is neverthe-
less relevant to make some comparative remarks.
In energy minimization it is assumed that the most probable state of the
system is the one with lowest potential energy. The property of interest is
completely determined by the various forces in the system. Highly refined
procedures are used to move monomers either individually or collectively in
order to find an energy minimum. Atomic motions are ignored so the
technique models a polymer at zero kelvin and entropy effects are not
included. At high densities the atomic movements are however severely
36 Molecular dynamics modelling of amorphous polymers

restricted and large scale changes in the chain configuration do not occur.
Indeed it is possible for the system to become 'trapped' in local metastable
states. The method therefore relies heavily on the initial choice of chain
configurations. Energy minimization methods have been used extensively
for detailed studies of the mechanical properties of solid polymer structures
near equilibrium (Theodorou and Suter, 1986; Gusev, Zehnder and Suter,
1995; Gusev, Zehnder and Suter, 1996) and some of the results are discussed
in Chapter 4.
Monte Carlo is a powerful method for sampling polymer configurations
according to Boltzmann statistics at a finite temperature. It has been
particularly successful in investigations of the properties of single chains in
diJferent solvent regimes, using mainly coarse grain models. It has also been
used very successfully in conjunction with lattice models to study, for
instance, phase behaviour in many-chain polymer systems (Binder, 1995). Its
application to the more realistic regime of polymer melts in continuous
space with atomistically detailed polymers has been more problematic. The
main difficulty has been in devising successful artificial methods for explor-
ing the intricate configurational fluctuations of polymers at liquid or glassy
densities. The source of the difficulty is the connectivity of polymers, which
means that moves have to involve several monomers at a time (e.g. by using
a crankshaft move or by rotating part of a chain around a pivot point) and
multiple high-energy overlaps are extremely difficult to avoid. As a result
acceptance rates can be vanishingly small. Considerable progress has been
made using algorithms which regrow monomers or groups of monomers at
the ends of chains (Siepmann and Frenkel, 1992). Even so, the efficiency is
limited and recent direct comparisons suggest a significant disadvantage
compared to molecular dynamics in terms of modelling a melt composed of
fairly short chains at equilibrium (Yong and Clarke, 1996).
The power of molecular dynamics is that it is basically very straightfor-
ward to apply. Assuming the applicability of classical mechanics, it involves
the simultaneous solution of the equations of motion for a sample of
interacting polymer chains at a finite temperature. There are several consist-
ency checks that can be used to monitor the progress and integrity of the
simulation. It is deterministic, the system follows a well defined trajectory in
phase space and it is the only reliable method for examining time-dependent
properties. Although technically more involved since it requires the compu-
tation of forces and not just energies, the advantage over Monte Carlo is
that monomer moves do not have to be devised; they occur spontaneously
and are by nature cooperative.
The application of molecular dynamics to study molecular fluids is well
established and many of the techniques can be transferred in a. straightfor-
ward way to polymer simulations (Allen and Tildesley, 1987). As mentioned
Ingredients of a computer simulation 37

in the previous section, however, the special feature of polymers which


requires careful attention if anything like a realistic simulation is to be
obtained is that the spontaneous configurational fluctuations in polymers
often cover a very wide range of time scales. Since atomistic modelling
techniques are currently limited to times of the order of nanoseconds, there
is a potentially serious problem of ergodicity here; simulations generally
cannot be run long enough to sample fully all the equilibrium fluctuations.
In practical terms, of course, long time fluctuations may not be important
for the property being examined, e.g. guest molecule diffusion in polymer
glasses. In this case we can avoid non-ergodicity problems either by using a
very large system ('" 100000 monomers; Brown et al., 1996) or by averaging
over many smaller but independent samples (Brown et aL, 1994b). Insuffi-
cient averaging of this type has been the shortcoming of many polymer
simulations to date.

2.2.2 MOLECULAR DYNAMICS FOR POLYMERS

Let us start by defining the external conditions. In the simplest case the total
energy E, the volume V and the total number of monomers N are kept
constant and we have a microcanonical or (N, V, E) ensemble. In this case
the atomic motions are Newtonian and the equations to be solved are
Hamiltonian. Whatever method is chosen, the equations of motion are
solved numerically using discrete time steps with the aid of well established
integration algorithms.
In practical simulations it is often much more convenient, however, to
control the pressure rather than the volume and, in particular, to maintain
constant the temperature rather than the energy of the system. These
conditions can be achieved by modifying the Hamiltonian or by using more
general Lagrangian equations of motion. A good deal of work has been put
into developing and testing such 'extended' molecular dynamics methods to
simulate isothermal (N, V, T) and isothermal-isobaric (N, P, T) ensembles
(Allen and Tildesley, 1987). In simple molecular fluids they all give very
similar results (Brown and Clarke, 1984) and most have been used at
various times for polymer simulations.
We shall now outline a method for controlling the pressure P or the
equivalent stress (J developed within a sample (Berendsen et al., 1984). Here
we shall use the convention that (J = - P. The method has proved extremely
useful for modelling polymer melts and glasses (Brown and Clarke, 1991)
and utilizes weak coupling of an externally applied tensorial pressure field,
P*, to the system through a simple feedback loop (Brown and Clarke, 1991).
It is assumed that, provided the coupling is loose enough, it will have an
insignificant effect on the first-order properties of the system. A very similar
38 Molecular dynamics modelling of amorphous polymers

scheme can be used also to control the temperature of a simulation. For the
detailed discussion of this method readers are referred to the original articles
cited above.
The coupling is implemented by allowing a matrix h, made up from the
basis vectors, a, band c which determine the shape of the primary dynamics
cell, to respond to imbalances between the internally measured stress tensor
and an externally applied pressure tensor. The equation for the rate of
change of the h matrix with time is then defined as
• (J + P*
h=--- (2.1)
M

where M is a coupling constant and (J is the internally measured stress


tensor, which in this case is defined in an 'atomic' frame of reference. In
terms of the momenta p, the positions r and the forces f acting at the N
interaction sites

(J = - -1
V i=l
L [1-PiPi
N
mi
+ rJ; ] (2.2)

Although the above equation is formally correct, great care has to be taken
in a simulation using periodic boundaries when calculating (J. Criteria that
can be used for choosing the value of the coupling constant M have been
discussed in detail elsewhere (Brown and Clarke, 1991).
A simple proportional scaling of coordinates is used to minimize local
disturbances. If we define a set of scaled coordinates s, by
(2.3)
differentiation gives the following equation of motion for the sites:

(2.4)

The motion is thus seen to be split into two contributions which are
integrated separately, that due to the momenta and that resulting from the
change in shape and size of the cell. The 'fast' motions due to the momenta
are dealt with in the usual way using a 'leapfrog' algorithm (Allen and
Tildesley, 1987) incorporating an iterative scheme to maintain the con-
straints whereas a simple first-order Taylor expansion is considered suffi-
cient to integrate the equation for the relatively 'slow' motion of the primary
cell

(2.5)
Ingredier of a computer simulation 39

It can then simply be shown that to first order the motion of the primary
cell results in a scaling of the position of a site:
(2.6)
Although the method is less rigorous than alternative techniques that have
been described in the literature (Parrinello and Rahman, 1985), it does have
at least one important practical advantage. This is that the pressure
imbalance is coupled to the first derivative of the basis vectors rather than
to the second derivative, which means that the motions of the primary cell
are overdamped and so there is little tendency for an unphysical oscillatory
response to changes in the applied pressure. For this reason this method
comes into its own for the calculation of non-equilibrium properties of
dense, highly viscoelastic systems.

2.2.3 FORCE FIELDS


As in any molecular level simulation, one of the first decisions to make is
what inter- and intramolecular force field to use. As far as the intrachain
contributions are concerned, realistic modelling of the chain flexibility,
controlled by valence angle and torsional potentials along the chain, is an
important requirement. Quantum mechanical calculations and spectros-
copic data on short chain homologues are the main source of data with
which to fit these potentials.
With regard to the long range force field (the so-called non-bonded
interactions which include interchain effects and also direct interactions
between monomers on the same chain), we have basically two choices. First,
we can set about bringing together as much information as possible from
experiment and quantum mechanical calculations to develop 'good' force
fields and in this way to aim for quantitatively accurate modelling. This
approach has been used, for instance, to model melts of the hydrocarbon
n-C44 H 90 and to make extensive and successful comparisons with a wide
variety of structural and dynamic experimental data (Smith et al., 1994). The
price paid is that such detailed fully atomic simulations are extremely
demanding on computer resources.
It is however possible to make some simplifications and still retain
reasonably realistic simulations. For instance, in one very popular (and
quite successful) model of hydrocarbon chains the hydroge.' atoms on each
of the methylene groups are subsumed into a single interaction centre on the
carbon atom. This so-called united atom model has been used with some
success to model properties of e.g. polyethylene; simulations using this
model have the advantage of consuming about an order of magnitude less
computing time than fully atomic models. The emphasis here is not on a
40 Molecular dynamics modelling of amorphous polymers

~ ~ a

~
. iM
.. ........... valence an Ie poIeniial
...

f e

j·tB
non· bonded poIenlilll CH 1 - unilcd 110m model

- am lropi
Hl uniled 110m model

'rit III - full lomie model

Figure 2.1 A typical scheme for modelling the intra- and interchain interaction
energies in a polymer simulation. As an illustration, three different schemes for
determining the non-bonded interactions between CH 2 groups are shown (all based
on the Lennard-Jones potential form). See text for more details.

highly detailed description of monomer interactions; in fact, what is more


important is the detail that one can leave out and still correctly reproduce
the essential behaviour of the system. This kind of simplification is quite
different to that of the coarse grain models where beads are used to
represent statistical units in a polymer chain.
A typical scheme for modelling these different interactions in polymers are
shown for a linear hydrocarbon chain in Figure 2.1. By way of illustration,
we will now briefly outline four different models that have been used for
amorphous polyethylene in recent simulations. We identify the models as
PEl to PE4 and in all four cases the monomer units are treated as single
interaction sites and given masses corresponding to CH 2 groups.
As far as the intramolecular potentials are concerned, there are small
differences between the models but these are unlikely to have a significant
effect on the calculated values of bulk properties. In models PEl (Brown and
Clarke, 1991) and PE2 (Brown et at., 1994b) neighbouring sites on the chain
are connected together by rigid bonds of length 0.153 nm using the method
of constraints, whereas harmonic springs are employed in models PE3
(Rigby and Roe, 1987) and PE4 (Pant and Boyd, 1993). The use of springs
usually forces a much shorter time step on the simulation unless an
artificially small force constant is used, in which case there is always a risk
Ingredients of a computer simulation 41

of spurious coupling between the bond vibrations and other degrees of


freedom.
Flexibility of the chains is modelled by incorporating a harmonic valence
angle potential, <1>(0), and a torsional potential, <I>(ex), into the model. For
PEl, PE2 and PE3 <1>(0) is of the form
1
<1>(0) = "2ko(cos 0 - cos 00)2 (2.7)

where ko = 520kJmol- 1, 00 = 112.813° for PEl and PE2 and ko =


500kJmol-1, 00 = 120° in the case of PE3. Use of the cosine of the angle
in the harmonic potential is a computational convenience. but it is possible
to express <1>(0) directly in terms of angle displacements as in the case of PE4
which uses ko = 482kJmol- 1 and 00 = 111.6°. For small displacements
there is very little difference in the two potential forms. [n the absence of
precise experimental data the exact choice of ko is in any case somewhat
arbitrary.
The torsional potential restricting internal rotations about a bond in the
chain can be parametrized in terms of the dihedral angle ex using experimen-
tal and ab initio data for short chain alkanes. The form used in PEl for
instance is given below (Steele, 1985):
<I>(ex)jJ mol- 1 = Co + C 1 cos ex + C 2 cos 2 ex + C 3 cos 3 ex (2.8)
where Co = 8832, C 1 = 18087, C 2 = 4880 and C 3 = - 3] 800.
There are much more significant differences between the force fields used
by the four models for interchain interactions and non-bonded interactions
(those between sites separated by at least three others). All use the
Lennard-Jones (LJ) 12-6 potential form

<l>LJ{I riil) = 48 {(I;jIY2 -C;jIY} (2.9)

but the details are different. For instance, in simulations it is common


practice to set the potential energy to zero beyond a certain distance (the
cut-off distance rc) in order to restrict the number of pair interactions that
have to be considered. In model PE2, for instance, this truncation is made
at the minimum of the potential, so there are no attractive interactions at
all. While this may seem a gross simplification, it is widely accepted that the
repulsive part of the potential is primarily responsible for structural effects
in dense fluids (Weeks, Chandler and Anderson, ]971). In PE2 the potential
is also raised by the well depth
<1>(1 riil) = <l>LJ{I rijl) +8 for Irijl ~ 21/6 a (2.10)
<1>(1 rijl) = 0 for Irijl > 21/oa
42 Molecular dynamics modelling of amorphous polymers

so that both the force and the energy go to zero and long range corrections
do not apply. This potential is computationally highly efficient; molecular
dynamics programs run about 3.5 times faster for PE2 than for PEL
For both PEl and PE2 e/kB = 57 K and (1 = 0.428 nm. In the case of PEl,
however, the potential is truncated at re = 2.5 (1 so as to include attractive
forces. To take account of the discarded interactions beyond r e , long range
corrections were made to the potential energy and the pressure at each step
assuming that the pair distribution function g(r) was unity for r > re;
attractive interactions are therefore fully represented in this model. It is of
course computationally less efficient, as mentioned above. Using the PEl
potential simulations gave a reasonable fit to the density of real polyethylene
at 500 K, as extrapolated from experimental data.
In PE3 e/kB = 57 K and (1 = 0.38 nm, and in this case the potential is
truncated and raised to zero at 1.5(1; this is a form previously used in studies
of glass formation in the monatomic Lennard-Jones fluid (Fox and Ander-
son, 1971). With the dominant repulsive forces, the measured values of the
pressure for both PE2 and PE3 are extremely high in molecular dynamics
simulations of the melt. The density has to be adjusted arbitrarily to fit
either the experimental data or other simulation data. In PE4 e/kB = 57 K
and (1 = 0.38 nm, but the interaction centre is offset from the centre of the
CH 2 unit by an amount 0.042 nm along the bisector of the obtuse C-C-C
angle. This is referred to as an anisotropic united atom potential (Toxvaerd,
1990) and it gave an improved fit to the density of molten laboratory
polyethylene over a fairly wide range of conditions (Pant and Boyd, 1993).

2.3 PREPARATION OF MODEL POLYMER MELT SAMPLES


It is relatively straightforward to prepare an equilibrium sample of a simple
molecular liquid from an arbitrary starting configuration (usually crystal-
line) using molecular dynamics or Monte Carlo methods. The reason is that
the time scales of the relaxation processes leading to equilibrium are usually
quite short. This simple procedure does not lend itself favourably to
polymers due, as has been mentioned above, to the long time scale required
for structural relaxation in these materials. The method has been used in
simulations of molten polyethylene, but only by starting the system at a very
high temperature and using highly reduced values for the intermolecular
forces during the initial stages to speed up the equilibrium before cooling to
the required temperature.
Most simulations to date have used a two-stage process for direct
preparation of amorphous samples at the required temperature. The two
stages are firstly chain growth to produce the basic topologies and then a
period of molecular dynamics, sometimes preceded by energy minimization,
Preparation of model polymer melt samples 43

to 'equilibrate' the excluded volume interactions. Again the device is


sometimes used of reducing the magnitude or range of th~~ van der Waals
interactions during the initial period of relaxation. The first stage consumes
a tiny fraction of the total computing time but is extremely important since,
once excluded volume interactions are fully introduced, the time scale for
further topological changes becomes extremely long.
The importance of fully documenting the preparation procedure in
polymer simulations cannot be overemphasized. Any physical property that
relates to the chain dimensions in the melt or glass is expected to exhibit a
marked sample history dependence, a phenomenon well known in experi-
mental polymer science.

2.3.1 INITIAL CHAIN CONFIGURATIONS


The usual approach here is to utilize the local energy approximation
mentioned in section 2.1 (Flory, 1988) to build an initial distribution of
single chain configurations from which a melt sample can be assembled at
the required density. There have been several implementations of this
approach. A particularly straightforward method is to use a Monte Carlo
pivot algorithm to sample chain configurations of a single chain of the
required length and at the desired temperature in a vacuum, ignoring all
long range interactions. Long range here refers to distance along the chain,
not through space (Brown et aI., 1994a). For alkane chains, for instance,
only interactions involving neighbouring torsion angles and backbone sites
separated by four bonds (the so-called pentane effect) are considered. In the
absence of any long range entanglement effects, such a system rapidly
achieves equilibrium. One merely stores samples of chain configurations
from the equilibrium distribution and these are then utilized to build a
starting structure for the melt. The important point is that this can be done
using exactly the same intramolecular force field as for the full molecular
dynamics simulations.

2.3.2 INTRODUCING EXCLUDED VOLUME EFFECTS


The required number of chains are introduced into the basic simulation cell
at the required density. An unavoidable side effect of this procedure is that,
at a typical melt density, there are bound to be a large number of overlaps
between sites for this initial configuration. In principle, energy minimization
could be used to remove the high energy contacts, but molecular dynamics
allows relaxation at a specified temperature. It is necessary, however, to
moderate the forces in the very early stages to avoid breakdown of the
molecular dynamics algorithm.
44 Molecular dynamics modelling of amorphous polymers

One method that has proved quite robust uses a 'truncated force'
potential (McKechnie, Brown and Clarke, 1992). In the initial stages of the
simulation the short range force for neighbours i and j, where li-jl > 5, is
constrained to be constant below a critical separation rtr' i.e.

-d<l>m(r)
dr
=F (= -d<l>(r
trdr
tr ))
(2.11)

The full definition of the resulting modified potential is then


<l>m(r) = <I>(r) for r> rtr (2.12)
<l>m(r) = <I>(r tr ) + (r tr - r)Ftr for r ~ rtr (2.13)
rtr must be sufficiently small that only a few pairs will be within this distance
in the equilibrium distribution (which at this stage is unknown), but not so
small that the large magnitude of Ftr causes breakdown of the algorithm.
The truncated potential is applied only to the 'long range' interactions
and not to the local interactions (section 2.3.1). The procedure is to decrease
continuously the value of rtr from 0.90" to 0.70" in an initial run of time length
tv using a time step of 1 fs and rescaling of particle momenta at each step to
remove the large amounts of thermal energy released. At this stage the
switch is made from the modified potential to the full potential. Satisfactory
results are obtained using PEl with tv = 3 ps. During this initial stage it is
easiest to carry out the MD simulations at constant volume with the
temperature kept close to 500 K using the loose coupling method with a
coupling constant tT ~ 0.1 ps. Subsequent relaxation can be carried out
either under controlled pressure or constant volume.

2.3.3 SAMPLE RELAXATION


One unfortunate consequence of the introduction of excluded volume effects
is some unavoidable perturbation of the carefully prepared chain configur-
ations! Dependent on the chain length, the sample can take an extremely
long time to relax back to equilibrium. The effect is shown for alkane-like
chains with 100 methylene groups in Figure 2.2a, where the radius of
gyration relative to the initial value (corresponding to the Flory model) is
plotted as a function of time for a melt sample of 640 chains at 500 K. These
simulations were carried out at constant volume but can equally well be
performed under controlled pressure conditions.
There are two striking features of this plot. First, the introduction of
excluded volume causes an immediate and substantial decrease in the radius
of gyration (the chain configurations were initially equilibrated using pivot
Monte Carlo so, within the statistical error of about 1%, the t = 0 value of
Preparation of model polymer melt samples 45

~
g 1.00 I---------N-....!-~I..l_,M_I
~
A-
~ 0.98
v
. . . . . 0.96
A
C\I
(/)
V 0.94

0.92 0~---L-..",20,L,0..,..0---JL...-....,.40...LO-0-.l...--6..JOO--O---J
4650 r----,----,----r--~-~__,~~

(b)
4600
20 bars pressure
..... discrepancy at 1 ns
CIS
~ 4550
a..

time / ps
Figure 2.2 The relaxation of (a) the square radius of gyration, 52, and (b) the
hydrostatic pressure at 500 K following the introduction of excluded volume in a
sample of a melt composed of 640 chains each with 100 monomers of model PE2
polyethylene (see text). Note the expanded vertical scales. The simulations were at
constant volume; 20 bar pressure discrepancy would be equivalent to a density
discrepancy at constant pressure of about 0.1 %. 52 values are shown relative to
those determined assuming complete screening of long range interactions (see text).
The density is 0.70 gcm- 3 . These results were obtained with a Fujitsu AP1000
massively parallel computer.

the ordinate was unity). Second, the relaxation back to equilibrium values
is (as might be predicted) extremely slow, even for these chains of only 100
monomers.
Of course, not all properties are sensitive to small changes in the
chain dimensions. In Figure 2.2b, for instance, we show the relaxation
of the pressure. If the simulations had been performed at constant press-
ure there would have been a corresponding relaxation of the density.
From estimates of the compressibility the calculated density discrepancy
{p( 00 )-(p(t)}/p(inf)} would be only about 0.1 % after 1 ns for this model
46 Molecular dynamics modelling of amorphous polymers

polymer with n = 100 (where n is the number of sites per chain). Since the
density is by far the most important property in determining local chain
motions and penetrant diffusion, for instance, equilibrium times of 0.5-1 ns
may still be adequate for many purposes.

2.3.4 SAMPLE SIZE EFFECTS


We briefly consider here the question of choosing a sample size that will
provide a good representation of bulk behaviour. The traditional require-
ment is that the size of the primary simulation cell is large enough to prevent
molecules interacting with images of themselves through the periodic
boundaries and is also large enough to contain all the important character-
istic structural fluctuations of the bulk system. In the case of molecular
liquids it is usually satisfactory to set the number of molecules N in the
range 100-1000. For a polymer system with a degree of polymerization
n = 1000, this would imply using up to 106 monomers in a simulation. We
could relax the criterion somewhat so that, for instance, the cell dimensions
are greater than the expected mean square end-to-end distance of a polymer
molecule. For n = 1000 we would still have to include about 30 chains and
a total of at least 30000 monomers in the simulation.
A more radical, although controversial, approach is to use a cell which is
only larger than the correlation lengths important to the phenomenon being
studied. In this case we might for instance use just one chain of 1000

Figure 2.3 Two-dimensional schematic diagram of the polymer model consisting of


a single chain replicated by the periodic boundaries. In three dimensions it corre-
sponds to a monodisperse polymer entangled with replicas of itself.
Preparation of model polymer melt samples 47

monomers to form a dense amorphous polymeric system through the


replicative properties of periodic boundaries. The primary chain spans many
neighbouring cells. The model is therefore one of a monodisperse polymer
entangled with replicas of itself. A two-dimensional schematic diagram of
this model is shown in Figure 2.3. Chain ends can be eliminated from the
model by arranging that the end of a chain is attached to the other end of
one of its periodic images, giving 'infinite' length (Weber and Helfand, 1979).
It must be remembered, however, that the repeat length is still the number
of monomers in the primary cell.
The effects of boundary conditions in this kind of model have yet to be
evaluated fully but we can expect that one important condition might be the
size of the unit cell in relation to the correlation length along the chain. For
small values of n there is no doubt that the model gives a poor representa-
tion of bulk behaviour, particularly for less flexible polymer chains, but as
n becomes larger we expect it to be an increasingly better approximation to
a dense amorphous system.

2.3.5 VALIDITY OF THE LOCAL ENERGY APPROXIMATION

As explained in section 2.3.1, this is an important assumption which


underpins most preparation methods for polymer simulation samples. It has
been widely used in conjunction with the rotational isomeric states theory
to predict chain configurations, and the results appear to be in accord with
a considerable body of experimental evidence (Flory, 1985). Computer
simulation provides new opportunities of making unambiguous compari-
sons between the dimensions of melt chains with those predicted using the
local energy approximation. Such comparisons are independent of any
assumptions concerning the detail of intrachain or interchain interactions.
As already indicated in Figure 2.2, accurate studies of the dimensions of
alkane-like chains with up to 100 methylene groups show that the radius of
gyration in the melt at 500 K with full interactions (obtained by molecular
dynamics) differs, at equilibrium, by only 1-2% from the predicted values.
This vindication of the model cannot, however, automatically be extended
to chains of different chemical composition. Although recent comparisons
show that a similar prescription for the local interactions proved satisfactory
in the case of isotactic poly(vinyl chloride) (Neyertz, Brown and Clarke,
1996), discrepancies of '" 30% have been found for poly(ethylene oxide)
(PEO), with the Flory model predicting more compact configurations than
actually found in the melt (Neyertz and Brown, 1995). The models used for
both PVC and PEO were fully atomistic with all atoms carrying partial
charges. The discrepancy in the case of PEO was found to be linked to the
strong preference for gauche configurations about C-C bonds and was
48 Molecular dynamics modelling of amorphous polymers

traced to the competition between attractive intramolecular and inter-


molecular C-H .... 0 interactions.
One may ask why this discrepancy was not observed in previous com-
parisons between experimental data and the prediction of RIS studies
(Smith, Yoon and Jaffe, 1993). The answer may be that such studies require
independent parametrization of the intramolecular potential so that some of
this agreement may arise from the empirical fitting procedures. A conclusion
to be drawn from the simulation studies is that there are limitations in the
application of the approximation due to difficulties in finding the correct
prescription for the required local energy.

2.4 CHARACTERIZATION OF CHAIN DYNAMICS IN DENSE


POLYMERS
In this section we discuss determination of a selection of dynamical
properties of polymer chains on different length scales. Analysis of fluctu-
ations of the end-to-end distance, of local chain conformations, of bond
orientations and of the density provide a fundamental characterization of
polymer melts and glasses. In addition, such motions are in principle
accessible to experimental measurement.

2.4.1 CONFIGURATIONAL FLUCTUATIONS


Chain configuration relaxation is conveniently described by the normalized
correlation function for the square end-to-end distance:

(2.14)

where for a linear polymer of n units


R2(t) = (r 1(t)-r n (t))2 (2.15)
For example, correlation functions obtained from polymethylene melts
composed of n = 50 and n = 100 chains at 500 K using the united atom
model are shown in Figure 2.4 (Brown et al., 1996). In the case of n = 100
a comparison is shown for a small system of 10 chains and a very large
system of 640 chains. Despite the limited range for the comparison, it
appears that there is no discernible system size dependence.
The correlation functions do not exactly fit a simple exponential form, but
if we can define a relaxation time tR as

tR = too CR(t) dt (2.16)


Characterization of chain dynamics in dense polymers 49

1.0

,,
\
----
.....
U==
\,
,,
\,
,
\,. - - n=l00, N=640
...\, • n=l00, N=10
'.\I' ------n=50, N=20

0.1
'"
0 0.5 1 1.5 2 2.5 3 3.5 4
time / ns
Figure 2.4 Correlation functions for the mean square end-to-end distance from
large scale simulations of alkane-like melts at 500 K. The samples contained N chains
each containing n methylene groups using the united atom model.

't"Rcan be interpreted as the average time taken for chain configurations to


be refreshed by the thermal motions within the melt. The values obtained
for 't"R are 1.94 ns and 0.36 ns for n = 100 and n = 50 respectively, and imply
an increase in relaxation times scaling greater than n2 . For n = 1000, which
is the lower limit of what experimentalists might consider should be the size
of a polymer chain, on this basis the relaxation time would be in excess of
'" 0.2 f.!s. In fact entanglement effects would probably lengthen the time even
further.

2.4.2 CONFORMATIONAL MOTIONS


The dynamics of local conformational changes in polymers is of fundamen-
tal interest and has attracted a good deal of attention in molecular dynamics
studies. Most of the studies have been concerned with hydrocarbon chains
where torsional motions around backbone C-C bonds gives rise to the
characteristic gauche +, gauche - and trans conformations.
One useful way to characterize conformational relaxation is to compute
relaxation functions for, say, the trans state of a particular bond (Brown and
Clarke, 1990). This is a particularly straightforward method of analysing
data on the fluctuating conformational states in an equilibrium simulation.
It was originally devised to analyse data on liquid butane but can equally
well be applied to longer chains.
50 Molecular dynamics modelling of amorphous polymers

One computes relaxation functions of the form


(2.17)
where HT [oc;j(t)] is the characteristic function of the trans state and oc;/t) is
the value of dihedral angle j of molecule i at time t. HT[ocit)] only takes
two values,
(2.18)
otherwise
(2.19)
The initial value of the relaxation function, RTT(O), is just the mean fraction
of angles in the trans conformation, <XT ), and at long times RTT(t) will tend
to <XT)2.
The normalized form of the relaxation function

C (t) = RTT(t) - <XT )2 (2.20)


TT <XT )-<XT)2
is shown in Figure 2.5 for methylene chains using the united atom model
for n = 8, 20 and 100 at 500 K and a typical melt density (Brown et aI.,
1994b). The relaxation time for n = 100 is '" 4ps. The functions are,

----

0.1
o 2 4 6 12
tips
Figure 2.5 Normalized relaxation functions for trans states in alkane-like melts at
500 K. Results are shown for chain lengths of n methylene groups using the united
atom model. The corresponding relaxation times are 3.3 ps for n = 8, 3.5 ps for
n = 20 and 4.5ps for n = 100.
Characterization of chain dynamics in dense polymers 51

however, clearly not exponential and cannot be explained using a simple


first-order mechanism such as was successful in the case of butane.
For topologically constrained chains, such as those found in a melt or
glass, it is obvious that interconversions between different states of one angle
will depend strongly on the states of at least nearest neighbour angles. This
effect has been investigated in detail by several groups for linear hydrocar-
bon chains (Boyd et ai., 1994; Neelov and Clarke, 1994; Smith, Yoon and
Jaffe, 1995). Correlations involving gauche+ --+ gauche- transitions be-
tween second nearest neighbours propagate more or less randomly through
the chains in a melt. Self and fourth neighbour correlations also occur quite
frequently while transitions involving nearest or third neighbours are very
infreq uen t.

2.4.3 BOND ORIENTATIONAL MOTIONS


Fast motions in amorphous polymers, such as the reorientation of main
chain bonds and pendant side groups, can in principle be probed by a
number of spectroscopic techniques such as infrared (Williams, 1979), NMR
(Spiess, 1991) and inelastic neutron scattering (Richter, 1992). Since molecu-
lar dynamics is particularly suitable for studying short time dynamics there
are opportunities for making useful comparisons with experimental data
and for evaluating theoretical models.
Most experimental techniques do not probe reorientational motions
directly but measure the time dependence of properties which themselves are
sensitive to orientation (Clarke, 1978). Thus in far infrared and Raman
spectroscopy one measures respectively dipole and polarizability fluctu-
ations with respect to particular normal modes of vibration. Although the
interpretation can be complicated, in cases where the fluctuations arise
predominantly from rotational motions of localized groups, correlation
functions of the first and second Legendre polynomials, C 1(t) and C 2 (t)
respectively, can be retrieved from the data.
Simulation studies are not limited in the above way and one can study
the reorientational motions directly and in great detail by determining the
full probability distribution function wee, t) for bond reorientation through
an angle e in a time t. The reorientational correlation functions C 1(t) and
C2 (t) can of course both be calculated from wee, t), e.g.

(2.21)

Various models have been proposed for reorientational motion in fluids,


giving rise to characteristic forms of C 1(t) and C2 (t). One of the simplest is
52 Molecular dynamics modelling of amorphous polymers

small step rotational diffusion which requires that both C1(t) and CZ{t)
should show exponential decays and that the ratio of the respective
correlation times should be 3 (Clarke, 1978). These predictions are for
isotropic motion, but in polymers the motion of a single bond is likely to
be highly anisotropic as a result of restriction arising from the connectivity
of the chains; this has previously been observed in Brownian dynamics
simulations of a polyethylene model (Weber and Helfand, 1983). Such effects
are likely to lead to a long time tail to the functions.
Neither C1(t) nor C 2 (t) for the backbone atoms in model amorphous
polyethylene shows a simple exponential relaxation (Roe, Rigby and
Furuya, 1992); there is a rapid initial decay over 10-20ps followed by a
much slower decay up to 100-200ps. The correlation functions could be
fitted to the Kohlrausch-Williams-Watts empirical equation
C;(t) = exp( - t/'rY (2.22)
This form for the correlation function fits a wide variety of data on dynamic
properties in complex fluids and, as in the case of many other properties, p
turns out to have a value close to 0.5. Equation 2.21 is often interpreted in
terms of a distribution of relaxation times, but in the case of model
polyethylene it has been shown, however, that the results cannot be
explained in this way. Instead the result is interpreted in terms of a model
for anisotropic motion in which the polymer chain is confined to a 'pipe'
formed by its neighbours (Takeuchi and Roe, 1991).

2.5 STUDIES OF THE GLASS TRANSFORMATION


The glass transformation temperature ~ is one of the most important
material properties of an amorphous polymer and it is not surprising that
glass formation has attracted a good deal of attention in simulation studies.
The accurate prediction of laboratory values of ~ is an attractive goal for
any modelling technique. Most of the molecular dynamics studies to date
have, following the usual experimental procedure, utilized cooling to pro-
duce glassy samples. One of the central issues to address, however, is the
effect of the enormous cooling rate used in the simulations (of the order of
10 10 Ks- 1). Under conditions where a material appears glassy in a simula-
tion, it might well behave as a liquid in the laboratory.

2.5.1 WHAT SHOULD WE EXPECT?


Cooling experiments in simulations can be performed either by reducing the
control temperature in a series of discrete steps, as shown schematically in
Figure 2.6, or alternatively by imposing a continuous decrease. In either case
Studies of the glass transformation 53

cooling SCHEMATIC
schedule
loss of equilibrium
during
stepwise cooling

time
r-

--~~~\J'
equilibrium not achieved
in step time

~ ---~~~
.'.::
I ~

- - - - -------------- -~ -~-~-~-~-~-~\
time temperature
Figure 2.6 Schematic illustration of an idealized stepwise cooling experiment,
showing the increasingly slower density relaxation as the temperature is decreased.
The shorter the step times, the higher the temperature at which the density falls out
of equilibrium and a glass transition is observed.

it is necessary to perform a series of constant temperature simulations in


order to allow relaxation, and subsequently to accumulate by time averag-
ing reasonably accurate values of properties such as the density. The current
practical limit for these simulations is - 1 ns.
At each temperature the rate of relaxation of the density will be controlled
mainly by a combination of the non-bonded interactions and the torsional
motions of the chains. We assume here that the chains are of sufficient
length that purely diffusion motions are frozen out by the topological
constraints. The motions become more restricted as a result of cooling or
compression and we reach a stage where full relaxation of the density is not
complete before the next step in temperature or pressure, so the material
falls progressively out of equilibrium and a glass transition is observed, as
shown in Figure 2.6. The faster the cooling rate, the higher is the tempera-
ture at which the material falls out of equilibrium (Turnbull, 1969).
54 Molecular dynamics modelling of amorphous polymers

By fitting a range of experimental data on real polymers using the empirical


Williams-Landel-Ferry equation, the magnitude of this effect has been
estimated from experimental data as about 3 K per decade increase in
cooling rate. If the difference in time scales between simulation and typical
laboratory measurements is taken as ten orders of magnitude, then on this
basis one might expect a discrepancy of about 30 K between experimental
and simulation data. Use of a simple linear prediction over such a wide scale
must, however, be subject to considerable uncertainty.
Experience with glass formation in atomic and molecular systems (Angell,
1988) suggests that there is an additional effect of using ultra-short time
measurements to characterize the glass transition (in the laboratory this
corresponds to the use of high frequency probes). It appears to result in a
diffuse rather than sharp transformation on the cooling curve. This kind of
behaviour has been observed in simulations of glass formation for the
atomic Lennard-Jones fluid (Angell, Clarke and Woodcock, 1981) using
cooling rates of the order of 10 12 K s - 1. It has also been observed for real
materials in the laboratory when using 'high frequency' probes of the
glass transition. For instance, Brillouin scattering measurements have been
used to probe (Angell and Torrell, 1983) the sound velocity at frequencies
of about 10 GHz in cooling experiments on the glass-forming liquid
2Ca(N0 3)2' 3KN0 3 . The high frequency longitudinal compliance and the
adiabatic compressibility derived from the measurements both show a
smearing out of the transition towards high temperatures whilst the more
familiar static measurements show a sharp transition at the low end of the
transformation range.
This smearing out of the transformation in simulations can be understood
from the temperature dependence of the relaxation times (Angell, 1988) as
shown, again schematically, in Figure 2.7. Simple linear polymers are
examples of 'fragile' fluids which are expected to show a marked tempera-
ture dependence of the activation energies (Ea) for structural relaxation
processes such as conformational changes. At low temperatures in the
supercooled regime, Ea becomes extremely large and relaxation times may
change by one or two orders of magnitude over a few degrees, thus
producing a very sharp transition. At high temperatures a much smaller
value of Ea means that the same change occurs over a much wider
temperature range, thus producing a broad transformation.
Finally, it is worth mentioning yet another reason why we might expect
a different character to the glass transition as observed on very short time
scales. The chain configurational motions that provide the primary mech-
anism for structural relaxation in a supercooled polymer are expected to
have a high degree of cooperative character. For instance, Rouse-like modes
appear to be a reasonable description of intrachain motion up to the
Studies of the glass transformation 55

lab
time scale

simulation
timescale

broad sharp
transformation tran ition
Figure 2.7 Schematic temperature variation of a characteristic structural relaxation
time in a supercooled polymer melt. It is assumed that a change of approximately
ten times in the relaxation time is required to give a significant change in some
associated property such as diffusion coefficient or expansivity. This change takes
place over a much wider range of temperature on the simulation time scale.

entanglement length in dense melts (Doi and Edwards, 1986) and it is likely
that these motions have much to do with viscoelastic and mechanical
properties. Each of the p Rouse modes will have a characteristic relaxation
time Tp and temperature dependence. They can only be 'active' in the glass
transformation if the cooling rate is much less than l / Tp • On the time scale
of current molecular dynamics experiments we can expect that only the
short wavelength modes will be accessible.

2.5.2 STUDIES OF AMORPHOUS POLYETHYLENE


The results of three independent cooling experiments on amorphous poly-
ethylene (PE) models are shown in Figure 2.8. The glaring differences
between the sets of results illustrate the sensitivity of the density- tempera-
ture relationship to the exact conditions of the model and the simulations.
For instance, the three sets of data use three different models for the
non-bonded methylene interactions as detailed in section 2.2.3. In the case
of PEl the Lennard-lones potential parameters were fitted to data for real
polyethylene at only one temperature (500 K), whereas model PE4 was fitted
over a wide range of temperatures.
56 Molecular dynamics modelling of amorphous polymers

0.95 ......
:::---.--.--.--.--,..--.----,
....
........ model
'" 0.90
'" / PE4
'E .... ,)I-
() 0.85 .., laboratory
C> .. , polyethylene

-
-- 0.80
>.
.~ 0.75
Q)
"'0 0.70

100 200 300 400 500 600 700


T/K
Figure 2.8 Cooling curves for three models of polyethylene as obtained by molecu-
lar dynamics simulation. Note the widely different behaviour of the density and the
glass transformation for the three models (see text for details).

For PEl five independent samples of a single chain with 1000 methylene
sites were relaxed for 500 ps at 500 K at an applied isotropic pressure of
1 bar. Samples at different temperatures were then obtained by cooling or
heating at an effective rate of ,. . ., 1 K ps - 1 to the desired temperature under
isotropic controlled pressure conditions (1 bar) followed by subsequent
periods of relaxation of order 1 ns. Using this procedure additional samples
were generated at 600, 400, 300, 200, 100 and 10 K.
As the temperature is lowered there is a gradual decrease in the thermal
expansivity of the polymer (obtained from the slope of the plot) towards
values typical of amorphous solids. For this model the transition is, as
expected, rather weak and smeared out and occurs at a higher temperature
than estimated from experiments (Brandrup and Immergut, 1989). For the
other two models the cooling schedule was not fully specified, but it is
apparent that not only is a transition observed at much lower temperatures,
but surprisingly for these very high cooling rates it appears quite sharp.
Direct observation of the freezing out of conformational transitions at a
temperature of about 400 K provides strong supporting evidence of a
glass-like transformation, in accord with the notion that the torsional
degrees of freedom are the dominant modes of relaxation in these systems
(Brown and Clarke, 1991). This is shown in Figure 2.9 which gives the
temperature dependence of the fraction of trans conformers averaged over
five samples of unperturbed chains at the various temperatures. A dihedral
angle is defined to be trans if it lies between ± 60° otherwise the angle is in
one of the two gauche states. The solid line in the plot is an extrapolation
Studies of the glass transformation 57

90
Initial samples
en prepared at 500 K
c:
~ 80
I-
cfl. t
"".""""""'"
70 conformations
"frozen In"

6OL.-"--L..-"--L..-"--L..-"--.L-...........l
o 200 400 600 800 1000
T/K
Figure 2.9 Variation with temperature of the percentage of trans conformers in the
cooling simulation of polyethylene model PEl. The theoretical equilibrium curve was
calculated according to equation (2.23).

to lower temperatures made on the basis of fitting the data at 500 K and
1000 K to the form

(X T ) = Aex (~<I» (2.23)


(XG ) p RT

where X denotes the fraction of trans (T) or gauche (G) states and A and
~<I> are the adjustable parameters. The actual values used for the curve
shown were A = 1.01 and ~<I>/R = 463 K. The value of ~<I> is about 15% less
than the energy difference between the gauche and trans wells for the
dihedral angle potential; this is probably the result of non-bonded interac-
tions.
It is worth mentioning here that polymer samples can be prepared directly
in the glassy state (without cooling from the melt) by using the preparation
techniques described in section 2.3 at the appropriate temperature. This
method has been used for modelling polypropylene (Theodorou and Suter,
1986) and polyethylene (McKechnie et ai., 1993). In the latter case careful
control over the preparation procedure was used to construct amorphous
samples of the same polymer in different configurational states in order to
examine the relation between mechanical properties and polymer structure
(these simulations will be discussed in more detail in section 2.6.2). Such
control is only possible, of course, for non-equilibrium states where the
chain topology is frozen in.
58 Molecular dynamics modelling of amorphous polymers

0.84

'?E 0.80
o

...
C)

-
heating from
;, 0.76 directly prepared
·w
c::
glass
•••••
Q)
"0 0.72 cooling ~"
from melt -~- '.

100 200 300 400 500 600 700


T/K
Figure 2.10 Comparison of the cooling curve starting from the 500 K melt with the
heating curve for samples prepared directly in the glassy state at 200 K, as obtained
in simulations of polyethylene model PE1.

The transformation to equilibrium liquid behaviour from such samples


can be observed by heating (Clarke and McKechnie, 1996) and some results
for the model PEl are shown in Figure 2.10. In this case the directly
prepared glass was obtained from chains grown at 200 K and these were
significantly more extended than those grown at 500 K. There is a noticeable
difference in the two values at 200 K which may be related to the different
relaxational histories; cooling from the melt over a period of '" 3 ns allows
monomers to achieve much more efficient packing and hence to achieve a
slightly higher density in the glass.
As the temperature of the directly prepared glass is raised at a rate of
1 K ps -1 (the same as that used in the cooling simulation of PEl shown in
Figure 2.10, but in smaller steps), increased thermal motions begin to allow
more relaxation in the directly prepared glass and the density discrepancy
between the two data sets gradually decreases. The change in slope that
occurs in the temperature range 350-400 K is indicative of a further increase
of monomer motions that would be expected as the glass transforms to a
liquid, and it is significant that this transformation occurs in the same
temperature and density range as in the cooling simulations.
Some general remarks are appropriate to conclude this section. The
sensitivity of the transition temperature to the details of the non-bonded
interactions suggests that the physical effect being observed here is equi-
valent to the oc-transition characteristic of amorphous laboratory poly-
mers. The transition temperature does appear to depend, however, on the
coupling of interchain motions with local conformational fluctuations; it has
Studies of the glass transformation 59

been shown in studies of alkane-like systems (Takeuchi and Roe, 1991) that
removal of the torsional potential caused significant lowering of ~. By way
of contrast, the result shown in Figure 2.10 implies that the transition (in
contrast to the stress-strain properties discussed in section 2.6) is not very
sensitive to the overall chain dimensions; the chains prepared directly in the
glass were over three times more extended than those cooled from the melt.
Once the structure has been arrested, there is no evidence of intrachain
processes that might contribute to fj-relaxation phenomena, as seen for
instance in dielectric relaxation experiments. The time scale of these experi-
ments is, however, several orders of magnitude longer and in any case the
relaxation is most often associated with side chain motions; structural arrest
in such polymers has not yet been examined in detail by simulation.

2.5.3 COMPARISONS BETWEEN EXPERIMENT AND SIMULAnON


Several results on different polymers are now available from ultrafast
cooling simulations; some of these data are plotted in Figure 2.11, which
compares experimental and simulation estimates of the glass transforma-
tion temperature (Han, Gee and Boyd, 1994). Only linear hydrocarbon
chains are included here, but nevertheless the comparisons seem reasonably
encouraging. As expected, generally higher temperatures are observed in the

400
/

at-P~ G'
350
/
/
300 /

g at-P~
/

°250 /
~ PIB
f..-OO (I) / ...... PE
200 / B
PBD~ /
150 /
/
/
l00y~~~~~~~~~~~~~~
100 150 200 250 300 350 400
Tg exp (K)

Figure 2.11 Comparison of glass transition temperatures obtained from experiment


and from molecular dynamics simulations. Displacements from the broken line show
the errors in the simulation results. The points marked f3 and y refer to experimentally
observed 'transitions' associated with amorphous regions in semicrystalline PE. See
text for details. (Reproduced from Han, Gee and Boyd, 1994, with permission.)
60 Molecular dynamics modelling of amorphous polymers

simulation but the discrepancies appear to be much smaller than estimated


from the discussion in section 2.5.1; the largest (for polyisobutylene) is only
about 30K.
In regard to the use of simulations to predict transition temperatures, it
should be noted that the results represented in Figure 2.11 were obtained
only for polymers where the parametrization could be adjusted to fit
experimental data on the polymer melts over a range of temperatures. We
should exercise a great deal of caution in attempting to predict transform-
ation temperatures in cases where there are few experimental data. In
addition we should always expect cooling rate effects as discussed in section
2.5.1. Finally, there is a chain length effect which so far has not been
mentioned.
It is known tht ~ increases to an asymptotic limit as the degree of
polymerization N increases to very high values (~20000 or more) and most
of the experimental data apply to this regime. But all of the simulations
discussed here have been performed on chains with N < 1000 and some
have used chains as short as N = 125; this is a regime where we might expect
a significant dependence of relaxation behaviour (and hence transformation
temperatures) on chain length and the effect will be to reduce the transform-
ation temperature as compared to very long chains. The favourable com-
parisons shown in Figure 2.11 may therefore be to some extent fortuitous.

2.6 STRESS-STRAIN PROPERTIES

2.6.1 UNIAXIAL TENSION SIMULATIONS


Mention has already been made that one of the advantages of the controlled
pressure molecular dynamics discussed in section 2.2 is that the form of the
applied pressure tensor p* can be used to impart strain to a sample as a
function of time in much the same way as in laboratory experiments.
Control of appropriate components of the applied pressure tensor can be
used to produce, to take just three examples, uniaxial tension, compression
or shear, as illustrated in Figure 2.12.
Current limitations on simulation times have meant that very high rates
of strain must be used in order to observe the system response. Nevertheless,
the general form of the results shows striking similarities to that obtained in
laboratory experiments performed on time scales many orders of magnitude
slower.
To give an example of what can be achieved in such simulations, we
discuss below the stress-strain behaviour as observed in simulations of
amorphous polyethylene using model PEl at a range of temperatures in the
glass and melt (Brown and Clarke, 1991). The sample size was 1000
monomers formed into a single linear chain as described in section 2.3. The
Stress- strain properties 61

~
":/::::====1)"1
'
~

""""atlon

[(J
- 1 . - 1_ _- "_ '

-I" 0 0

P:.~~~M
+p' O 0
pO. 0 1 0
o 0 1

.Mar f==\\
-lP' O~
pO. P' l 0
001

Figure 2.12 Examples of mechanical deformations that can be induced using loose
coupling molecular dynamics with an applied pressure tensor po, (see section 2.3 for
method).

cooling curve for these samples is shown in Figure 2.8. In order to obtain a
representative picture, the results were averaged over five independent
samples.
The prepared samples were each subjected to a gradually increasing
uniaxial tension by changing the y component of the applied pressure
tensor, P;y, at a constant rate
dP;y .
- - = -r (2.24)
dt
where the tension application rate i used was either 0.5 MPa ps - 1 or
0.1 MPa ps - 1, and the minus sign accounts for the fact that tension is a

negative pressure. Employing two different values for i should give some
indication of the extent to which the measured properties are rate depend-
ent.
The applied tension is best considered as a control variable which
produces a change in the strain. The response is given by the measured
tension, i.e. G yy ' within the sample, calculated using equation (2.2). The
method is preferable to direct control of the strain since there is no way
a priori of predicting how the shape or density of such a small sample will
respond to a change in the external conditions.
The simulations were continued until the sample had extended by
50- 100% of its original length. Extensions beyond about 100% were not
62 Molecular dynamics modelling of amorphous polymers

possible due to the contraction in the transverse direction. For the size of
sample used, the latter effect leads eventually to violation of the criterion
that the truncation diameter for the site-site potential should be less than
the smallest dimension of the primary cell. The important information which
results from these tension experiments is the response of the h matrix,
defining the size and shape of the cell, and that of the measured stress tensor
(1. These together allow us to elucidate the stress vs. strain behaviour. If L

is the length of the primary cell, then the elongational strain is defined as
L- Lo
YL = - - (2.25)
Lo
The average response of the five independent samples at six temperatures
from 10 to 500 K are shown in Figure 2.13 for a tension application rate of
0.5 MPa ps - 1. There is clearly a wide range of behaviour observable in the
model system; at low temperatures the material can support the tension up
to strains of -- 20% before undergoing yield and at progressively higher
temperatures there is a gradual change in behaviour until at 500 K it is
unclear whether there is any elastic response at all.
The essential difference in character between the elastic low-temperature
behaviour and the viscous high-temperature response is shown by plotting
the extension as a function of time (Figure 2.14). On a log-log scale an
elastic response should have an asymptotic slope of 1 at low strains for a

300
-0.........,
250 "'"-c ~ 10K
~---a
100K
~
a.. 200
200 K
~
....... 150
en
en
....
Q)
...... 100 400
C/)

50 500 K

0
0 0.2 0.4 0.6 0.8 1.0
strain
Figure 2.13 The measured stress (rJ yy ) as a function of elongational strain for an
applied tension of 0.5 MPa ps -1 for model amorphous polyethylene. The force field
is PE 1 (see text) and there is a single chain of 1000 methylene groups. The data at
each temperature represent the average behaviour over five independent samples.
Stress-strain properties 63

o 10 K
I] 100 K
6 200 K
c • 300 K
o • 400 K
·00 II 500 K
a510
+-'
x
W
~
o

lr----+--+-~~~~----~~~~~~

lO 100 1000
tips
Figure 2.14 The percentage extension as a function of time for the amorphous
polyethylene samples subjected to a tension application rate of 0.5 MPa ps -'. On the
log-log plot a slope of 1 indicates an elastic response to the applied tension,
whereas a slope of 2 is that expected of a viscous material (see text for details).

system with a well defined Young's modulus E,


iTyy = EYL (2.26)
Alternatively, if the response to the applied tension is viscous, i.e.
(2.27)
where 17e is an elongational viscosity coefficient, then it is easy to show that,
for the experiment performed here, the strain should increase quadratically
in time and hence give a slope of 2.
Both types of behaviour are evident, confirming the trend from elastic (at
low strain) to viscous response as the temperature is increased. The
estimated extensional viscosity at 500 K from the simulations is of the order
of 0.01 Pa s. Although this is much lower than the equilibrium extensional
viscosity of polyethylene, it is known that the viscosity does decrease
significantly with increasing strain rate (Bird et al., 1987). At the extension
rates used in our simulations, 10 7 -10 9 s -1, the behaviour is expected to be
strongly non-Newtonian.
One working definition of the yield stress in the laboratory is the true
stress at the observed maximum tension (Ward, 1985). For convenience we
have chosen to define the yield stress as the measured tension at a strain of
20%, which corresponds closely to observed maxima in the load for those
samples that show a maximum. The resultant values are plotted in Figure
64 Molecular dynamics modelling of amorphous polymers

300

: 'y
250 .. " Applied tension
~ 0.5 MPa/ps

100/~ Applied tension J.


50
0.1 MPa I ps
o~----~----~----~--~~----~
o 100 200 300 400 500
T/K
Figure 2.15 Stress at 20% extension ('yield stress') as a function of temperature for
amorphous polyethylene simulated using model PE1 with a single chain of 1000
methylene groups. Squares and circles refer to tension application rates of 0.5 and
0.1 MPa pS-1 respectively. Open symbols indicate data for which no discernible yield
was observed; these points are excluded from the curve fits and extrapolations to
zero tension. The error bars shown are the standard deviations in the results for the
five independent samples.

2.15 and the behaviour is very similar to that found in real systems where
the yield stress decreases approximately linearly with increasing temperature
(Ward, 1985).
The data cover a very wide temperature range which may account for the
slight nonlinearity. It has also been shown in laboratory experiments that
extrapolating to zero yield stress results in convergence close to the glass
transition temperature. If we ignore the points above 300 K for which there
is no discernible yield point, our data extrapolate to zero yield stress at
around the same temperature at which there is a change in expansivity
(Figures 2.8 and 2.14). As for laboratory measurements, there is a depend-
ence on the rate of application of the tension, with the lower rate leading to
consistently lower values of the yield stress and hence a lower extrapolated
temperature of zero yield stress.
It will not have gone unnoticed that the observed values of the
yield stress and strain are much larger than those typically observed in
the laboratory for a glassy polymer. There are several possible reasons
for this. For instance, the response of the system to very high rates of
strain will not include slow relaxation processes such as creep which are
important in laboratory experiments. Also, simulation samples are perhaps
too homogeneous; they do not contain the mesoscale heterogeneities which
Stress- strain properties 65

probably occur in laboratory samples and which may playa significant role
in determining the overall relaxation behaviour.
The behaviour of the system densities as the strain is increased at the
different temperatures is quite striking. As seen in Figure 2.l6, at high
temperatures (400- 500 K), where the flow process is predominantly viscous,
there is a hardly perceptible change in density during the extension of the
samples. In contrast, at low temperatures there is a noticeable dilation effect
as the tension is applied and the density decrease continues until just beyond
the yield point. Once the material yields, the density remains relatively
constant as plastic flow takes place. It appears that, for a given tension
application rate, this apparent critical density for yield and plastic flow is
independent of temperature over the range 10- 300 K.
The decrease in density that occurs under extension at the lower tempera-
tures is entirely consistent with the typical values of the Poisson ratio (J1)
for amorphous polymeric solids which are generally in the range 0.3-0.4.
Indeed the estimates of the Poisson ratio from the extensional and contrac-
tile strains in the simulations,

·
11m" - ')Ie
J1 = --+0 - - (2.28)
' L fL
give values of about 0.41 at the lowest temperatures.

0.84

0 10 K
'? tI
100 K
.to 200K
E
U
0>
-....
.• 300K
400K
0.78
. ......• :'l. ... •
Ie
SOOK
~ .to
'iJi " rl-
c
Q) 0.76 D
,.0 • .to • • 0

°D
0 0
0

0.74
xX·· x JCx • • x • • x • • JC Ie Ie

0.72
0 0.2 0.4 0.6 0.8 1.0
Strain
Figure 2.16 Behaviour of the density during the extension experiments for a tension
application rate of 0.5 MPa ps -1 for the same samples as represented in Figures 2.13
and 2.14. For the low temperature glassy samples note the sharp decrease in density
with increasing extension, which contrasts with the nearly constant density of
post-yield flow and with the viscous flow observed at high temperatures.
66 Molecular dynamics modelling of amorphous polymers

We should be a little cautious, however, in drawing too close a parallel


between these simulation results and laboratory experience since the above
behaviour of the density has, apparently, not been observed experimentally;
in fact, there appears to be evidence that the density may increase slightly
at yield (Ward, 1985). Also, it is worth recalling the well known experimen-
tal result that yield can also occur in amorphous polymers under compres-
sion. Simulations of model polymers under compression have yet to be
reported.

2.6.2 STRESS-STRAIN BEHAVIOUR AND CONFIGURATIONAL


PROPERTIES
As mentioned in section 2.1, one of the great advantages of model simula-
tions is the ease with which system parameters can be controlled in order
to study some physical effect. This approach was exploited in a recent study
of the ways in which the phenomenon of strain hardening depends on the
configurational properties of an amorphous polymer (McKechnie et al.,
1993). Samples of the same glassy polymer were prepared each with different
configurational properties, the aim being to avoid any ambiguities introduc-
ed by comparing polymers with different chemical structure.
Although it is not very evident from Figure 2.13, at high strains there is
nearly always an increase in the modulus, which is referred to as strain
hardening. This latter property is of great practical importance since its
extent is associated with the susceptibility of the material to necking and
crazing, phenomena which can also complicate the interpretation of experi-
mental data since they make true stresses and strains difficult to determine.
Elongation of an amorphous polymer can be achieved either by uncoiling
the overall chain configuration, as measured for instance by the radius of
gyration or by changing the local conformations within the 'tube' formed by
neighbouring chains in the entangled structure. This alternative 'local'
mechanism corresponds in our model to segments of chains being converted
from the gauche (,short') form to the long ('trans,) form; the latter is the
lowest energy state and has a planar zigzag structure. For the polyethylene
model used, it follows that chains which are either less coiled or contain
fewer gauche states to begin with will be harder to deform. In an equilibrium
polymer melt the overall configuration and the fractions of conformers are
inextricably linked, but this restriction does not apply for non-equilibrium
glassy states where changes in the preparation procedure can provide some
measure of independent control of the two properties.
For the study in question, the persistence length a was used as a measure
of the overall configuration. This property measures the correlation in the
orientation of successive monomers as we move along a polymer chain. One
useful definition which is easily applied to the polyethylene model used here
Stress-strain properties 67

is (Flory, 1988)
(fJ

a = bo L <ei'ei+k) (2.29)
k=O

where ei is a unit bond vector and bo is the backbone bond length. It can
be further shown that

(2.30)

where the characteristic ratio C oc is given by

C. -- l'1m <r;) _ «r r+k)2


--2 -
i - i
2 (2.31)
, k~ ex. kb o kb o
The above equations imply that a can only be obtained accurately from an
asymptotic limit. This limit is only reached when k > 1000 for the polyethy-
lene PEl model (McKechnie et ai., 1993), and an accurate result cannot be
achieved with only small samples. Nevertheless, correlation lengths were
deduced from averages over five configurations for values of k up to 100
(these are referred to as a 1 00) and it is reasonable to assume that to a good
approximation the ratios of these values reflect the ratios of the true
persistence lengths.
A set of molecular dynamics simulations was carried out on samples
prepared using the direct method described in section 2.3. All four sample
sets of polymer glass were based on the PEl model but have different
preparation histories. Each sample was relaxed at 200 K with P* =
0.1 MPa for ~ 1 ns. The associated correlation lengths and fractions of trans
conformers obtained from the final 200 ps of these runs are shown in Table
2.1.
Sample set A was formed by cooling from the melt. Set B was produced
by growing chains at 200 K and relaxing them for 1 ns at the same
temperature. Set C was obtained by 'flash' heat treatment of set B; this
involves raising the temperature instantaneously to 1000 K for 100 ps
followed by rapid cooling back to 200 K. This was not sufficiently drastic to
alter the overall configuration of the chain, but it did allow conformational
transitions to occur and the overall result of this treatment is a net decrease
in the trans fraction. Direct examination of configurations showed a notice-
able contraction of the polymer in the tube formed by its neighbours
(McKechnie et ai., 1993).
Sample set D was grown at 200 K with the torsional and 1-5 site
interaction potentials scaled down by a factor of 50; this has the effect of
increasing the gauche fraction and produce a highly coiled chain. Van der
Waals interactions were then introduced and subsequent relaxation with the
68 Molecular dynamics modelling of amorphous polymers

Table 2.1 The percentage of trans conformers (% trans)


and correlation length, a100 , calculated for each of the
four sample sets

Sample % trans a100(nm)

A 78 0.50
B 82 1.55
C 77 1.20
D 70 0.58

full interaction potentials produced a polymer glass with a much reduced


fraction of trans states.
The stress-strain behaviour of all four sets of samples was obtained by
subjecting them to an externally applied uniaxial tension which was in-
creased at a rate of 0.5 MPa ps - 1, in exactly the same way as described in
the previous section. As a result of the small system size, data were obtained
only up to extensions of '" 100%. In Figure 2.17 the load on these samples
is plotted as a function of the strain, the load being determined from the
product of the stress and the cross-sectional area. Since the samples are very
small, the loads are also extremely small (of the order of 10 - 9 N).
The pattern of the stress-strain plots is similar to that discussed in the
previous section - there is an initial elastic response followed by yield and
plastic flow. In detail, however, the four sets of samples show rather different
behaviour; the differences begin to be noticeable after about 10% extension.
Samples Band C both show enhanced resistance to extension beyond the
point at which the A samples yielded ( '" 20% extension). The high per cent
trans-high correlation length samples (set B) show the largest extent of
strain hardening and in particular produce significantly more stress than the
set C which, within the error, has a similar correlation length but lower trans
fraction. Conversely, sample set C has practically the same per cent trans as
set A, so the differences here must be due to their contrasting configurational
structure. Set D shows the lowest resistance of all to the applied tension, as
was expected from its highly coiled structure with a smaller fraction of trans
conformations. What these results suggest is that both an increase in the
fraction of trans states and an increase in the persistence length can
independently contribute to strain hardening.
It is interesting to note the strain dependence of the fraction of trans
conformers. These data are shown in Figure 2.18 and they reveal a common
feature of all uniaxial tension simulations to date, namely that there is
a consistently linear dependence of the per cent trans upon the extension.
The lack of any discontinuities or breaks in the plots at the yield points
Stress-strain properties 69

2.5

2.0
B c
Z
c:
...... 1.5
"C

--
ctl
0 1.0
....J

0.5 1 200 K I

0.00
0.2 0.4 0.6 0.8
strain
Figure 2.17 Measured load plotted as a function of elongational strain for four
glassy samples of the PE 1 polyethylene model directly prepared at 200 K by different
procedures as discussed in the text. The configurational properties are given in Table
2.1. The samples were subjected to a uniaxial tension increasing at a rate of
0.5 MPa ps -'. Data for each set are averaged over five samples.

confirms the conclusion that this mechanical property has nothing directly
to do with the onset of transitions between different conformational states.
Technical developments have now made it possible to measure true
stress-strain curves and so to derive reliable values of strain hardening
coefficients, as shown in Chapters 4 and 5. The results generally agree with
the conclusions reached here. For instance, it is found that the level of strain
hardening increases with persistence length. A striking example is that of the
polyisocyanates, which are known to have extremely high values of the
persistence length (or alternatively the Kuhn length) and also show signifi-
cant strain hardening, as exemplified by their behaviour in a conventional
tensile test where uniform deformation is observed (Owadh et ai., 1978;
Haward, 1993).
To conclude this section it is interesting to examine the topological
changes that accompany uniaxial strain. Figure 2.19 shows single configur-
ations of the chain from one of the highly coiled D samples before and after
extension. The view is in a plane perpendicular to the direction of strain and
shows the contours of the continuous primary chain consisting of 1000 sites.
This chain passes through many of the replica cells.
It is evident that there is a significant expansion of the chain in the
deformation direction and a concomitant shrinkage in the orthogonal
directions. What is striking, however, is that the overall pattern of chain
topology has not changed significantly during the extension. This is particu-
70 Molecular dynamics modelling of amorphous polymers

g
90

85 -8
, ---c
," -0

-
(J)

...
t: 80
ro ~.-::-.-

-r....r;.-;:--
~
0 75

70
0 20 40 60 80 100 120
O/o"{
Figure 2.18 Percentage of trans conformers plotted as a function of the percentage
strain for sample sets A-D, as in Figure 2.17.

larly noticeable in Figure 2.19B, which shows the same data except that the
final, deformed configuration has been subjected to a linear transformation
that maps the coordinates of the vertices of the deformed unit cell onto those
of the undeformed cell. This is a very general result, as is confirmed by visual
examination of the deformation of many different polymer glasses.
The above result lends support to the notion of a 'tube' formed by
interactions with entangled neighbouring chains. It also suggests that the
deformation is highly affine in the strain, although it is possible that
the simulations are too short to observe significant non-affine motions.
The effect must be caused by topological constraints which have relaxation
times many orders of magnitude longer than the time scale of either
simulations or even of some laboratory experiments (Ferry, 1980). This may
be the underlying reason for the basic similarity of stress-strain phenomena
in the two cases. As can be seen from Figure 2.19, there are regions where
the chain structure has changed on a local scale of tens of monomers, but
after laborious examination of the structures there is no clear evidence for
the existence of geometrically defined 'entanglements'.

2.6.3 STRAIN RECOVERY


Real glassy polymers retain a remarkable memory of their original structure,
even when subjected to substantial mechanical deformation beyond the
yield point. Hence it is often found that most of the strain is readily
recovered if a sample is heated to just above the glass transition tempera-
ture. The explanation is usually given in terms of the extremely long
Stress- strain properties 71

A. ormal cartesian coordinates

extension
direction

B. Afle r an affine transformation of coordinates which


superimposes the deformed and undeformed unit cells

extension
direction

Figure 2.19 Examples of configurations of the primary chain from sample set D
before extension (thin line) and after 107% uniaxial strain (thick line). (A) In regular
cartesian coordinates; note the contraction of the chain perpendicular to the
direction of strain. (B) After applying a linear transformation to the strained
configuration coordinates such that the deformed cell basis vectors map exactly onto
those of the original. The right-hand portions of the pictures are enlarged by the
perspective.

relaxation times associated with disentangling the polymer chains in


such amorphous materials. On distance scales less that the characteristic
entanglement length, deformation can still be accommodated by 'local'
changes in chain configurations which, although they remain frozen in the
glass, relax much more rapidly just above I'g.
The same phenomenon can be demonstrated in computer simulations and
some recent results are shown in Figure 2.20 for the directly prepared
extended chain sample set B (Clarke and McKechnie, 1996). What is shown
here is the elongational extension of this glassy polymer plotted as a
72 Molecular dynamics modelling of amorphous polymers

function of time. Also included are the corresponding behaviour of the


density and the fraction of trans conformers during extension. When the
extension has reached about 50% Gust before the end of the simulation
described in section 2.6.2) the tension is removed and the properties are
monitored, first for the next 0.5 ns at 200 K and subsequently for a further
0.5 ns after raising the temperature to 500 K.
The 'strain recovery' occurs in two distinct regimes. In regime 2, still at
200 K, there is a very rapid recovery of the essentially elastic (Hookean)
deformation. This is accompanied by an extremely fast relaxation of the
density to a value close to that at the start of the simulation. While some of

~:
= I
tension _
Appl.ied ;

T/K :~~====I
60
c: I
0

-
'Ci) : relaxatiop
c: 40
Q) I
x
Q) 20
I
I
I
~
0 I
0
0.80
C?
E 0.78
(.)
C) 0.77
...... 0.75
0-
---~
0.73
86
en

-
c: 83
....as
79
ffl
76
0 450 900 1350 1800
time/ps
Figure 2.20 The response of PEl model polyethylene to uniaxial extension under
load and the subsequent recovery phenomena following first the release of tension
and second raising the temperature above the glass transformation region. The data
apply to set B (see Table 2.1 and text for details). The arrow indicates the long time
value of the density at 500 K.
Penetrant diffusion 73

the strain recovery can be ascribed to the effects of van der Waals forces
between monomers, it is evident from the bottom graph of Figure 2.20 that
part of the recovery is also associated with the return of local chain
conformations to their pre-tension values.
If the temperature is now raised above the glass transformation range
(regime 3), there is further rapid recovery of the strain and an indication of
a much longer relaxation which is not fully accessible on the simulation time
scale. Similarly, in this regime the density and trans fraction begin to relax
towards new values appropriate to 500 K. The magnitude of the elastic
recovery in regime 2 is much larger than would be obtained in laboratory
polymers but, as before, this can be explained in terms of the ultra-short
time scale of the simulation experiments, which limits the contribution of
relaxation processes such as viscous flow and creep during the deformation.

2.7 PENETRANT DIFFUSION


An account of the diffusion of gases and liquids in glassy polymers was given
in the first edit on by Hopfenberg and Stannett (1973). Since then much
progress has been made in simulation studies. The permeation of small
molecules such as He, Oz, CO 2 and CH 4 through amorphous and partially
crystalline polymers is of great practical interest. For instance, polymer
materials are becoming increasingly important on the one hand as selective
membranes in gas separation technologies and on the other hand as highly
impenetrable but flexible barrier membranes. It is not surprising therefore
that this phenomenon has already attracted a great deal of attention in
simulation and related studies. Although there are well established experi-
mental methods for measuring permeation properties of polymers, computer
simulation studies can in principle provide important insight into the
mechanisms of small molecule transport and hence, in combination with
experimental data, contribute to the molecular design of host polymers.

2.7.1 SOLUBILITIES AND DIFFUSION COEFFICIENTS


Self-diffusion coefficients in fluids are most commonly obtained in molecular
dynamics simulations using the Einstein relation

(2.32)

The diffusion coefficient D is then obtained from the slope of a plot of the
mean squared displacement <.1r 2 ) against time t. The problem is deciding
when the limiting linear regime has been reached. Times of the order of
100 ps are usually sufficient to attain this condition in simple fluids. The
74 Molecular dynamics modelling of amorphous polymers

situation may not be so straightforward in more complex systems. In viscous


molecular fluids, for instance, the slope has been observed to change
significantly over the time range 50-300ps (Brown and Clarke, 1987).
Gas solubilities can also be determined by molecular dynamics simula-
tions (Muller-Plathe, 1991; Muller-Plathe, Rogers and van Gunsteren, 1993)
using the Widom test particle insertion method (Widom, 1963) to calculate
the excess chemical potential or free energy /lex of the penetrant molecules.
If <I> is the interaction energy of a virtual penetrant molecule with the
polymer inserted at random within the sample (the molecule is 'invisible' to
the polymer) then
/lex = -RTln(exp( -<I>/kT» (2.33)
The solubility can then be obtained from /lex using Henry's law.
In dense materials the original implementation of this method can be
quite inefficient and a better way to sample the structure is to bias the choice
of insertion points towards those that have an inherently higher probability;
this can be done using an accessible volume map (Deitrick, Scriven and
Davis, 1989; Sok, Berendsen and Van, 1992).
Comparisons with the extensive experimental data available on a wide
range of polymers provide a stiff test not only of the accuracy of force fields
but also of how representative the polymer microstructure is of the real
material. Molecular dynamics simulations of penetrant diffusion coefficients
have so far been performed for polypropylene, polyethylene, polyisobutylene
and polydimethylsiloxane. Some of the results are shown in Table 2.2. It is
seen that the agreement between experiment and simulation is not startling,
although several of the experimental trends involving different penetrants in
the same polymer are reproduced. Reliable comparisons with experimental
data are made difficult by uncertainties as to the true diffusion coefficients
in pure amorphous polyethylene, which have to be extracted by calculation
from data on the partially crystalline material.
In nearly all cases the diffusion coefficients and solubilities from simula-
tions exceed the experimental values, and there are several possible causes
of this phenomenon. The use of the united atom approximation to model
interaction involving methylene groups has, for instance, been criticized;
inefficient packing of the spheres may lead to far too much spare volume in
the model structures. It appears that better results can be obtained using
either the anisotropic united atom model (Pant and Boyd, 1993) or an
all-atom description (Mueller-Plathe, Rogers and van Gunsteren, 1992).
Another possible source of discrepancies is incomplete relaxation of the
polymer sample densities; one expects diffusion coefficients to be extremely
sensitive to density in this regime and this again emphasizes the important
issue of how model polymer samples are best prepared. In one study
(Muller-Plathe, Rogers and van Gunsteren, 1993) it was pointed out that
Penetrant diffusion 75
Table 2.2 Comparisons between MD simulation and experiment for diffusion
coefficients (D) and gas solubilities (given as the Henry law constants, H) for a range
of penetrants in amorphous polymer samples

Polymer Penetrant D/(1 0-6 cm2s-') H/(bar')

Simulation ExperimenF Simulation Experiment g

Atactic H2 44a 5.7 0.78" 0.13


polypropylene O2 4.0a ~1.5 4.0b
CH 4 OA8 a ~0.6 10.0 b
Poly(dimethyl- He 18 c 10 0.2c 0.046
siloxane) CH 4 2.1 c 2.0 11.0c 0.50
Polyethylene CH 4 0.5 d 0.3~0.6
CH 4 1e
Polyisobutylene He 301 5.93 0.19 f 3.2
H2 9.21 1.52 19 9.7
O2 0.047-0.171 0.081 181 11
CH 4 OA' (350 K) 1.7(375 K) 9.7 1 0.12

'Muller-Plathe, F. (1992) f. Chern Phys., 96, 3200.


bMulier-Plathe, F. (1991) Macromolecules, 24, 6475.
'Sok, R.M., Berendsen, H.J.C. and van Gunsteren, W.F. (1992) f. Chem. Phys., 96, 4699.
dPant, P.V.K. and Boyd, R.H. (1992) Macromolecules, 25, 494.
'Pant, P.v.K. and Boyd, R.H. (1993) Macromolecules, 26, 679.
fMulier-Plathe, F., Rogers, s.c. and van Gunsteren, W.F. (1993) f. Chern. Phys., 98,9895.
gBrandrup, J. and Immergut, E.H. (eds) (1989) Polymer Handbook, 3rd edn, Wiley, New York.

the bulk of the thermodynamic solubility of O 2 in the sample of model


polyisobutylene was contributed by a single large 'hole' in the structure,
emphasizing the need for adequate sampling of the way in which a penetrant
samples the natural structural fluctuations in the polymer, either by using
very large systems or by averaging over many independent simulations.
It must be emphasized that the limited quantitative success so far
achieved has been with penetrants and system conditions where the diffusion
rate is quite high, so that the process can be characterized during the time
scale of a simulation experiment. It will be much more of a challenge to
compute reliable diffusion coefficients in glassy polymers and effective
barrier systems where spontaneous diffusion is required to be extremely
small.

2.7.2 THE DIFFUSION MECHANISM


One of the attractions of molecular dynamics studies is the potential insight
it might provide concerning the mechanism of penetrant molecule diffusion
in polymers and this has been investigated in several recent studies.
76 Molecular dynamics modelling of amorphous polymers

8.0 , - - - - - - , - - - - . - - - . - : - - - - - - - , - - - - - - ,

H,

time/ps

Figure 2.21 Displacements against time for selected molecules of hydrogen, oxygen
and methane in amorphous polypropylene at 300 K as determined by molecular
dynamics simulations. The initial displacements are arbitrary. (Reproduced from
Gusev et a/., 1994, with permission.)

There is accumulating evidence that diffusion proceeds by a some kind of


'rattle and jump mechanism'. The simplest and most direct way of analysing
the simulation data is to select a 'typical' penetrant molecule and plot the
displacement of its centre of mass as a function of time. This is shown in
Figure 2.21 for various diffusing molecules in polypropylene at 300 K. In the
case of O 2 there are long periods of small vibrational-like motions (and
large angle orientational motions; Pant and Boyd, 1993) interrupted by
rapid displacements of 0.5-1.0 nm. For methane in PE at 300 K typical rms
jump lengths appear to be about 0.5 nm with the time between jumps of the
order of 0.5 ns. At higher temperatures the hopping rate increases and the
magnitude of rattling motions increases, so the mechanism is less well
defined.
Recent simulation studies seem to suggest that the jump events may be
associated with the opening up of temporary channels between cavities of
fluctuating size and shape (Muller-Plathe, Rogers and van Gunsteren, 1993;
Sok, Berendsen and Van Gunsteren, 1992) so the local mobility of the
polymer chains is intimately involved. There is direct evidence of the
participation of local motions of the polymer chain in the process of
diffusion. Removal of the torsional potential in polyethylene considerably
increases the chain mobility and gave an increase of about a factor of 2 in
the diffusion coefficient of O 2 (Takeuchi, Roe and Mark, 1990). Conversely,
Penetrant diffusion 77

decreasing the magnitude of the torsion angle fluctuations in a polyethylene


model drastically reduces the diffusion coefficient of methane (Pant and
Boyd, 1993). It would be a mistake to assume, however, that jumps are
necessarily correlated with conformational transitions; certainly there is no
clear evidence of this to date. In any case, for glassy polymers such
transitions are rare events on the time scale of molecular dynamics simula-
tions.
Penetrant diffusion has been widely discussed in the literature in terms of
free volume theories of polymers, and it is not surprising to see this idea
taken up in simulation studies. The problem is that although geometric free
volume is a physical concept that is easy to comprehend, it suffers from lack
of precise definition. In the simulation studies to date it has been identified
with the unoccupied volume in the sample calculated as an average over
many static configurations. Note that this definition is different from the
accessible volume to penetrant molecules discussed earlier in this section.
This kind of analysis has been performed for amorphous polyethylene,
polyisobutylene and polypropylene models, for instance, by prescribing
spheres of a chosen diameter around each interaction site, or by using
Voronoi polyhedra to define the unoccupied space (Rigby and Roe, 1990).
The free volume fraction fv can be defined by

(2.34)

where Do is the volume of the simulation unit cell and rc is the volume
occupied by the polymer segments. In the case of united atom models with
a Lennard-Jones representation of the van der Waals forces, the (J size
parameter is taken as the hard core diameter in order to calculate Dc. Rather
surprisingly, values obtained for Iv are quite high, giving the fraction of
volume unoccupied as 30-40% in model PE and polyisobutylene (Pant and
Boyd, 1993), although only a much smaller fraction, typically '" 0.1 %, is
accessible to penetrants.

2.7.3 TRANSITION STATE MODELS OF DIFFUSION


The conclusion from molecular dynamics studies that penetrants appear to
diffuse by a sequence of activated 'hops' has led to the development of the
transition state approach (TSA) (Gusev et aI., 1994) which shows particular
promise for extending studies of diffusion to the microsecond regime or even
beyond - much longer times than would be feasible with direct simulation.
This method (Gusev and Suter, 1993) extends the classical ideas introduced
more than 30 years ago that diffusion in fluids is related to the redistribution
of free volume. It utilizes atomistic simulation to characterize the energy
78 Molecular dynamics modelling of amorphous polymers

surface controlling the motion of the penetrant in a particular polymer, and


is based on the assumption that the dynamics of small molecules dissolved
in dense polymers are coupled to the elastic thermal motion of dense
polymers, but not to their structural relaxation. This accords with the
evidence from molecular dynamics that there is no direct correlation with
conformational changes, at least for small penetrant molecules. The assump-
tion is likely to break down in cases where penetrants cause a significant
deformation of the polymer structure.
The actual procedure in TSA is in fact quite involved (Gusev and Suter,
1993), although still less computationally intensive than full molecular
dynamics simulations. Central to the method is the determination of a solute
spatial distribution function p{r) at a fine grid of points r, where

p{r) = f
1) dL\~ exp ( - 2(~~) - Uk~~~») (2.35)

Here (L\;) is the mean square thermal deviation for each of the (local) host
polymer atoms IX; it represents the elastic motion of the polymer and is
similar to the Debye-Waller factor encountered in X-ray crystallography.
Without the dependence on L\~, the first term in brackets disappears and one
has the simple expression for the solute distribution in a 'frozen' host
structure.
Values of p{r) are needed to compute configurational partition functions
Z, both for solute sites and the activated states ij between i and j. The rate
constant Rij of the associated solute transition is then calculated to be
(Gusev, Zehnder and Suter, 1994)

R.. = ( kTZij )1/2 (2.36)


lJ 8nmZ.1 1

Neglecting differences between the behaviour of different sites an average


smearing factor (L\2) for a specific penetrant in a particular host polymer
structure can be determined by molecular dynamics. Monte Carlo sampling
is used to determine up to 1000 possible stochastic solute trajectories from
which mean square displacements can be calculated from the relevant
combinations of Rij values.

2.7.4 ANOMALOUS DIFFUSION


We end this section with a brief discussion of the phenomenon of anomalous
diffusion, which is a feature of all modelling studies of diffusion to date and
which has been extensively discussed (Muller-Plathe, Rogers and van
Gunsteren, 1993; Gusev, Zehnder and Suter, 1994; Takeuchi and Okazaki,
1996). In the true (long time) diffusive regime conforming to the Einstein
Penetrant diffusion 79

ii,,,
1'-' '~~ ,
I

::: f
'~~
N
f
l'TSA I
.~ 103 r
r
'-..... ,
Cl "
m
~ 102 r .....
...... " " "
"

1
~ "
"

~
10
,,'
f
10 0 t II
10- 13 10- 12 10- 11 10- 10 10-9 10- 8 10- 7

tis
Figure 2.22 Time dependence of the mean square displacement for helium in
amorphous bisphenol A polycarbonate at 300 K and 1 bar pressure, as determined
by molecular dynamics (MD) simulations (this is an average over ten solute molecules)
and by the transition state approach (TSA, averaged over 500 walks). The compari-
son shows reasonable agreement up to the limit of the molecular dynamics data
(- 0.5 ns). The straight line drawn up to 10 -7 s shows the diffusive regime which is
only accessible by TSA. (Reproduced from Gusev et a/., 1994, with permission.)

relation equation (2.32) then <i1r 2 ) is linear in time. In fluid systems, short
time departures from this behaviour due to vibrational motion are always
found. In the case of amorphous polymers, however, for rms displacements
at longer times in the range - 4- 20 A it is found that <i1r 2 >is proportional
to something less than the first power of t; this regime is clearly distin-
guished for helium in amorphous bisphenol A polycarbonate in the double
logarithmic plot in Figure 2.22.
This so-called anomalous diffusion can be related to the spatial in-
homogeneity of diffusion paths, i.e. on medium length scales the penetrant
motion is not truly random. Unfortunately the periodIcity of polymer
models emphasizes this inhomogeneity on the same distance scales, and in
order to remove the possibility of this artefact it would be necessary to
perform very large scale molecular dynamics calculations involving about
ten monomers.
From Figure 2.22 it is comforting to note the agreement between the
predictions of TSA and molecular dynamics as far as the long time
(diffusive) motion is concerned. Indeed both techniques show a similar
anomalous diffusion regime, but it must be recalled that both calculations
are based on atomistic simulations of only moderately sized samples.
80 Molecular dynamics modelling of amorphous polymers

2.8 CONCLUSIONS AND FORWARD LOOK

The aim of this article has been to outline the ingredients of atomistic
molecular dynamics simulations as applied to dense polymers, to discuss
some of the applications relevant to glassy polymers and to indicate how
such simulations can be used to improve our understanding of the physical
behaviour of these materials.
The discussion has been limited to simulations which attempt to model
the atomistic detail of polymers. Many aspects of polymer behaviour can
however be studied successfully using simpler 'coarse grain' models and it is
relevant to mention some of this work here. In these models the atomistic
detail is ignored and a polymer chain is considered as a connection of freely
jointed 'statistical units' or Kuhn segments; the size of such units can be
equated to the magnitude of the persistence length of a real polymer chain.
The use of such models to study entanglement effects and some properties
of networks has recently been reviewed (Kremer and Grest, 1995). More
relevant to the current article such coarse grain models have also been used
to examine the relative contributions to the stress from the bonded interac-
tions and the van der Waals forces in polymer networks subjected to
extensional strain (Gao and Weiner, 1991). More recently, studies have been
extended to stress relaxation in entangled melts (Gao and Weiner, 1995).
In any atomistic simulations, glasses are formed at enormously high
effective cooling rates (beyond what is possible even with splat quenching
techniques in the laboratory) and this has often led to their description as
'configurationally arrested states' in an attempt to distinguish them from
glasses formed in the laboratory. It has been shown in this article, however,
that phenomenologically there are many similarities between material
responses observed in simulations and those measured for experimental
polymer glasses. In any case the dependence of physical properties on the
rate of formation is part of the definition of a glass and the materials formed
in simulations can be considered a limiting case.
A much more important question is how we can bridge the gap in time
scales (and to a lesser extent the spatial scales) between simulation and
experiment. This is likely to be a very active and fruitful area for research in
the next decade. Several approaches seem possible. So-called 'hybrid' methods
may have a role to play; in this case Monte Carlo or long time-step molecular
dynamics can be used to move a system rapidly through phase space between
periods of atomistic level dynamics. There are likely to be developments in the
use of coarse grain (see above) and meso scopic simulation methods although
these will have real value to experimentalists only if ways are found for relating
such models quantitatively to the atomistic detail.
Conclusions and forward look 81

There is no doubt that the number of simulation studies of dense


amorphous polymers will increase significantly with the more widespread
availability of the enabling software and hardware technologies. These
studies are likely to make their strongest impact when close comparisons
can be made with experimental data from e.g. neutron scattering, NMR and
spectroscopic techniques; the opportunities for increased insight and under-
standing into dynamic properties are particularly exciting as the time scales
accessible to these experiments and to atomistic simulations begin to
overlap.
An additional stimulus to experimental and simulation work is the
recent applications of mode coupling theories to study polymer dynamics
near the glass transition (Gotze and Sjogren, 1992; Schweizer, 1991). These
theories claim to provide a first-principles explanation for the rapid slowing
down of the dynamics of polymers as the temperature is decreased. The
mechanism offered is reasonably straightforward. Phenomenologically it
is clear that the behaviour of transport properties such as the viscosity
are strongly related to density. What these theories do is to provide a
statistical mechanical description of the coupling of fluctuations of the
stress and density and how the associated feedback mechanisms can result
in the dramatic increase or divergence of viscosity at some finite tempera-
ture. There are some major approximations and difficulties involved in
mode coupling theories, but some of the predictions have been broadly
confirmed by recent simulation studies on amorphous polyethylene (Roe,
1994).
Currently, molecular dynamics is the modelling technique with the
greatest potential, although energy minimization methods may continue
to prove useful for some polymer applications as discussed in Chapter 4
of this book. There are of course many commercial and public domain
molecular dynamics codes available, but mention should be made of
a highly versatile program DL_POLY which was developed by the
Collaborative Computing Project Nt). 5 of the UK EPSRC (Smith
and Forester, 1996) [Copies can be obtained from the authors, Dr W.
Smith and Dr T. Forester at the Daresbury Laboratory, Warrington,
WA4 4AD, UK. See also the WWW reference: http://www.dl.ac.uk/
TCS/Software/DL_POLY/main.html.] Atomistic Monte Carlo simulations
have yet to make an impact in the field of dense polymers; there
appear to be major difficulties in designing suitable moves which are
statistical mechanically acceptable and also provide efficient configur-
ational sampling. Such definitive studies that are available suggest that
Monte Carlo offers no advantage over molecular dynamics for dense
systems.
82 Molecular dynamics modelling of amorphous polymers

REFERENCES
Allen, M.P. and Tildesley, D.J. (1987) Computer Simulation of Liquids, Clarendon
Press, Oxford.
Angell, c.A. (1988) J. Phys. Chem. Solids, 49, 863.
Angell, C.A. and Torrell, L.M. (1983) J. Chem. Phys., 78, 937.
Angell, C.A. Clarke, J.H.R and Woodcock, L.V. (1981) Adv. Chem. Phys., 48, 397.
Berendsen, HJ.C., Postma, 1.P.M., van Gunsteren, W.F. et al. (1984) J. Chem. Phys.,
81,3684.
Binder, K. (ed.) (1995) Monte Carlo and Molecular Dynamics Simulations in Polymer
Science, Oxford University Press, New York.
Bird, RB., Curtiss, C.F., Armstrong, R.C. and Hassager, O. (1987) Dynamics of
Polymeric Liquids, Vol. 1, Fluid Mechanics, 2nd edn. John Wiley, New York, p.
173.
Boyd, RH., Gee, RH., Han, 1. and Jin, Y. (1994) J. Chem. Phys., 101, 788.
Brandrup,1. and Immergut, E.H. (eds) (1989) Polymer Handbook, John Wiley, New
York.
Brown, D. and Clarke, J.H.R. (1984) Mol. Phys., 51, 1243.
Brown, D. and Clarke, 1.H.R (1987) J. Chem. Phys., 86,6446.
Brown, D. and Clarke, 1.H.R. (1990) J. Chem. Phys., 92, 3062.
Brown, D. and Clarke, 1.H.R (1991) Macromolecules, 24, 2075.
Brown, D., Clarke, 1.H.R., Okuda, M. and Yamazaki, T. (1994a) J. Chem. Phys., 100,
6011.
Brown, D., Clarke, 1.H.R., Okuda, M. and Yamazaki, T. (1994b) J. Chem. Phys., 100,
1684.
Brown, D., Clarke, 1.H.R., Okuda, M. and Yamazaki, T. (1996) J. Chem. Phys., 104,
2078.
Clarke, J.H.R (1978) in Advances in Infrared and Raman Spectroscopy, Vol. 4, (eds.
R.E. Hester and J.H.R. Clarke, Heyden, London, p. 109.
Clarke, J.H.R and McKechnie, 1.1. (1996) submitted for publication.
de Gennes, P.G. (1979) Scaling Concepts in Polymer Physics, Cornell University
Press, Ithaca, NY.
Deitrick, G.L., Scriven, L.E. and Davis, H.T. (1989). J. Chem. Phys. 90, 2370.
Doi, M. and Edwards, S.F. (1986) The Theory of Polymer Solutions, Clarendon
Press, Oxford.
Ferry, J.D. (1980) Viscoelastic Properties of Polymers, John Wiley, New York.
Flory, P.J. (1985) Polym. J., 17, 1.
Flory, PJ. (1988) The Statistical Mechanics of Chain Molecules, Hanser Publishers,
New York.
Fox, H. and Anderson, H.C. (1971) J. Phys. Chem., 88, 4019.
Gao,1. and Weiner, 1.H. (1991) Macromolecules, 24, 1519.
Gao,1. and Weiner, 1.H. (1995) J. Chem. Phys., 103, 1614.
Gotze, W. and Sjogren, L. (1992) Reports Prog. Phys., 55, 241.
Gusev, A.A. and Suter, U.W. (1993) J. Chem. Phys., 99, 2228.
Gusev, A.A., Zehnder, M.M. and Suter, U.W. (1994) Macromolecules, 27,615.
Gusev, A.A., Zehnder, M.M. and Suter, UW. (1995) Macromolecular Symposia, 90,
85.
Gusev, A.A., Zehnder, M.M. and Suter, UW. (1996) Phys. Rev. E.: Condensed
M atler, 54, 1.
Gusev, A.A., Muller-Plathe, F., van Gunsteren, W.F. and Suter, UW. (1994) Adv.
Polym. Sci., 116, 207.
References 83

Han, 1., Gee, R.H. and Boyd, R.H. (1994) Macromolecules, 27, 7781.
Haward, R.N. (1993) Macromolecules, 26,5860.
Hopfenberg, H.B. and Stannett, V. (1973) in The Physics of Glassy Polymers (ed.
R.N. Haward), Applied Science Publishers, London, Ch. 9.
Kremer, K. and Grest, G.S. (1995) in Monte Carlo and Molt?cular Dynamics
Simulations in Polymer Science (ed. K. Binder), Oxford University Press, New
York.
McKechnie, J., Brown, D. and Clarke, 1.H.R. (1992) Macromolecules, 25, 1562.
McKechnie, J.I., Haward, R.N., Brown, D. and Clarke, 1.H.R (1993) Macro-
molecules, 26, 1982.
Muller-Plathe, F. (1991) J. Chem. Phys., 94, 3192.
Muller-Plathe, F., Rogers, S.c. and van Gunsteren, W.F. (1992) Macromolecules, 25,
6722.
Muller-Plathe, F., Rogers, S.c. and van Gunsteren, W.F. (1993) J. Chem. Phys., 98,
9895.
Neelov, I.M. and Clarke, 1.H.R. (1994) Macromolecular Symposia, 81,55.
Neyertz, S. and Brown, D. (1995) J. Chem. Phys., 102, 9725.
Neyertz, S., Brown, D. and Clarke, 1.H.R. (1996) J. Chem. Phys., 105, 2076.
Owadh, A.A., Parsons, I.W., Hay, 1.N. and Haward, R.N. (1978) Polymer, 19, 386.
Pant, P.Y.K. and Boyd, R.H. (1993) Macromolecules, 26, 679.
Parrinello, M. and Rahman, A. (1985) Phys. Rev. Lett., 4069.
Richter, D. (1992) Physica B, 180-181 (Part A), 7.
Rigby, D. and Roe, R.J. (1987) J. Chem. Phys., 87, 7285.
Rigby, D. and Roe, R.J. (1990) Macromolecules, 23, 5312.
Roe, R.J. (1994) J. Chem. Phys., 100, 1610.
Roe, R.J., Rigby, D. and Furuya, H. (1992) Comput. Polym. Sci., 2. 32.
Schweizer, K.S. (1991) J. Non-Crystalline Solids, 131,643.
Siepmann, 1.1. and Frenkel, D.A. (1992) Mol. Phys., 75, 59.
Smith, G.D., Yoon, D.Y. and Jaffe, R.L. (1993) Macromolecules, 26,5213.
Smith, G.D., Yoon, D.Y. and Jaffe, R.L. (1995) Macromolecules, 28.5897.
Smith, G.D., Yoon, D.Y., Zhu, W. and Ediger, M.D. (1994) Macromolecules, 27,
5563.
Smith, W. and Forester, T.R. (1996) J. Molecular Graphics., in press.
Sok, R. M., Berendsen, H.J.C. and Van Gunsteren, W.F. (1992) J. Chem. Phys., 96,
4699.
Spiess, H.W. (1991) Chem. Rev., 91, 1321.
Steele, D. (1985) J. Chem. Soc. Faraday Trans. II, 81, 1077.
Takeuchi, H. and Okazaki, K. (1996) Molecular Simulation, 16, 75.
Takeuchi, H. and Roe, R.J. (1991) J. Chem. Phys., 94, 7458.
Takeuchi, H., Roe, R.J. and Mark, J.E. (1990) J. Chem. Phys., 93, 9042.
Theodorou, D.N., and Suter, V.W. (1986) Macromolecules, 19, 379.
Toxvaerd, S. (1990) J. Chem. Phys., 93, 4290.
Turnbull, D. (1969) Con temp. Phys., 10, 473.
Ward, I.M. (1985) Mechanical Properties of Solid Polymers, John Wiley, Chichester.
Weber, T.A. and Helfand, E. (1979) J. Chem. Phys., 71, 4760.
Weber, T.A. and Helfand, E. (1983) J. Phys. Chem., 87, 2881.
Weeks, J.D., Chandler, D. and Anderson, H.c. (1971) J. Chem. Phys., 54, 5237.
Wid om, B. (1963) J. Chem. Phys., 39, 2802.
Williams, G. (1979) Chem. Soc. Rev., 7, 89.
Yong, c.w. and Clarke, 1.H.R. (1996) submitted for publication.
Relaxation processes
and physical aging 3
1. M. Hutchinson

3.1 INTRODUCTION
Glassy materials have been of service to mankind for centuries, from the
earliest soda-lime-silicate glasses to the present sophisticated developments
in, for example, metallic, fluoride and superionic glasses. Throughout this
time, artefacts have been produced with a very wide range of functions:
objects of exquisite beauty, such as the Venetian 'mille fiori' ornaments and
the stained glass windows of many ancient cathedrals, high technology
applications such as optical fibres for communications, 'smart' glass for
controlled transmission, and slow-release glasses for regulating drug dos-
ages, as well as the more commonplace requirements for containers and
glazing. What is perhaps surprising about this remarkable success of glass
in general is that the glassy state should still remain today so difficult to
define with any precision. Even the seemingly simple question of whether a
glass should be considered as a liquid or as a solid remains open to
discussion.
The state of matter which constitutes the important class of glassy
materials is one which displays mechanical and physical properties similar
to those of crystalline solids, whilst retaining a molecular or atomic
arrangement more characteristic of a liquid, with no long range order. One
important aspect of this disordered state is that it can confer upon the glass
a large mobility, not only macroscopically in respect of the bulk flow
properties of the liquid, but also microscopically in that localized molecular
motions are possible even in an apparently rigid solid. These latter give rise
to the relaxation processes which can have a profound effect on many of the
properties of glass. Good examples may be found from quite different fields.
Amongst the inorganic glasses, the class of materials known as fast ion

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
86 Relaxation processes and physical aging

conductors, such as the silver iodomolybdates, show a remarkably high


ionic mobility in the solid state. This involves the motion of relatively large
ions along some kind of pathways through the glassy network structure,
which is believed to occur by the cooperative relaxation of the network. For
polymeric glasses, a particularly important relaxation phenomenon is that
of physical aging, discussed in more detail later. The process of aging leads
to time-dependent changes in physical and mechanical (and other) proper-
ties, with important consequences for the design and use of components and
structures fabricated from these materials.
The selection of examples from inorganic as well as organic glasses is
deliberate, as there is a remarkable universality in many aspects of relax-
ation behaviour. Indeed, the range of materials within the scope of non-
crystalline solids would nowadays include, amongst others, elements (e.g.
amorphous selenium), oxides, cha1cogenides, metallic glasses and polymer
glasses. In nearly all cases, the most common procedure for transforming
from the liquid to the glassy state is to cool, or quench, the material at a
sufficiently high rate so that crystallization is suppressed, as has been known
since the pioneering work of Tammann [1,2]. In effect, as the temperature
is reduced on cooling, the rate of molecular configurational rearrangements
reduces dramatically until, at and below a certain temperature referred to
as the glass transition temperature I'g, the structural state becomes 'frozen
in' on the time scale of the imposed cooling rate. An important exception to
the need for a fast cooling rate is the class of glassy polymers whose
amorphous structures result from a steric hindrance to crystallization,
introduced by their atactic chemical structure; nevertheless, these materials
still display a freezing in of a glassy structure at a glass transition tempera-
ture which is dependent on the rate of cooling.
Besides cooling from the melt or liquid, other procedures can be used to
arrive at the glassy state. These include the application of high pressures,
vapour deposition, the sol-gel process, and chemical reactions such as the
crosslinking of thermosetting polymer resins. Whatever the technique of
vitrification, the glass thus produced is in a non-equilibrium thermodynamic
state, as was first demonstrated by Simon [3, 4] and which can be illustrated
by reference to Figure 3.1. On cooling from the melt, the volume (and the
enthalpy and entropy) departs from equilibrium in a temperature interval
around the glass transition temperature I'g, which depends on the cooling
rate, and asymptotically approaches a glassy structure with excess values of
volume, enthalpy and entropy. This is a purely kinetic effect, and results
from the increasing time scales for molecular motion as the temperature is
reduced, as mentioned above, resulting in a freezing in of the glassy
structure. The question of the existence and classification of an underlying
Introduction 87

c
,
/
/

,,
/

I
Ta

Temperature, T

Figure 3.1 Schematic illustration of the temperature dependence of the volume V


(or of the enthalpy H) of an amorphous material in the glass transition region. The
temperature ~ represents the crystallization temperature, had the crystalline state
been allowed to form, and is usually about 1.5 Tg . Two different cooling rates, q, and
q2' are seen to give rise to two different glass transition temperatures, Tiq,) and
T9(q2) , defined by the intersections of the equilibrium liquid line and the asymptotic
glassy line, and to two different states A and B at 7;,. Isothermal annealing of glass
B reduces the volume (and enthalpy), for example to state C, which is defined by the
fictive temperature Tf .

fundamental thermodynamic transition has long been debated, ever since


the famous Kauzmann paradox [5J demonstrated that the entropy of the
glass would be less than the entropy of the corresponding crystal, an
impossible situation, unless some kind of thermodynamic transition inter-
vened. Excellent discussions of these thermodynamic aspects may be found
in the early work of Davies and Jones [6,7J and Kovacs [8J as well as in
more recent reviews [9,10l
In practice, the relaxation behaviour in the glass transition region is
monitored by the isothermal decrease of the excess volume, or more
commonly of the excess enthalpy, from their initial values towards the
metastable equilibrium state at the annealing or aging temperature T.
(Figure 3.1). This 'unfreezing' of the frozen-in glassy state, at A or B in
88 Relaxation processes and physical aging

Figure 3.1 dependent upon the cooling rate, can take place isothermally
because time is now allowed for the severely retarded molecular motions to
occur. In fact, the isothermal approach to equilibrium is accompanied by a
further severe retardation, so that time scales for molecular motion again
become very long. As a result it is rare to be able to achieve an equilibrium
state more than, at most, 15 or 20°C below ~ within any reasonable
experimental time. The details of these structural changes and of their
impact on, in particular, mechanical properties are referred to as physical
aging, and will be discussed in later sections.
The notion that molecular mobility is frozen in at ~ is conceptually
appealing but is clearly an oversimplification. Whilst it is true that large
scale translational and rotational rearrangements are energetically not
available below ~, the complex structure of polymer molecules, indeed of
most glass-forming liquids, means that more localized molecular or atomic
rearrangements are possible below, and often well below, ~. These second-
ary relaxations, denoted {3, y, etc. as they appear in order of decreasing
temperature, show characteristic behaviours which are quite distinct from
that of the main glass transition, or oc-relaxation, and confer upon the glass
many interesting optical, electrical and mechanical properties. The origin
and interpretation of physical aging and of secondary relaxations in a wide
variety of glasses has increasingly become the subject of lively debate in the
last 10 years or so, and a series of workshops and discussion meetings has
recently been devoted to this topic [11-14].
Whilst the detection of the main glass transition by volume or length
dilatometry, or more usually by differential scanning calorimetry (DSC), is
very common and rather straightforward, the use of these techniques for
detecting and, to an even greater extent, investigating secondary relaxations
is much more limited. There are some good dilatometric [e.g. 15-21]
examples of the identification of a {3-transition in a variety of amorphous
polymers, by a change in the slope of a volume-temperature plot or by
measuring isothermal contraction rates, and even evidence for a low
temperature y-transition and a liquid-liquid transition above ~ for some
polymers [19]. On the other hand, DSC has not yet been reliably reported
to have detected any {3-transition; interestingly, though, adiabatic calori-
metry was recently reported to have detected a {3-transition in an inorganic
silver iodophosphate glass [22], resulting from the freezing in of the
positional arrangement of Ag + ions.
Instead of using dilatometry or calorimetry, therefore, investigation of
secondary relaxations is made mainly through dielectric and dynamic
mechanical spectroscopy. The classic review of this subject by McCrum,
Read and Williams [23] has recently been republished, and remains a prime
source of information.
Structural relaxation in the glass transition region 89

3.2 STRUCTURAL RELAXATION IN THE GLASS


TRANSITION REGION

3.2.1 GENERAL ASPECTS


Figure 3.1 shows the existence above the glass transition temperature of an
equilibrium, which is often referred to as the melt, liquid-like or rubber-like
state. Although there is continuing debate about the existence of transitions
above I'g, for example the so-called liquid-liquid transitions, it is an
experimental fact that a reproducible state, defined for example by the
specific volume, can be attained in this temperature range provided that
other effects such as crystallization are avoided in systems for which these
can occur. The starting point for any study of structural relaxation in the
glass transition region should always be, therefore, to equilibrate the
material above I'g. In the simplest case, the structural relaxation process can
be initiated by cooling, preferably at a controlled rate, to an aging tempera-
ture 1',. at which the state of the sample (e.g. at A or B in Figure 3.1) will be
determined only by the cooling rate. Thereafter, during isothermal relax-
ation at 1',., the state of the glass will approach that of equilibrium, defined
uniquely by the temperature and pressure. This approach to equilibrium can
be followed either dilatometrically or calorimetrically, and the resulting
respective volume and enthalpy relaxation behaviours are considered separ-
ately below.
The bulk response of the glass determined in either of these ways can give
considerable insight into the overall relaxation process, but provides limited
information about the microscopic molecular motions contributing to the
relaxation. The complementary information available from scattering and
spectroscopic techniques can be revealing, and is briefly reviewed.
Finally in this section, histories more complex than the simple quench
introduced above and illustrated in Figure 3.1 are considered. In particular,
the combined effects of pressure and temperature histories can lead to rather
complex responses.

3.2.2 VOLUME RELAXATION


The classic study of volume relaxation in amorphous polymers was made
by Kovacs over 30 years ago [8], and demonstrated unequivocally the
important aspects of structural relaxation. These aspects relate not only to
volume but also to enthalpy relaxation, and furthermore are found to be
quite general to a very wide range of glass-forming systems. An illustration
of the typical isothermal volume contraction curves that are obtained at
various temperatures 1',., following the cooling stage to either of states A or
B in Figure 3.1, is shown for atactic polystyrene in Figure 3.2.
90 Relaxation processes and physical aging

0·004

atactic PS
10soC\T

aGQ

,
aa
Ga

Ga
a.
a GG
aG
10 100

Figure 3.2 Dilatometric isothermal volume contraction data for atactic polystyrene
(anionic fraction Mw ~ 160000) following a quench from equilibrium at 105°( to
the aging temperatures Ta indicated. The parameter b is the relative departure from
equilibrium, defined by equation 3.1. The initial time t i , of the order of 36 s, is the
time required to establish thermal equilibrium in the 8ekkedahl-type mercury-in-glass
dilatometer. The arrows indicate the initial values of b at time t = t i •

The ordinate lJ is the excess volume relative to the volume in equilibrium


Voo , at the same temperature:

(3.1)

and for an instantaneous quench through a temperature change Ll T would


take the value Ll()( Ll T, where Ll()( is the increment in thermal expansion
coefficient at the glass transition, Ll()( = ()(l - ()(g, between glassy and liquid-
like values «()(g and ()(l respectively). For this polystyrene, Ll()( '" 3.4 X
10- 4 K -1 [24], and it can thus be seen that the quenches to initiate the
volume contraction isotherms in Figure 3.2 are far from instantaneous. This
is one of the problems of quantitative interpretation of such data.
Two important features of structural relaxation are evident from these
data. First, the approach to equilibrium (lJ = 0) is much slower than
exponential, as can be seen from the very shallow slopes of the inflectional
tangents. This non-exponentiality is universally observed in glass-forming
Structural relaxation in the glass transition region 91

0·0010

0'0005

c.Q
Q)
E
:::J
g -0'0005
l!l
~
~
.~
iii
Q)
II:

0'1 10

Figure 3.3 Dilatometric isothermal volume expansion data for atactic polystyrene at
e
1000 following up-quenches from volumetric equilibrium states at 92.5, 95.0 and
97se as indicated. The parameters band ti have the same meaning as in Figure
e
3.2. Also shown for comparison is the isothermal contraction isotherm at 1000 after
a down-quench from equilibrium at 10Soe (one of the curves shown in Figure 3.2).
The arrows indicate the initial values of b at time t = t i•

systems. Second, the time required to achieve equilibrium increases dramati-


cally as T" reduces; here, for example, volumetric equilibrium is established
within approximately 100 h down to only 10°C below the initial temperature
of lO5°C.
If sufficient time is allowed for equilibrium to be reached at some lower
temperatures, it is possible to study an interesting experimental counterpart
to the isothermal volume contraction data shown in Figure 3.2, namely
isothermal volume expansion. An example of such expansion isotherms for
the same polystyrene sample is shown in Figure 3.3.
The important feature shown by these data, noted many years ago by
Kovacs [8], is the asymmetry of the responses in isothermal contraction and
expansion. This is most clearly seen by comparing the contraction isotherm
in Figure 3.3 with the expansion isotherm from 97SC, for both of which
92 Relaxation processes and physical aging

the initial departures are approximately equal and of opposite sign. In


general, the contraction isotherms appear self-retarding while the expansion
isotherms are autocatalytic. This aspect is often referred to as non-linearity
of the response.
The characteristic features of structural relaxation illustrated by the data
in Figures 3.2 and 3.3, together with some other features (most notably the
spectacular 'memory effects' [8] observed following complex thermal his-
tories), have shown conclusively that the volume relaxation process in the
glass transition region is both non-exponential and non-linear, and any
relevant theory of structural relaxation must include both of these aspects.
The non-exponentiality may be treated in one of two ways. The relaxation
may be considered as an inherently non-exponential process, usually involv-
ing the stretched exponential relaxation function cf>(t) associated with the
names Kohlrausch-Williams-Watts [25,26]:

cf>(t) = exp[ -(t/r)P] (3.2)

in which r is the relaxation time and P(O ~ P~ 1) is a stretching exponent.


Alternatively, the relaxation may be described in terms of a discrete
distribution of relaxation times [27] such that the departure from equilib-
rium (j may be considered as the sum of N elementary contributions:

(3.3)

with each element (\ being associated with a relaxation time rio


The non-linearity may be introduced in a variety of ways. In his
pioneering work half a century ago, Tool [28] recognized the importance of
this concept, and coined the term 'fictive temperature' I; as a parameter
to characterize the structure of the glass. The definition of I; is illustrated
with respect to the glass in state C in the schematic volume-temperature
diagram in Figure 3.1: it is the temperature at which the glass would
apparently have its equilibrium volume if it were instantaneously removed
to that temperature. For any glass formed by cooling at constant rate, the
fictive temperature initially is equal to the glass transition temperature
corresponding to the cooling rate. On aging isothermally, I; reduces until,
in equilibrium, it is equal to the aging temperature T.. This concept is often
used to define the temperature and structure dependence of the relaxation
time, and hence introduce non-linearity, through the empirical Tool-
Narayanaswamy-Moynihan equation [29,30]:

A [XAh* (1 - X)Ah*]
r = exp RT + R I; (3.4)
Structural relaxation in the glass transition region 93

in which the parameter x (0 ~ x ~ 1) defines the relative contributions of


temperature and structure to the relaxation time [31], and is often known
as the non-linearity or Narayanaswamy parameter, and Ah* is the apparent
activation energy for structural relaxation.
An alternative expression based upon the relative excess volume b
(equation 3.1) has been used in the so-called KAHR model [27] for
structural relaxation, which refers to a discrete distribution of relaxation
times !i:
!i = !irexP[ -8(T - 7;)] exp[ -(1 - x)8b/A~] (3.5)
In this expression !ir is the value of !i in equilibrium at a reference
temperature 7;, and 8 is a temperature coefficient related approximately to
the apparent activation energy Ah* by
Ah*
8 ~ RT2 (3.6)
g

An important point to note in equation 3.5 is that the individual relaxation


times !i depend upon the global value of b, and not upon the elementary
values bi .
These expressions for the relaxation time (equations 3.4 and 3.5) can be
combined with a simple constitutive equation defining the rate of approach
to equilibrium as proportional to the departure from equilibrium:
db b db i bi
-= -- or - = - - (3.7)
dt ! dt !i
to provide a complete description of the volume relaxation behaviour.
Whilst individual isotherms can be described in this way with great
precision, sets of isotherms such as those illustrated in Figures 3.2 and 3.3
are seen to deviate in a systematic way from the theoretical predictions
[8,27]. One of the most sensitive ways of presenting the data in order to
illustrate these discrepancies is by means of the 'effective relaxation time',
!eff' defined by [8]
1 1 db
(3.8)
b dt
Comparison with equation (3.7) shows that !eff is the instantaneous or
effective value of the relaxation time as it changes throughout the course of
the isothermal relaxation, and which can be evaluated from the time
evolution of b. The data for polystyrene shown in Figures 3.2 and 3.3,
together with other isothermal relaxation data not shown there, are repre-
sented in Figure 3.4 in the form of log !eff as a function of b.
94 Relaxation processes and physical aging

100

92·5

3
105
en
"oc:
&len
.;; 4

-2 o 2 6
Relative excess volume,6x10 3

Figure 3.4 Dependence of log (effective relaxation time) on relative excess volume
for atactic polystyrene. Curves to the left of the central line represent expansion
isotherms, and those to the right represent contraction isotherms, each at the
temperature r.
(OC) indicated in the adjacent box. The initiation of each isotherm
was effected by a quench from an equilibrium volumetric state at the temperatures
indicated against the individual curves.

The classic example of this form of data representation is the work of


Kovacs on polyvinyl acetate [8], but the present data for polystyrene shows
the same remarkable feature, namely the 'T eff paradox'. The paradox can be
seen from a consideration of the equilibrium relaxation times at the aging
temperatures of 100 and 105°e: it appears that the relaxation time in
equilibrium is not uniquely defined by the temperature, as it should be for
any thermorheologically simple system, and as equations 3.4 and 3.5 in
particular would predict, but depends on the path by which the equilibrium
state was reached. There are few such data in the literature, and the
measurement of (j becomes increasingly imprecise at very small values (i.e.
Structural relaxation in the glass transition region 95

very close to equilibrium), but the case appears strong for the real existence
of a so-called 'expansion gap'. The only approach that to date has success-
fully described this behaviour is the coupling model [32,33], though
reasonable agreement can be achieved by the Robertson-Simha-Curro
(RSC) model [34].
The essence of the coupling model is that the relaxation of a 'primitive
species', the fundamental relaxing unit, does not occur in isolation but is
coupled to its surroundings, with the strength of the coupling measured by
a parameter n (0:( n :( 1). Thus the concept of cooperativity of molecular
processes at the glass transition, which is fundamental to the KAHR [27]
and RSC models [34], and also to the Adam-Gibbs model to be discussed
below, arises naturally in the coupling model. One of its supposed strengths
is that it leads directly to a relaxation function of the K WW stretched
exponential form [35], with f3 = 1 - n (equation 3.2). One can therefore
identify strongly coupled systems as having small values of (3.

3.2.3 ENTHALPY RELAXAnON


Dilatometric experiments offer the advantage that the structural relaxation
process can be followed throughout its duration, for example in the
isothermal volume contraction and expansion experiments illustrated by the
results in Figures 3.2 and 3.3. The situation is not so simple in enthalpy
relaxation: first, the isothermal relaxation of enthalpy from positive values
of excess enthalpy, equivalent to isothermal volume contraction, is not
usually measured continuously, but discrete values of the excess enthalpy
are inferred from differential scanning calorimetry (DSC) scans at constant
heating rate; and second, not only is the same true for isothermal enthalpy
relaxation from negative values of excess enthalpy, equivalent to isothermal
volume expansion, but such experiments are very rarely performed. This
latter point in particular can have important consequences, as the kinetics
of the structural relaxation process from positive departures does not, in
itself, evidence the non-linear aspect of the behaviour which is so clearly
manifest by a comparison of the volume expansion and contraction iso-
therms. Overlooking the non-linearity inherent in the kinetics can result in
a misleading interpretation of the data, and particularly in the evaluation of
the stretching exponent f3 [36].
The usual procedure for studying isothermal enthalpy relaxation is to
scan in the DSC at constant heating rate from the initial annealed glassy
state (e.g. C in Figure 3.1) until equilibrium is established at a temperature
above ~. A second scan from an initial unannealed state obtained immedi-
ately after the cooling stage (e.g. B in Figure 3.1) provides a reference from
96 Relaxation processes and physical aging

Tp

40 50 60 70 80 90 100
Temperature (OC)

Figure 3.5 DS( heating scans on poly(propylene isophthalate) (M n = 11 000) at


10 0 ( min -1 after annealing at Ta = Tg - 15°( = 60 0 ( for 24 h (curve 1) and zero
hours (reference scan, curve 2). Tp is the temperature of the endothermic peak.
(Redrawn from reference 37.)

which to evaluate the enthalpy loss on annealing at T. (e.g. between states


Band C in Figure 3.1). An example of such DSC scans on annealed and
unannealed glasses is shown in Figure 3.5 for the particular case of a linear
polyester, poly(propylene isophthalate) [37]. The enthalpy loss between
states Band C (Figure 3.1) is found from the difference in areas under the
two curves in Figure 3.5. From a series of similar experiments, the depend-
ence of enthalpy loss on aging time may be found, and the results for the
same linear polyester are shown in Figure 3.6.
A comparison of Figure 3.6 with the volume relaxation data shown
in Figures 3.2 and 3.3 reveals that the measurement of isothermal
enthalpy relaxation by DSC is subject to significantly more scatter than
is volume relaxation; this applies not only to the data shown here, but is
generally observed. Nevertheless, it is possible to identify the same import-
ant features of non-exponentiality and non-linearity in enthalpy relaxation,
and these can be described theoretically in an analogous way to volume
relaxation: the former through the stretched exponential decay function
(equation 3.2) or by a discrete distribution ofrelaxation times, and the latter
by equation 3.4 or a modified equation 3.5:
(3.5a)
in which AC p is the difference between the specific heats of the liquid and of
Structural relaxation in the glass transition region 97

3
o

O~--~--~--~--~ __- L_ _-L__ ~ __~


-1 o 2 3

Figure 3.6 Dependence of enthalpy loss on log (aging time) for poly(propylene
isophthalate) aged at Ta = Tg - 15°( = 60°C. (Redrawn from reference 37.)

the glass (L\C p = Cpl - Cpg), and (5H is the excess enthalpy:
(5H = R - Roo (3.9)
where Roo is the enthalpy in equilibrium at r;..
Another, rather ~ifferent, approach follows from the original model of
Gibbs and DiMarzio [38], which requires the macroscopic molar configur-
ational entropy of the melt Sc to reduce to zero at a thermodynamic
transition temperature T2 (the Kauzmann temperature) below I'g. By mak-
ing use of the concept of 'cooperatively rearranging regions', structural units
in which configurational changes can take place without affecting their
surroundings, Adam and Gibbs [39] derived an expression for the relax-
ation time:

(3.10)
98 Relaxation processes and physical aging

where NA is the Avogadro constant, s~ is the configurational entropy of the


smallest cooperatively rearranging region, t'1J1 is the energy barrier to
cooperative rearrangement of monomer segments, k is the Boltzmann
constant and Sc is given by

S =
c
IT t'1CT' p dT' (3.11)
T2

The temperature dependence of t'1C p will determine the form of the express-
ion for Sc, and hence for the relaxation time. Hodge [10] has shown, from
an analysis of the compilation of data by Mathot [40], that t'1Cp for most
polymers has a temperature dependence lying between a constant value

t'1C p = C = constant (3.12)

and a hyperbolic relationship:

(3.13)

where C is the value of t'1C p at Tz. Inserting these equations for t'1C p into
equation 3.11 leads respectively to the following expressions for the relax-
ation time:

(3.14)

for constant dC p ' where

(3.15)

and

r = Allexp[-Q-] (3.16)
T- Tz
for the hyperbolic dependence of t'1C p • Equation 3.16 is the Vogel-Tam-
mann- Fulcher (VTF) expression for the temperature dependence of r, and
equation 3.14 is a very close approximation to it.
The extension of these ideas to the glassy state [10] is usually accom-
plished by assuming that the macroscopic configurational entropy depends
not on the actual temperature T but on the fictive temperature 7;, which is
a simple descriptor of the structural state of the glass. Hodge has shown that
this leads to expressions for r in which there are clear contributions from
Structural relaxation in the glass transition region 99

both temperature and structure:

r = A1exp [ Q ] (3.17)
Tln(Tr/T2)
and

r -- A" exp [ T(l - QT2 /T;) ] (3.18)

for the constant and hyperbolic dependences of LlC p respectively, and has
denoted these as AGL (for Adam-Gibbs logarithmic) and AGF (for
Adam-Gibbs-Fulcher) forms [41].
The numerous expressions given above for the temperature and structure
dependence of the relaxation time(s) (equations 3.4, 3.5, 3.5a, 3.17 and 3.18)
can be shown [10,42] to be approximately equivalent in a small tempera-
ture interval around ~, and can also be related to the WLF equation [43]
and to the free volume theory of Doolittle [44] and Cohen and Turnbull
[45,46].

3.2.4 RELAXATION PARAMETERS FROM VOLUME


AND ENTHALPY DATA

A common objective of many researchers in recent years has been the


evaluation of the parameters describing structural relaxation in glassy
polymers, and this has met with mixed success. Because of the convenience
of commercially available DSC equipment, and the corresponding inconven-
ience of dilatometry, the vast majority of the data relate to enthalpy
relaxation, and an excellent and comprehensive review has recently been
given by Hodge [10]. Two different approaches have been used in the
evaluation of these parameters: the 'curve-fitting' and 'peak-shift' methods.
The curve-fitting method uses the appropriate constitutive equation (e.g.
equation 3.7) together with both non-linearity and non-exponentiality to
predict the response to any prescribed thermal history. This method has
been applied exclusively to DSC data, with the constant heating rate treated
as a sequence of instantaneous T-jumps followed by an isotherm. The set of
parameter values, in particular the apparent activation energy Llh*, the
non-linearity parameter x and the non-exponentiality parameter p, are
found by achieving a best fit of the theoretical model predictions to the
available experimental data.
The peak-shift method [47,48] examines the dependence of the peak
endotherm temperature I;, (Figure 3.5) in DSC, or the peak in the thermal
100 Relaxation processes and physical aging

expansion coefficient in dilatometry [49], on the experimental variables, and


in particular its dependence on the enthalpy lost (5H' or relative excess
volume lost (5, during isothermal aging at 1'... This dependence is written as
a normalized shift:

(3.19)

for cycles in which the cooling and heating rates, ql and Q2' and the aging
temperature 1'.. are all maintained constant; only the aging time varies. This
shift can be shown theoretically [47, 48] to have a strong dependence on the
non-linearity parameter x, following a 'master curve' which is essentially
independent of the form of the distribution of relaxation times or of the
non-exponentiality parameter p.
The mixed success of these approaches relates to two aspects in particular:
the ability or otherwise of the theoretical models to describe enthalpy
relaxation behaviour under a variety of experimental conditions using a
unique set of parameter values, and the agreement or otherwise amongst
results reported from different workers. The former aspect concerns the
curve-fitting method, which has been found to give inconsistent results on a
number of occasions. For example, O'Reilly et al. [50] observe a significant
variation in the value of x derived from DSC data for polycarbonate as a
function of both annealing time and temperature in the region close to 1'g
(Ta in the approximate range 120-140°C). Similar discrepancies have been
reported for polystyrene [51,52], poly(methyl methacrylate) [52-54] and
poly(oxy-2,6-dimethyl-1,4-phenylene) [52], which have raised serious
doubts about the validity of the theoretical models. This problem is
particularly evident when the relaxation is occurring far from equilibrium
[55], and is most commonly attributed to an inadequate description of the
non-linearity of the relaxation, especially when the Tool-Narayanaswamy-
Moynihan formalism (equation 3.5) is adopted. A possible explanation for
these discrepancies is that the basic assumption of thermorheological
simplicity is not valid. Thermorheologically complex models have been
proposed, such as Ngai's coupling model [32, 33, 35] or Rekhson's multi-
relaxation processes with distributions of energy barriers and Kauzmann
temperatures [56,57], but their advantages in describing and understanding
structural relaxation in glasses have yet to become clear.
The latter aspect of the mixed success of the theoretical models concerns
the variability in the parameter values obtained by different workers. As an
illustration, values of x, p and I1h*/R are collected in Table 3.1 for
polystyrene, atactic poly(methyl methacrylate), poly(vinyl chloride), poly-
carbonate, poly(vinyl acetate) and epoxy resins, and a comparison is made
Table 3.1 Comparison of values of x, f3 and Ah*/Robtained for several polymer glasses by the curve-fitting and peak-shift methods
Curve-fitting Peak-shift

Polymer x P A~ * (kK) Ref. x II ,1~* (kK) Ref.

PS 0.43 0.68 82.5 58 0.46 0.456<P<0.6 70 61,62


0.49 0.74 80 41 0.46" 49
0.44 0.55 76-110 59
0.23-0.34 0.58 126 52
0.52 0.80 53-71 60
PMMA 0.19 0.35 138 41 0.37 0.3 <P <0.456 105 65
0.25-0.4J 0.35-0.45 105 63
0.34 0.27 132 64
PVC 0.10 0.23 225 41 0.27 P<O.3 135 67
0.11 0.25 225 66
PC 0.19 0.46 150 41
PVAc 0.27 0.51 88 41
0.41 0.51 71 68
Epoxy, fully cured 0.42 0.3<P<0.6 132 69
0.44 70
Epoxy, 70% cured 0.41 0.3 < P < 0.456 74 71
Epoxy with reactive diluent 0.42 Pr::!O.3 96 72

·Values obtained from volume relaxation experiments by dilatometry.


102 Relaxation processes and physical aging

between the values found by curve fitting and by the peak-shift method.
Although the curve-fitting results show considerable scatter, and there are
rather few results obtained by the peak-shift method, it is nevertheless
possible to draw some conclusions and make some inferences regarding the
structural relaxation process.
The activation energies can be seen to be very large, often of the order of
1000 kJ mol- 1 (I:!h* / R = 120 kK), and are even significantly greater than
polymer bond dissociation energies, of the order of 350 kJ mol- 1 (I:!h* / R =
42 kK) for the C-C single bond [73]. The explanation for such high values
is usually attributed to a high degree of cooperativity necessary for the
molecular motions to occur. This idea of cooperativity is fundamental to all
the theoretical treatments discussed above: through the non-exponentiality
parameter [3, through the dependence of 'i in the KAHR model on the
global value of relative excess volume (j (equation 3.5), through the par-
ameter n in the coupling model or the equivalent in the RSC model, and
through the idea of cooperatively rearranging regions in the Adam-Gibbs
treatment. In fact, the prediction from the coupling model of a stretched
exponential decay function implies that strongly coupled systems have low
values of [3 ([3 = 1 - n), broad distributions of relaxation times and highly
cooperative molecular processes. Furthermore, the coupling model also
predicts that the apparent activation energy I:!h* is related to the funda-
mental activation energy Ea for molecular processes of the so-called primi-
tive species by the simple relationship [35]

(3.20)

Thus the smaller the value of [3, the wider the distribution of relaxation
times, the greater the degree of cooperativity, and the higher the apparent
activation energy for structural relaxation.
To a certain extent this is confirmed by the data in Table 3.1, where
the decreasing values of [3 in the order polystyrene, polycarbonate,
poly(methyl methacrylate), poly(vinyl chloride) are associated, on the whole,
with corresponding increases in the apparent activation energy. Indeed,
strong correlations between the three parameters [3, x and I:!h* have been
reported by Hodge [10,41,74]. These may be summarized as an inverse
correlation between x and I:!h*, such that the product xl:!h* remains
relatively constant in comparison with the variation in I:!h* alone, and a
direct correlation between x and [3. Clearly these two lead to an inverse
correlation between [3 and I:!h* such that the product Ea (equation 3.20)
remains rather constant.
Recently, attempts have been made to rationalize these observations on
the basis of the concept of 'strength' and 'fragility' of glass-forming melts.
Structural relaxation in the glass transition region 103

Strong
-5
<II
U
I:
a
o
:ll
.!:
A
~
V
o
ci
.3 -10

-15

o 0·2 0·4 0·6 0·8 1·0


Normalised reciprocal temperature. Tg/T

Figure 3.7 Classification of strong and fragile glass-forming systems in terms of the
deviation from Arrhenius behaviour at temperatures above Tg . The ordinate is the
average relaxation time on a logarithmic scale, and the abscissa represents the
reciprocal temperature normalized to Tg , such that it ranges between zero (at infinite
temperature) and unity (at Tg).

This concept was introduced by Angell [75] in terms of the temperature


dependence of viscosity, but may conveniently be represented in terms of the
average relaxation time (or), as illustrated in Figure 3.7. The upper (straight)
line in this figure corresponds to Arrhenius behaviour, with the relaxation
time reducing from 100 s at ~ to 10- 14 s at high temperature; this latter
represents the average period of vibration of the liquid lattice, and can be
deduced from far infrared studies. The lower curve corresponds to non-
Arrhenius behaviour, which may be described by, for example, the Vogel-
Tammann-Fulcher equation (equation 3.16). The term 'fragile' is used to
describe departure from Arrhenius behaviour, and can be seen to involve a
large apparent activation energy at ~, where the relaxation time changes
very rapidly as the glass transition is approached from the melt. This may
104 Relaxation processes and physical aging

be thought of as a rapid freezing in or blocking of molecular motions on


cooling. In contrast, strong glass-forming liquids display no such catas-
trophe on cooling.
It is interesting to examine some glass-formers which exhibit these
limiting behaviours of Arrhenius and markedly non-Arrhenius dependences.
The classical examples of melts showing strong behaviour are those inor-
ganic liquids with tetrahedrally coordinated network structures, such as
silica (Si0 2) and germania (Ge0 2 ), where the three-dimensional network is
believed to provide the strengthening effect. At the other extreme, orthoter-
phenyl appears at present to be the most fragile liquid, with the majority of
glass-formers in fact clustering towards this limit [75]. The linear amor-
phous polymers are no exception, and are found rather close to this extreme.
The inverse correlation between {3 and Ah* noted by Hodge [10,41,74]
implies that fragile behaviour is associated also with small values of {3, in
other words with highly cooperative relaxation processes. The observation
that most linear polymers can be classified as fragile glass-formers empha-
sizes again the important role of cooperativity in these systems.
A recent attempt [76] has been made to quantify these different behav-
iours through a parameter (m) termed 'fragility', defined as the limiting
slopes of the curves shown in Figure 3.7:

dIOg(r>1 (3.21)
m = d(~/T) T=T,

and previously known as the steepness index [77]. The minimum value of
m is mmin '" 16 and corresponds to the limit of strong behaviour illustrated
in Figure 3.7; increasing values of m are indicative of increasing fragility.
Table I in reference 76 places many polymer glass-formers at the head of the
list, with values of m approaching 200. These include poly(vinyl chloride)
(the most fragile, with m = 191), poly(methyl methacrylate), polystyrene
and polycarbonate, all with correspondingly low values of {3 (in the range
0.24 - 0.35). Interestingly, though, a crosslinked epoxy resin is found to be
very close (m = 151) to the most fragile of these systems, whereas one might
have anticipated that its three-dimensional network structure would have
introduced a strengthening in a similar way to that for the inorganic
network glass-formers such as silica (m = 20) and germ ani a (m = 20). The
results shown in Table 3.1 would seem to confirm the trend of increasing
fragility with the introduction of the chemically crosslinked network struc-
ture in epoxies; compared with the fully cured system, the activation energy
is significantly reduced in the partially cured (70%) resin and in the resin
with reactive diluent, for both of which the network structure is looser than
for the fully cured system.
Structural relaxation in the glass transition region 105

In contrast, the common observation that the non-linearity parameter x


and the apparent activation energy Ah* are inversely correlated [78J seems
more difficult to substantiate, in particular when these epoxy systems are
considered. According to this correlation, highly fragile systems would be
anticipated to have low values of x, and yet not only are the values of x for
the epoxies in Table 3.1 almost as large as any found for the linear
amorphous polymers, but they also appear essentially independent of the
changes introduced into the network structure. The interpretation of x as a
measure of fragility is therefore difficult to sustain. Rather, it has been
argued [79J that it more revealingly can be interpreted as a measure of the
continuity of the liquid-like structure as the glass transition is traversed on
cooling to form the glass. Thus, while f3 and Ah* represent measures of
liquid state properties above ~, x on the other hand is an indicator of
continuity in melt/glass behaviour at ~. Indeed, this is evident from the
expression for the relaxation time given by equation 3.4; the equilibrium
liquid (Tr = T) has activation energy Ah*, whereas the glass of fixed
structure (Tr = constant) has activation energy xAh*. A similar measure of
continuity of melt/glass behaviour may be afforded by AC p ; the greater the
discontinuity at the glass transition, the more restricted become the molecu-
lar processes in the glass, and hence the smaller their contribution to the
glassy specific heat capacity, with a corresponding increase in AC p • Thus, for
example, the increasing values of x as the AgI content is increased in silver
iodomolybdate glasses [79J are accompanied by decreasing values of AC p
as the melt/glass behaviour becomes more continuous.

3.2.5 MICROSTRUCTURAL ASPECTS


One of the most important aspects to emerge from the study of the
structural relaxation processes at ~ by volume and enthalpy relaxation is
the non-exponentiality, also interpreted as resulting from a distribution of
relaxation times. The physical meaning of this may be, for example, that a
distribution of lengths of molecular segments are involved in the relaxation
process, or that these segments relax in a variety of different environments.
Bulk responses, such as in volume and enthalpy relaxation, can, and indeed
do, show the existence of such distributions, but provide little insight into
the nature of their origin. Thus the study of the bulk relaxation behaviour
can be complemented usefully by microstructural techniques such as X-ray
scattering and probe spectroscopy.
The concepts of free volume [44-46J and of free volume distribution have
long been used to provide a physical picture of the relaxation behaviour.
The techniques of small angle X-ray scattering and of light scattering are
able to measure the mean square value of the thermal density fluctuations,
106 Relaxation processes and physical aging

which may be considered to be related to the distribution of free volume


[80]. For example, Curro and Roe [81] estimate a weight average 'hole'
volume of 0.105 nm 3 for poly(methyl methacrylate) at 106°C; this corre-
sponds to the volume of 0.74 monomer units. Above ~ the fluctuations
of the sample in equilibrium are determined by the thermal energy kT and
by the isothermal compressibility, whereas below ~,in the glassy state, they
are proportional to the thermal energy and to the compressibility in
equilibrium at the fictive temperature. Thus, on cooling through the glass
transition region, the fluctuations exhibit a transition at ~. With the
reduction in fictive temperature on aging at a temperature below ~, as
observed by volume and enthalpy relaxation, a corresponding reduction in
the density fluctuations would be expected. Surprisingly, however, it has
been found for polycarbonate and poly(methyl methacrylate) that the
density fluctuations are very insensitive to thermal history, and that they
change insignificantly during aging below ~ [81,82], even though the bulk
volume decreases measurably. Similar results for polystyrene [83] have led
to the suggestion that the bulk volume from dilatometric experiments yields
information about an average free volume while density fluctuations from
small angle X-ray spectroscopy reveal the distribution of free volume.
Methods other than aging for producing reductions in bulk volume lead
also to surprising results in terms of density fluctuations. Milller and
Wendorff [82] have shown that cold drawing of polycarbonate considerably
increases the absolute value of the thermal density fluctuations compared
with the isotropic sample, and that both the density fluctuations and the
bulk volume relax isothermally on aging at much faster rates than the
isotropic samples, despite the fact that the effect of cold drawing is to reduce
the volume by about 0.3%. The enthalpy relaxation behaviour is also
significantly enhanced in the cold drawn sample. One interpretation of these
results would be that the free volume is increased by the applied stress
despite the large reduction in overall volume, hence giving rise to increased
mobility.
Similar surprising results have been obtained by Song and Roe [84], who
obtained 'pressure-densified' samples of polystyrene by the application of
pressure above I'g, and then cooling under pressure to freeze-in the structure
before releasing the pressure. Small angle X-ray scattering intensities showed
that the degree of short-range ordering (on a scale of 0.5 nm and less) was
greatly reduced in the pressure-densified samples, even though they had a
much higher density than the ambient pressure-produced glasses. In other
words, the reduced specific volume in these densified samples is achieved
with a reduced rather than enhanced short-range ordering. In contrast, the
reducton in specific volume which occurs on isothermal aging of normal
Structural relaxation in the glass transition region 107

(ambient pressure) glasses is accompanied by an increased short range


ordering, as would be expected intuitively.
These results suggest that the simple and widely used concepts of free
volume and distribution of free volume need rather more careful consider-
ation than is often afforded. Indeed, this problem was highlighted many
years ago by Haward [85] in his consideration of the various possible
definitions of free volume.
Nevertheless, the concept continues to be popular, and has been used to
interpret the results of various spectroscopic techniques. Some particularly
interesting data are emerging from the use of photochromic probes and
labels. The absorption of light over a range of frequencies induces a
reversible photochemical trans~cis isomerization of a chromophore, which
can be detected by UV ~vis spectroscopy. This photoisomerization requires
a critical free volume in the immediate vicinity of the chromophore. Lamarre
and Sung [86] analysed the kinetics of the photoisomerization as the sum
of two processes: a fast process with rate constant A I and a slow process
with rate constant A 2 . The interpretation was that rapid isomerization
occurs when the free volume in the vicinity of the chromophore exceeds a
critical value, whereas the slow process occurs when the free volume is less
than this critical value. The fraction rt. of the fast processes is therefore
assumed to be the cumulative area under the free volume distribution curve
for free volumes greater than the critical size for that chromophore. By
photochromically labelling different sites on the polystyrene molecule (chain
end, side chain and chain centre) and comparing with free photochromic
probes situated in the regions between the polystyrene chains, Yu et al. [87]
were able to estimate free volume distributions in different sites, and to
follow their variation during physical aging following quenches from above
Y'g (Tg - lOO°C) to temperatures of 70 o e, 80 e and 90o e.
0

The rate constants Al and A2 differ by about two orders of magnitude,


but remain essentially constant with aging time and are approximately equal
for all samples. On the other hand, the fraction rt. of fast photoisomerization
processes is dependent on the labelling site and reduces as a function of
aging time, as illustrated in Figure 3.8. The reduction in rt. values with
increasing aging time, for all sites and for the free probe, implies a shift in
the free volume distribution curve towards smaller sizes as aging proceeds.
More detailed examination of the data in Figure 3.8, and of similar data
obtained for aging temperatures of 70 e and 90 o e, shows that the rate of
0

physical aging, as quantified by the slopes of the lines of rt. versus log t, is
greater for the chain end and side chain sites than it is for the chain centre
site or for the free probe (Table 3.2). More clearly, the rt. values decrease in
the order probe> chain end> side chain> chain centre.
108 Relaxation processes and physical aging

.• +.
1.0
III III
Free Probe

0.8
!i!
Ii! o+P+~c. .. Chain End

0.6 0
+ ~ ODD
a + + + + Ii! Side Chain

0.4

0.2

III Chain Center


a Dl-c'" lt~
0.0
1 2 3 4 5 6 7 8

log(Aging Time. 5)

Figure 3.8 Dependence of the fraction (I( of fast photoisomerization processes on


aging time for three labelled polystyrenes (chain end, side chain and chain centre)
and for the free probe in polystyrene. Squares (0) represent results obtained after
a quench from 120 to 80°C; crosses (+) represent a quench from 110 to 80°C.
(Reproduced from reference 87, with permission.)

These trends have been interpreted in terms of changes in free volume


distribution as illustrated schematically in Figure 3.9. The distributions are
indicated for 100 s after quenching to 90°C (denoted by q) and after
100000 s of aging at 90°C (denoted by a). The shaded areas represent those
portions of the free volume distribution with fractional free volumes greater
than the critical value Ie for photoisomerization. For the free probe, which
has (X values close to unity (Figure 3.8), nearly all the sites occupied have a
fractionl free volume greater than !C. In contrast, the chain centre labels
have (X values close to zero, and nearly all the sites are smaller in volume
than the critical size. The shift of the distribution to smaller free volume on
aging would therefore be expected to influence only slightly the (X values of
these two, the free probe and the chain centre site. On the other hand, the
chain end and side chain labels have (X values such that a significant
proportion of the distribution is influenced by the shift on aging.
These results have been modelled [87] using the Robertson-Simha-
Curro theory for relaxation processes [34], and very satisfactory agreement
obtained except for aging at 30°C below ~, i.e. at 70°C, where discrepancies
can already be noted in respect of the aging rate for chain end labels in
Table 3.2. Interestingly, this is the region, far from equilibrium, in which
Structural relaxation in the glass transition region 109

Table 3.2 Rate of physical aging (10 2dw'd 10g[t.lsJ) as


measured by the reduction in 0( as a function of log
aging time (ta) for polystyrene. The thermal history is a
quench from 110°C to the temperature indicated. Data
taken from reference 87.
Probe/label site 90°C 80°C 70°C
Probe 0.8 0.7 0.7
Chain end 1.7 1.7 0.8
Side chain 2.7 2.0 1.0
Chain centre 0.7 1.5 1.4

difficulty has been encountered in modelling the enthalpy recovery behav-


iour. Nevertheless, this appears to be a powerful technique for probing the
microscopic aspects of relaxation processes, at least close to ~, and this is
confirmed by the work of Torkelson and coworkers [89-92].
The complementary technique employed by these authors was to use
photo chromic probes which required different volumes for the photoisomer-
ization to proceed. These volumes range from 0.127 nm 3 for azobenzene to

III
a . aged
q . quenched

fC
Fraction of Free Volume

Figure 3.9 Schematic illustration of the distribution of free volume in the three
labelled sites (chain end, side chain and chain centre) and for the free probe in
polystyrene. The curves denoted by q represent the distributions soon after the
quench from above Tg to the aging temperature; the curves denoted by a represent
distributions after physical aging. The critical fractional free volume for photoisomer-
ization of the chromophore (azobenzene) is denoted ~, and has previously been
estimated as 0.038 [88]. (Reproduced from reference 87, with permission.)
110 Relaxation processes and physical aging

0.571 nm 3 for 4,4'-diphenylstilbene, as calculated from the van der Waals


area of a probe during geometrical rearrangement, and can be used to
characterize the free volume distribution in glassy polymers. For example,
it is estimated [89] that, in an unannealed polystyrene glass at 25°C, 90%
of the free volume exists in sites larger than 0.120-130 nm 3 , while no site is
larger than 0.400nm 3 • The median value of 0.260-0.280 nm 3 can be com-
pared with the volume of a polystyrene repeat unit, about 0.180 nm 3 .
The effects of physical aging on the free volume and its distribution have
been investigated by means of these probes in polystyrene [89,90], poly-
(methyl methacrylate) [91] and polycarbonate [92]. For polystyrene and
poly(methyl methacrylate), although there were differences in respect of
some of the details according to molecular weight and aging temperature,
the same general trend was observed for both, whereby physical aging
reduces the fraction of local free volume of large sizes more than it does for
the smaller free volume fractions. In other words, larger pockets of free
volume decrease in number more so than do smaller pockets, resulting in a
narrowing of the distribution of free volume on aging. Indeed, aging
polystyrene at 60°C was even found to increase the fraction of local free
volume between 0.166 nm 3 and 0.262 nm 3 [89]. For poly(methyl metha-
crylate), aging at 68°C for 100 h was found to reduce the contribution of the
largest free volume pockets (0.571 nm 3 ) virtually to zero. These observations
are consistent with the predictions of the multiparameter KAHR model
[27], in which the bulk specific volume may be considered to consist of a
distribution of elements covering a range of specific volumes. It is those
elements with the highest specific volume which relax the fastest, and hence
there is a change in the distribution of the elementary contributions, as there
is for the distribution of free volume pockets found by Torkelson and
coworkers [89-91]. However, this should not be confused with a change in
the shape of the distribution of relaxation times, which simply shifts to
longer times without change in shape, as required by the fundamental
assumption of thermorheological simplicity, as the global specific volume
reduces on aging [93]. Quantifying the rate of loss of volume from the larger
free volume sites shows it is an order of magnitude larger than the loss in
bulk volume found by dilatometry, implying that the redistribution of free
volume from larger to smaller sites occurs at a much faster rate than the net
loss of free volume during aging.
A comparison of the polystyrene and poly(methyl methacrylate) data
[89-91] also shows that poly(methyl methacrylate) has a broader distribu-
tion of free volume sizes than does polystyrene, for both annealed and unan-
nealed glasses. For example, no local free volume greater than 0.400 nm 3 is
detected in polystyrene, whereas volumes as great as 0.571 nm 3 are found in
poly(methyl methacrylate). This observation is consistent with the broader
Structural relaxation in the glass transition region 111

distributions of relaxation times found for enthalpy relaxation in poly-


(methyl methacrylate) compared with polystyrene, as evidenced by the
values of the parameter f3 being smaller in the former (Table 3.1 and
references quoted therein).
In contrast, the local free volume distribution in polycarbonate is broader
than that for poly(methyl methacrylate) when determined from the photo-
isomerization results [92], which does not correlate with the value of f3 from
Table 3.1, which is greater for polycarbonate (i.e. narrower distribution)
than for poly(methyl methacrylate). However, it should be noted that as the
activation energy for polycarbonate is greater than that for poly(methyl
methacrylate), one would anticipate from the usual correlations that f3
would be smaller for polycarbonate. The observation that physical aging in
polycarbonate affects all of the local free volume sizes almost equally also
marks a currently unexplained difference with respect to the results for
polystyrene and poly(methyl methacrylate).
A similar technique to the use of photo chromic probes and labels is
fluorescence spectroscopy. In this technique a suitable fluorescent molecule
is used as a probe; when the non-radiative bond rotational motion within
the molecule is hindered by lack of mobility, the preferred mechanism
becomes that of photon emission and fluorescence. Thus the fluorescence
intensity is a measure of the reduction in molecular mobility, or reduction
in free volume, brought about for example by physical aging.
In his early work Loutfy [94] proposed a simple expression for the
fluorescence intensity I in terms of the free volume Vr of the sample:

(3.22)

where Vm is the critical volume for the relevant molecular motion of the
probe, lois the intensity for unhindered motion of the probe and b is a
constant dependent upon the polymer and probe types.
There are some remarkable similarities between the relaxation kinetics
identified by fluorescence spectroscopy and the bulk relaxation behaviour
observed in dilatometry or calorimetry. For example, as poly(vinyl acetate)
is cooled at constant rate from equilibrium at 50°C (i.e. above 1'g), the
fluorescence intensity first increases with a constant slope and then shows a
transition to a smaller slope in the glassy region, with a transition tempera-
ture of 28°C, close to that determined by differential scanning calorimetry
[95]. For samples quenched from above 1'g to aging temperatures below 1'g,
the intensity is seen to increase with aging time. A typical illustration of
this is shown in Figure 3.10, though it would have been preferable had
the original data been displayed as a function of aging time on a logarith-
mic rather than linear scale. The relaxation kinetics determined by these
112 Relaxation processes and physical aging

300~-------------------------------------,
>-
~ 290
z
UJ
.
..,~.
. ~' II,.m::.m~
f-
~ 26
UJ
u
z
~ 270
<f)
UJ
a
g 260
oJ
IL

2500~----~5~
O------~
IO~0------~
I50
~----~2~OO~----~250

AGEI NG TIME (m,nults)

Figure 3.10 Dependence of fluorescence intensity on aging time for poly(vinyl


acetate) following a quench from equilibrium at 60 0 ( to the aging temperature of
25°C. (Reproduced from reference 95, with permission.)

spectroscopic data are compared with those derived from the aging of the
mechanical stress relaxation response, and a good one-to-one correspon-
dence is found, with only small deviations occurring for the lowest aging
temperature (20.3°C).
Even more remarkably, the fluorescence intensity is found to mirror the
bulk volumetric relaxation kinetics when more complex thermal histories
than a simple quench from above to below Tg are employed. Defining a
relative excess fluorescence intensity DF in an analogous way to the defini-
tion of relative excess volume D(equation 3.1),

(3.23)

where 100 is the intensity in equilibrium, Royal and Torkelson [96] find,
again for poly( vinyl acetate), all of the characteristic features of bulk volume
relaxation: an asymmetric response is seen for initial departures of equal and
opposite sign, illustrated by the results shown in Figure 3.11; memory effects,
involving the spontaneous departure of DF from an apparent equilibrium,
are seen when the sample temperature is increased from an initial non-
equilibrium state; and the equivalent of the expansion gap or r eff paradox
in volume recovery (cf. Figure 3.4) is seen for the isothermal approach to
equilibrium from negative values of DF , visible in fact from a close inspection
of Figure 3.11.
Furthermore, the fluorescence intensity isothermally increases approxi-
mately linearly with log time following a simple down quench, until the
Structural relaxation in the glass transition region 113

0.06 iii II" , I I i ,III

0
0.04
0
l!.
l!.
0
0.02 l!. 0
l!. 0
l!. 0
l!.l!.~
Of 0.00
tj)
000
0 00 0
-0.02 0 0 ~O
0
0
00
0
0
-0.04 0
0

-0.06
10 100
Time (mins)
Figure 3.11 Isothermal recovery at 35°( of relative excess fluorescence intensity, OF
(equation 3.23). following quenches from equilibrium at the following temperatures:
0, 55°(; /:0., 40°C; 0, 32.5°(; D, 30°C. (Reproduced from reference 95, with
permission .)

equilibrium value Ioc is approached at each aging temperature. This is


strikingly similar to the enthalpy (and volume) relaxation behaviour, with
the time scales for reaching equilibrium at 40°C and 37SOC being very
similar in both fluorescence and enthalpy relaxation for poly(vinyl acetate)
[96]. Such linearly increasing fluorescence intensity with logarithmic time
during aging has been reported also for polystyrene, poly(methyl metha-
crylate) and poly(isobutyl methacrylate) [97]. An attempt has been made to
correlate these fluorescence aging rates with aging rates determined
dilatometrically [97J, both of which are found to depend upon the aging
temperature, passing through a maximum as the aging temperature is
reduced. As Royal and Torkelson point out, the remarkable similarities, in
many respects, of physical aging detected by fluorescence spectroscopy and
dilatometry (and calorimetry) indicate that bulk relaxations are a direct
manifestation of molecular scale processes.
Another, increasingly powerful, technique for the detection of these
molecular scale processes is positron annihilation lifetime spectroscopy
(PALS). In this technique, high energy positrons (e+, the antiparticle of the
electron) from a radioactive source such as 22Na or 58CO enter the glassy
polymer sample, and may penetrate by as much as 2 mm. Within this
114 Relaxation processes and physical aging

distance, the positrons are slowed down and annihilated in one of various
possible ways.
Most simply, the positron may annihilate directly with an electron to
form a y-ray. Alternatively, the positron may capture an electron to form a
bound state known as a positronium atom (Ps), with an atomic radius
similar to hydrogen. Only a fraction of the positrons form the bound state,
for which there are two possibilities: para-positronium (p-Ps) is formed by
the combination with an electron of opposite spin, which may subsequently
self-annihilate with the emission of two y-rays, while ortho-positronium
(o-Ps) is formed by the combination with an electron of parallel spin, with
subsequent annihilation by an electron with anti parallel spin to form three
y-rays in what is called the 'pick-off' process.
The lifetimes corresponding to these three annihilation mechanisms are
very different. The OOps lifetime, '3' is the longest at 142 ns in vacuum,
compared with only about 0.12ns ('2) for poPs and less than 0.5ns ('1) for
free positron annihilation. In condensed matter, however, the 0- Ps lifetime
'3 is considerably reduced, to only of the order of a few nanoseconds.
The positron annihilation lifetime spectrum is therefore commonly con-
sidered to consist of three components of different time scales. With regard
to physical aging, it is the OoPs lifetime that is the most important, the
shorter lifetime components being rather independent of structural changes
in the material [98]. The OoPs is preferentially formed or trapped in regions
of low electron density, generally interpreted as regions of local free volume
(or 'holes'). The annihilation rate or lifetime of 0- Ps is a very sensitive
function of the hole size or of the free volume associated with these sites. In
addition, the intensity 13 of the 0- Ps component will be a measure of the
number of these sites. Thus both free volume and free volume distribution
are, in principle, available from this technique of PALS.
The PALS technique has proved extremely useful in studies of the effects
of pressure on polypropylene and epoxy resins by Jean and coworkers
[99-102]. These authors found that the hole size, as measured by '3' was
reduced as the pressure increased; for example, in polypropylene '3 de-
creased from 2.3 to 0.7 ns as the pressure was increased from zero to
14.7 kbar [100], and from 1.6 to 0.5 ns for the same pressure range in epoxy
[101]. In addition, increasing pressure also resulted in a narrowing of the
hole size distribution.
The technique would therefore seem to offer exciting possibilities for the
study of physical aging of polymers, being able in principle to probe hole
sizes, usually assumed to be spherical, with radius in the range of 0.05 nm
to an upper limit of the order of 1.0 nm [100]. Interestingly, early results
show much less sensitivity of '3 with respect to aging than for the effects of
Structural relaxation in the glass transition region 115

pressure. For example, even though the temperature dependence of !3 for


poly(vinyl acetate) displays a rather well defined Tg ( = 32°C) from the
intersection of glassy and liquid-like asymptotes [103, 104], with a variation
from about 2.4 ns to about 1.8 ns over a temperature range from 80°C down
to - 30°C, the effect of aging at 20°C, some 12°C below ~, is a reduction
in !3 from about 1.90 ns to only about 1.88 ns over a period of 100 h. (It is
interesting to note that the plot of ! 3 versus temperature shows a complex
behaviour in the ~ region, identical to that found for the same material in
fluorescence spectroscopy [95].) On the other hand, the intensity 13 appears
much more sensitive to aging. Similar results are reported also for both
polystyrene [105] and polycarbonate [106-108], from which the conclusion
may be drawn that the consequences of physical aging and of temperature
change are qualitatively different; with reducing temperature the hole size
decreases, whereas as aging proceeds the number of holes decreases with
apparently little reduction in hole size.
Such interpretations of the data must be made with extreme care,
however, in the light of some recent observations on the complicating effects
of radiation from the radioactive source. Such effects have been observed in
a number of ways, but all lead to the same conclusion that a reduction in
13 may be induced by radiation. O'Connor et al. [109]. for example,
examined the response of polystyrene to PALS in respect of the reduction
of 13 with aging time for samples with very different aging histories. In one
experiment the response of two almost identical samples was. compared; one
sample was tested shortly after a quench from above to below ~, the other
after an aging period of 2 years. For both samples, 13 decreased with time,
over a period of about 80 h, by approximately the same amount and at
approximately the same rate. Kluin et al. [110] compared the response of
'as-received' (i.e. of unknown but presumed long aging time) and rejuven-
ated polycarbonate samples, and found that both exhibited similar decays
of 13 with aging time up to 60 h. Li and Boyce [111] adopted a different
approach in which they analysed successively by PALS the two sides of the
same sample, only one side being exposed to radiation at any time. For
polystyrene and polycarbonate, these authors found that, when 13 is
measured as a function of time for irradiation of the second side of the
sample, the same rate of decay from the same initial value is obtained.
Furthermore, when a stronger source (21 !lCi in place of 7 !lCi) is used, the
rate of decrease of 13 is faster. In contrast, when these PALS experiments
were repeated on poly(methyl methacrylate), no decrease of 13 with aging
time was observed.
All of these results indicate that the changes in 13 often reported to result
from physical aging may well be due in part, and possibly in large part, to
116 Relaxation processes and physical aging

radiation effects in the sample. The presence of an external electric field has
been shown to reduce the probability of positronium formation in several
polymers, and some of the reduction in 13 is therefore probably a result of
the build-up of an electric charge during the PALS measurements [112]. It
is noteworthy that poly(methyl methacrylate), for which Li and Boyce found
no significant decrease of 13 with aging time, is one polymer for which
positronium formation appears unaffected by the presence of an electric
field, possibly because it is strongly polar [112]. This problem of extraneous
irradiation effects in PALS could be resolved, at least partly, by removing
the sample from the radiation source when measurements are not being
made. The interpretation of intensity data from PALS should therefore be
made after careful consideration of whether or not the results were obtained
with such intermittment exposure of the sample.
3.2.6 COMBINED TEMPERATORE AND PRESSURE HISTORIES

It is well known that glasses may be formed not only by cooling from the
melt, but also by the isothermal application of pressure, or indeed by many
combinations of temperature and pressure histories. For reasons of experi-
mental facility, most studies of glassy relaxation behaviour have been made
isothermally following temperature jumps at ambient pressure. There have
been, however, some studies of isobaric glassy state relaxation following
pressure jumps, notable examples being the early work of Rehage and
coworkers [9, 113-116] and more recently that of Tribone et al. [117].
Qualitatively similar results are obtained from these pressure jump experi-
ments as those obtained using the more usual temperature jumps. In
particular, the features of non-linearity, non-exponentiality and memory
effects are all exhibited, indicating that the structural relaxation process is a
consequence of the non-equilibrium glassy state, which may be induced
equally well by either pressure or temperature changes.
In fact, a thermodynamic consideration of the freezing-in process at the
glass transition induced by both pressure and temperature changes leads to
some interesting observations. For example, continuity ofthe state functions
volume (V) and entropy (S), assumed to depend only upon temperature T
and pressure P both above and below the glass transition, leads to the
famous Ehrenfest relations:
dT .1"
(3.24)
dP .11X
and

dT TV.11X
= (3.25)
dP .1Cp
Structural relaxation in the glass transition region 117

where AK is the difference between the isothermal compressibilities of the


liquid and of the glass.
Equating these two expressions, one can write

(3.26)

where the term on the left-hand side is known as the Prigogine-Defay ratio.
Davies and Jones [6] were the first to show that for glass-forming materials
this ratio was not in fact equal to unity, but always exceeded it. The
explanation for this discrepancy lay in the inability of the so-called single
parameter model to describe the freezing-in process under these conditions,
and led to the idea of multiple 'ordering parmeters'. This concept may be
considered to be equivalent to that of a distribution of relaxation times [27],
and it follows from this that equation 3.26 is replaced by the inequality

AKACp ~ 1 (3.27)
TVAct 2 '"

One consequence of the existence of multiple ordering parameters is the


memory effect in glasses, by which the response of the glass is a function of
the whole of its previous thermal and mechanical (pressure) history since its
most recent equilibrium state. In the present context, this implies that the
freezing-in process by different pressure-temperature paths will lead to
different glassy structures and hence different glass transition temperatures.
The evaluation of d ~/dP will therefore depend upon how the glass is
formed, and equations 3.24 and 3.25 will not in general give the same result.
This can be well illustrated by the careful dilatometric measurements made
on poly(vinyl acetate) under pressures up to 800 bar by McKinney and
Goldstein [118], some of whose results are reproduced in Figure 3.12.
Three different glass formation histories were employed. In the first
(Figure 3.l2a), the glass is formed by cooling isobarically from the melt at
various pressures, using a cooling rate of 5°C h - 1. In the second (Figure
3.l2b), the glass was formed by cooling at 5°C h -1 at atmospheric pressure
(P = 0 bar), and then volume changes in the glassy state were measured by
temperature-pressure jumps applied to the glass. The third history (Figure
3.12c) is similar to the second except that the glass is formed by cooling
at 800 bar rather than at atmospheric pressure. In each case, the glass
transition temperature is identified by the intersection of the glassy and
liquid-like linear volume-temperature relationships, and is denoted by ~(P)
for the formation history of Figure 3.l2a, and by ~*(P,O) and ~*(P,800) for
the formation histories of Figures 3.12b and 3.12c, respectively. It is clear
that ~(P) appears to be uniquely defined by P, as it is a function of only one
118 Relaxation processes and physical aging

P (bars)
.89 0

.88
(a)

- 0 /)
.......
..,
.87

.86
00

E
.....,
u
v .85
E
:3
"0 .84
:>
u
to:: .83
'5
v
0..
en .82
.81

.80

·30 ·20 '10 0 10 20 30 40 50 60 70 80 90 100

Temperature (0C)

.89 P (bors)
0

~ ~ ~ 0 ~ 20 30 ~ ~ ro 70 80 90 00
Temperature eC)
Structural relaxation in the glass transition region 119

@
.89 P(borsJ

.88
(c)

- .87

,.,.!i!l
E
0
'-'
v
E
~
~~
~
"0
>
0
~
.~
Q.
f/l

.81

00

-30 -20 -10 0 10 20 30 40 50 60 10 80 90 100


Temperature (0C)

Figure 3.12 Specific volume as a function of temperature at different pressures, as


indicated, for different glass formation histories. The filled circles represent data
obtained during isobaric cooling at SoC h-1 from an initial equilibrium liquid state.
The open circles represent data obtained by temperature-pressure jumps. In (a), all
pressure changes are made in the liquid, whereas in (b) and (c), all pressure changes
are made in the glass, formed at either atmospheric pressure (b) or at 800 bar (c).
For further details, see reference 118. (Reproduced from reference 118, with
permission .)

independent variable (P), the cooling rate being held constant; on the other
hand, ~* is not uniquely defined by P, being dependent upon both P and
the manner in which the pressure was applied. The dependence of ~ and
~* on pressure is illustrated in Figure 3.12 by the full lines with negative
slope which join the intersection points in each case.
The change in ~ or ~* with pressure is not linear, and it is commonly
found that d ~/dP and d ~* /dP decrease with increasing pressure. The data
of McKinney and Goldstein give d Tg/dP decreasing from O.0266°C bar - 1 to
O.0149°C bar-I, while d ~*ldP also decreases over the same pressure range.
More notably, though, the values of d Tg/dP and d ~* /dP differ by a factor
of almost two. These results and other data from the literature are collected
in Table 3.3. There is considerable variability in these results, due in part to
Table 3.3 Pressure dependence of the glass transition temperature, and comparison with the Ehrenfest relationships (equations
3.24 and 3.25)

dfg/dP(O( bar- 1)

Volumetric Volumetric
Material isothermal isobaric Other 8.Kllla. (O( ba r - 1) TVlla.lllCp (O( bar-1) Ref.

Polystyrene 0.375 0.024 6


0.050 0.047 9
0.017 0.025 119
0.02 120
0.03 121
0.030' 0.061 122
0.030 0.05, 0.115" 0.034 123
0.048,0.019" 0.040 124
0.03 0.071-0.083 125
0.025 0.030 (at P = 0) - 126
0.032 0.073 0.034 127
0.079 L 0.0316 128
0.032 V 129
0.0742L 0.0316 0.078b 130,131
0.032D 132
0.031 0.030 133
0.020-0.033 134
0.027 RB 135
Poly(vinyl chloride} 0.014 0.046 15
0.0135 0.0335 121
0.Ol6v 129
0.016° 136c
0.017v 137
0.038l 0.013 0.0281 0.032-0.048 138
0.0215° 139
0.011-0.025 v•d - 140
0.018 0.0219 141 e
0.019-0.011 0.045 0.025 142
Poly(vinyl acetate} 0.0266-0.0149 - 0.0487-0.0317 118
0.044 122
0.021 0.022° 0.025 136
0.037 l 0.015 0.0221 0.026-0.051 138
0.0221 141 e
0.020 0.024c.c 143
0.Ol8v 144
0.0148-0.0264 - 145
Poly(methyl 0.018 15
methacrylate} 0.020 120
0.023 0.03 121
0.065 122
0.029v 129
0.022 0.032 133
0.024v 137
0.0545 e 0.0233 e 141
0.0211f 146
0.0236 147
0.02 148
Table 3.3 (Continued)

d TidP (QC bar- 1)

Volumetric Volumetric
Material isothermal isobaric Other 11K1l1rx (QC ba r -1) TVMJI1Cp (QC bar- 1) Ref.

PVDC9 0.016 v 140


PrxMS h 0.05-0.01 0.052 0.052 149
Polyester 0.022 0.026 119
PEr 0.016 V 144
PET i 0.025 150
PBTj 0.010 151
Polycarbonate 0.044D.V.c 136
Polycarbonate 0.0461" 141
Polysulphone 0.055 L 0.057 0.060 152
PPO k 0.05 T 153
PCHMA1 0.0224 147
PnBMAm 0.0204 147
HDPE n 0.005 v 154
Polypropylene 0.020" 0.0238" 141
0.009 v 154
0.018 v 155
PEEKo 0.057 0.059 0.054 0.024 156
PVc/PVAcP 0.011-0.014v 144
PS/PFSq 0.02-0.03T 153
BAMO/THFr O.OlO D,T 157
Polyisobutene 0.067 0.025 119
0.0241" 141
0.019 v 158
0.014N 159
0.035 160
Polybutadiene 0.013 160
Ethylene/propylene 0.010 160
cis-Polyisoprene 0.020 161
0.0212 162
0.016 163
0.018 0 164
Natural rubber 0.119 0.018 6
0.038 0.018 119
0.026- 0.0248- 141
0.029 160
0.024c 0.027 162
Polyurethane 0.018~0.01 OSD 0.037 ~0.017 16S
Glycerol 0.004D•e 0.004 e 136
n-Propanol 0.007 D•e o.oose 136
Phenolphthalein 0.019 133
Sucrose 0.008 0.007 133
Glucose 0.023 0.007 6
Salicin o.oose o.oose 136
Colophony 0.013 e 0.011' 136
820 3 0.020e O.027 e 136
(KN03)o.~(Ca(N03)2)o5~ ~ 0.003 0.003 133
42.8AgI 0.0040 166
so A/ 0.0046 166
0.0036 167
60 Aqlu 0.0037 166
Selenium O.013 C 0.011 136

"Different values of 1!!.K/t1rx result from different !J,.K values; !>from figure 10 of ref. 127 at P = 0; <quoted in Table I of ref. 136; see original cited
reference for details; dvalues quoted for several grades of PVC, including plasticized PVC; equoted in Table 8 of ref. 141; see original cited reference
for details; fisotactic PMMA; gpoly(vinylidene chloride); hpoly(rx-methyl styrene); ipoly(ethylene terephthalate); ipoly(butylene terephthalate);
kpoly(phenylene oxide); 'poly(cyclohexyl methacrylate); mpoly(n-butyl methacrylate); nhigh density polyethylene; °poly(ether ether ketone); Ppoly(vinyl
chloride)/poly(vinyl acetate); qpolystyrenelpolyfluorostyrene; 'bis(azodimethyl) oxetaneltetrahydrofuran; >Silver iodomolybdate, 42.8 mol% AgI; 'silver
iodomolybdate, 50 mol% Agl; "silver iodomolybdate, 60 mol% AgI; 'isochoric, dilatometric; Ldilatometric, low pressure glass; see reference 128 for
details; vviscoelastic; Ddielectric; RBRayleigh~Briliouin scattering; Cdynamic compressibility; Thigh pressure differential thermal analysis; NNMR.
124 Relaxation processes and physical aging

the experimental difficulties, but also reflecting the fact that different glassy
structures are obtained dependent upon the glass transformation history.
Nevertheless, it is possible to verify that the inequality 3.27 is satisfied in the
large majority of cases, with exceptions probably being ascribed to errors in
the measurement of the change in compressibility ~K at the glass transition.
Furthermore, in those cases where a direct experimental comparison has
been made [e.g. polystyrene [128,130,131], poly(vinyl chloride) [138] and
poly(vinyl acetate) [138]], the isothermal variation of I'g with pressure is
much greater, by a factor of two or more, than the isobaric variation of I'g,
in agreement with McKinney and Goldstein [118]. The reason for this is
clearly seen in a comparison of Figures 3.l2a and 3.12b; because the
compressibility of the melt is much greater than that of the glass, the
intersection points of glassy and melt behaviour which define I'g* lie on a
line of much shallower slope in Figure 3.12b than they do in Figure 3.12a.
This difference can be, and indeed has been, the cause of some confusion,
but need not be so if a rigorous definition of the glass transition temperature
is adopted, namely that I'g is the temperature at which the equilibrium melt
transforms to a glass, at whatever pressure, as the temperature is reduced.
In this sense, the isobaric dI'g/dP above is the logical definition of the
pressure dependence of the glass transition temperature. The application
of pressure to the non-equilibrium glass yields a different structural state
which, strictly, cannot be defined by a glass transition temperature as it
was not formed, either by pressure or temperature, from an initial equilib-
rium state. Accordingly, such a state would be better defined in terms of a
fictive temperature and a fictive pressure, being respectively the temperature
and pressure at which an equilibrium state would apparently be achieved
by, respectively, an instantaneous isobaric or isothermal change in these
variables.
It follows immediately from these considerations that glasses of different
structure (i.e. of different fictive temperature and pressure) may be obtained
at the same experimental temperature and pressure by formation following
different temperature-pressure histories, and this is the procedure for the
preparation of'densified' glasses. Consider the schematic volume-tempera-
ture plot in Figure 3.13, which shows two routes to the formation of a
glass at temperature Tl and pressure PI' The first route starts from
equilibrium at T2 and PI (point A, usually atmospheric pressure), and cools
at constant rate and at constant pressure PI through the glass transition
(point B) until the glassy state at temperature Tl is achieved (point C). The
second route is to start from the initial equilibrium state (A), but to apply
pressure P2 to the melt isothermally at T2 (point D) before cooling
isobarically at pressure P2 and at constant rate. This route passes through
the glass transition temperature at pressure P2 (point E) before reaching the
Structural relaxation in the glass transition region 125

>
ai D Pressure
E P2
g
:::l

C
G

T, T2
Temperature, T

Figure 3.13 Schematic illustration of pressure-temperature cycle in the formation


of densified glasses. The line BE defines d T/dP, the change in isobaric glass transition
temperature with pressure.

glassy state at temperature TI (point F), whereupon the pressure Pz is


released isothermally to the initial value Pl' The volume responds by
following the path to point G, which has a lower volume than the glass in
state C at the same temperature and pressure.
Experimental results have shown significant increases in density for
densification pressures (Pz in Figure 3.13) usually in the range 3-5 kbar. For
polystyrene, for example, average increases of between 0.3% kbar- l and
0.6%kbar- 1 have been observed [120,133,168-171], and for poly(methyl
methacrylate) of about 0.5% kbar- l [133,172]. Similarly, Bree et al. [173]
found increases of between 0.44% kbar- l and 0.59% kbar- l for polystyrene,
poly(methyl methacrylate), poly(vinyl chloride) and polycarbonate. Interest-
ingly, much smaller density changes, less than 0.1 % kbar- \ were found for
the inorganic iodomolybdate glasses [166,167], which also have much
lower values of d 'Fg/dP than the organic polymer glasses (Table 3.3). This
could result, though, from a levelling out of the effects of densification
as the pressure is increased, as has been observed in polymer glasses
[133,168,171,172], since the density change in these inorganic glasses was
measured only after densification at 50 kbar pressure.
Whilst densification of polymer glasses appears to result in a monotoni-
cally increasing density with pressure, at least up to a certain pressure
126 Relaxation processes and physical aging

5.----------------------------.

-I~------~L---------L-------~
a 100 200 300
Po IMN/m2 1

Figure 3.14 Dependence on pressure of the enthalpy, relative to a sample cooled


at '" 1°C min-' at atmospheric pressure, of densified poly(methyl methacrylate)
glasses formed by cooling from the melt at pressures up to 3 kbar. (Reproduced from
reference 172, with permission.)

(usually of the order of 3 kbar) at which the effect levels off, this is not
mirrored in the same dependence of the enthalpy. In early work, Allen et al.
[174J and shortly thereafter Ichihara et ai. [175J found the surprising result
that densification of polystyrene, at pressures up to 1.2 and 0.6 kbar
respectively, produced significant increases in density but very little, if any,
changes in enthalpy. Despite the results of Kimmel and Uhlmann on
poly(methyl methacrylate) [176, 177J which do not follow the usual trend,
possibly because of the different ways in which the densifying pressure may
be applied, there is now strong evidence that the enthalpy change is very
small and possibly slightly negative, for the first kilobar or so of densifying
pressure, but thereafter the enthalpy increases with increasing pressure in
contrast to the reduction in volume [133,168, 170, 178, 179]. A typical illus-
tration is shown in Figure 3.14. Clearly the early interpretation, by Shishkin
[120] and Mackenzie [180J for example, that densified glasses are closer to
equilibrium, is not correct.
The enthalpy changes are commonly found by DSC on the pressurized
and normally prepared glasses, and typical DSC traces are shown in Figure
3.15 for densified polystyrene. Densification introduces an endothermic peak
at a temperature below 1'g, and this peak shifts to lower temperatures with
increasing densification pressure. This behaviour appears to be quite general
for polymeric glasses, although rather different behaviour has been observed
in some inorganic glasses [166,167].
Structural relaxation in the glass transition region 127

2·0,-----------------,

Cp = 1·7----

320 360 400


T (K)
Figure 3.15 DS( traces at 20 0 ( min-' for polystyrene glasses formed by cooling
from the melt at l°(min-' at various pressures: A, 3.1 kbar; B, 1.4kbar; (,
reference. (Reproduced from reference 170, with permission.)

The increase in enthalpy produced by densification has been attributed by


Weitz and Wunderlich [133] and by O'Reilly and Mosher [181, 182] to the
freezing in of high energy conformations (trans-gauche), and the conforma-
tional changes have been followed in poly(vinyl chloride) by Fourier
transform infrared spectroscopy [179,181,182]. On the other hand, the
endothermic peak below ~, which has an associated enthalpy change of
the order of 1 J g-l calculated from the area of the peak, is believed to
result from the relaxation of frozen-in volume strain, involving bond
length changes and bond angle distortions; a volume strain of 0.4% and an
internal pressure of 3 kbar would account for the magnitude of the enthalpy
change.
A similar argument was proposed by Destruel et al. [183], and was based
upon the idea of the modification of the Lennard-Jones potential energy
function. The process of densification is considered to have two simulta-
neous effects: a reduction in the fractional free volume (equivalent to the
bond length and bond angle changes above), and the development of an
internal stress which increases the internal energy (equivalent to the confor-
mational changes above). Their analysis predicts that the former effect
dominates when the densification pressure is small, whilst the latter develops
quickly at higher pressures. The transition from one mechanism to the other
occurs at about 1 kbar, in pleasing agreement with the kind of behaviour
illustrated in Figure 3.14.
128 Relaxation processes and physical aging

Finally in this section, it is interesting to note that the dependence of glass


structure on the whole of its pressure and temperature history may lead to
the formation of new structural states. Some recent work on the effects of
pressure and of densification on some silver iodomolybdate glasses [166]
has shown that, under certain vitrification circumstances, irreversible
changes can occur, associated with disproportionation reactions, and lead-
ing to the appearance of biphasic structures which are not observed in
normally prepared glasses. This suggests that new kinds of glassy structure
may be produced by an appropriate combination of pressure and tempera-
ture history.

3.3 SECONDARY RELAXATIONS

3.3.1 INTRODUCTION
Structural relaxation in the glass transition (ex-relaxation) region, described
above, results from molecular motions which involve relatively large seg-
ments of the polymer backbone chain. In contrast, the various secondary
(P, y . ..) relaxations, which occur at lower temperatures, are generally
believed to involve much more localized molecular motions. This is intu-
itively expected as the glassy state below 1'g reduces considerably the
molecular mobility. In spite of this, there often appears to be a surprising
amount of mobility, and these molecular motions can have a significant
effect on the macroscopic properties of polymers, manifest for example as
stepwise changes in the mechanical and physical properties such as the
modulus, the dielectric constant and the thermal expansion coefficient. An
illustration of the possible types of molecular motion which may give rise to
secondary relaxations is shown in Figure 3.16, following the original
classification of Heijboer [184], and these molecular motions are outlined
below.
• Type A. This involves a rather localized motion of a small segment of the
main chain, in which rotation about certain backbone bonds occurs. Such
a possibility was proposed many years ago as a general mechanism, and
is known as the so-called crankshaft mechanism, with variants proposed
by Schatzki and Boyer (see reference 23 for further details). This type of
motion has been attributed to the p-relaxation in rigid poly(vinyl chlor-
ide), to the y-relaxation in polycarbonate, and to secondary loss peaks in
polysulphones and polyesters [184].
• Type B. In this type of molecular motion, the whole of the side group
rotates about the bond linking it to the main chain. In so doing, the group
typically moves from one potential energy minimum to another, without
Secondary relaxations 129

Figure 3.16 Schematic illustration of the possible modes of molecular motion in


secondary relaxation in glassy polymers. (After Heijboer [184], with permission.)

necessarily making a complete rotation, and may require a certain amount


of cooperative movement of the main chain. The best examples of this
type of molecular motion are afforded by the poly(n-alkyl methacrylate)s .
• Type C. Here, it is only the very localized internal motion of a unit within
the side group which is contributing to the secondary loss. Again, the
polymethacrylates provide a good example of this type of molecular
motion, in this case giving rise to the y-relaxation peak.
Perhaps one of the most important of these secondary relaxation effects
is the well known toughness or impact strength of polycarbonate, usually
considered to be associated with the low temperature y-relaxation process
in this polymer. The molecular origins of these secondary relaxation
processes in a wide variety of glassy polymers have been the subject of much
debate over a number of years, and were admirably reviewed in 1967 by
McCrum, Read and Williams [23]. In the 30 years or so that have passed
since then, considerable further advances have been made, some of which
are outlined in the following section, beginning with the example of
polycarbonate cited above.

3.3.2 POLYCARBONATE AND OTHER PHENYL RING-CONTAINING


MAIN CHAIN POLYMERS
Polycarbonate (4,4'-dioxydiphenyl 2,2'-propane) was first studied by Illers
and Breuer [18SJ using a torsion pendulum at 1 Hz, and they identified two
principal relaxation regions: the main glass transition at 155°C, and a
prominent y-relaxation peak at - 97°C, though there is the appearance at
about - 30°C of a slight shoulder on the y-relaxation. In the more
commonly used bisphenol A polycarbonate, the J1-relaxation at about 80°C
130 Relaxation processes and physical aging

appears as a shoulder on the low temperature part of the a-relaxation [186].


The p-relaxation is found to depend strongly on thermal history, being
significantly reduced by annealing, and has been attributed to either internal
orientational stress [185] or to packing defects in the glassy state [187].
It is the molecular origin of the y-relaxation which is the more interesting,
however. For many years this relaxation has been associated with phenyl
group 'flips'. Conformational calculations show that for isolated chains of
both polycarbonate and poly(phenylene oxide) at room temperature, phenyl
rings are free rotors. Experimental NMR studies have shown that polycar-
bonate in the glassy state exhibits 180° flips by hindered rotation, in contrast
to poly(phenylene oxide) for which only small amplitude motions are seen
[188]; similarly, by dynamic mechanical analysis, a damping peak is evident
in polycarbonate at about -100°C (1 Hz) whereas poly(phenylene oxide) is
rather featureless in this temperature range [189]. The activation energy for
the y-relaxation in polycarbonate is 54 kJ mol- 1 [187] which, according to
Schaefer et al. [188], would imply a frequency of about 200 kHz at room
temperature, the same frequency as the phenyl ring flips, though this would
in fact require a slightly lower activation energy, about 43 kJ mol- 1 . More
recently, however, it has been suggested [190,191] that phenylene 180° flips
are not as important as had previously been considered in the y-relaxation
of polycarbonate, or of phenyl ring containing main chain polymers in
general, and that cooperative intermolecular as well as intramolecular
motions are involved.
In a comparison of a number of polyarylates (polyester, polyethersul-
phone, polyaryletherketone and polyetherimide) with 0, C 0, CO-O-
and S02 between the phenyl units of the main chain, Schartel and Wendorff
[191] have measured the activation energies for the y-relaxation, as well as
more detailed features such as the width of the relaxation, its asymmetry and
variations with the different chemical structures considered. All the activa-
tion energies are very similar, 46 ± 3.5 kJ mol- 1, as are the pre-exponential
factors, and compare closely with that for polycarbonate, which might
suggest that phenyl ring flip is common to all systems. In addition, though,
the relaxation was found to be rather broad (as measured by the KWW
parameter p), which would not be anticipated if the relaxation were wholly
intramolecular, as the relevant molecular unit would then always relax in
the same surroundings, with a consequently narrow relaxation. Further-
more, the actual chemical structure of the polyacrylate has little influence
on the relaxation kinetics, which suggests again that the intramolecular
reactions are not very important.
The conclusion is that intermolecular interactions are at least as impor-
tant, and the asymmetry of the relaxation indicates an extent of intra-
molecular correlation involving only about one repeat unit in length.
Secondary relaxations 131

trans-trans trans- trans trans-cis trans-trans

trans -trans trans- cis trans-trans trans- trans


Figure 3.17 Schematic illustration of the bisphenol A polycarbonate chain. The (-0
bonds with asterisks indicate the points of bond rotation. The phenyl rings which flip
when the (-0 bonds rotate are numbered. (After Jones [192], with permission.)

A molecular model which combines the intramolecular motions with inter-


molecular coupling to the bulk glassy polymer has been proposed by Jones
[192], and a schematic illustration is shown in Figure 3.17.
The primary trans-trans polycarbonate chain undergoes correlated con-
formational interchanges between two neighbouring carbonate groups, with
the exchange of a trans-cis with a trans-trans conformation. This requires
rotation about the c-o bonds indicated, with the simultaneous flip of the
phenyl group about the C 1 C 4 axis. This motion produces a volume
fluctuation as a result of the translation of the bisphenol A group, and a
change in shape of the carbonate unit; together, these can diffuse down the
chain length with successive conformational interchanges to provide a
mechanism for the rapid dissipation of strain, and hence for energy absorp-
tion in impact. Furthermore, the model is consistent with the geometric
requirements of NMR data and with the existence of both dielectric and
dynamic mechanicalloss.

3.3.3 POLY(METHYL METHACRYLATE) AND RELATED POLYMERS

In one of the earliest studies of dynamic mechanical loss behaviour of


methacrylate polymers, Hoff et al. [193] identified both f3- and y-relaxation
regions in addition to the main et-relaxation glass transition. These authors
noted that the f3-relaxation could be attributed to rearrangements of the
132 Relaxation processes and physical aging

-COO- portion of the side group by rotation around the link to the main
chain, and that the magnitude of the relaxation decreased, and eventually
disappeared, as the length of the side chain was increased in the series
methyl, ethyl, n-propyl and n-butyl. These observations have subsequently
been confirmed, most notably by Heijboer [e.g. 194], who finds the p-peak
at about lOoC at 1 Hz, independent of the length of side chain. The contrary
observations of other authors, whereby the p-peak is affected by length of
side chain, is attributed to the problem of the merging of a and p-relaxations
for the higher n-alkyl methacrylates, which distorts the location and appear-
ance of the p-peak. This occurs in these polymers because the a-relaxation
(I'g) decreases with increasing length of side chain, explained by the
simultaneous increase in free volume which enables the steric hindrance
caused by neighbouring chains to be overcome more easily.
Although it is widely accepted that the p-relaxation in methacrylate
polymers is associated with an intramolecular rotation of the COOR group,
where R depends upon the particular homologue, there is evidence of a
certain amount of intermolecular interaction. For example, the dielectric
p-relaxation in poly(n-alkyl methacrylate)s is much broader than the a-
relaxation, and this has been attributed to the different glassy state environ-
ments in which the side group motion occurs [195, 196]. Similarly, dielectric
relaxation studies of poly(dimethylphenyl methacrylate)s suggest not only a
coupling of the local side group motions to motions of a small segment of
the main chain but also that these motions are influenced by intermolecular
interactions [197].
A lower temperature, y-relaxation process, at about -150 a C at 1 Hz, was
identified by Hoff et al. [193] in n-propyl, n-butyl and higher homologues
of the methacrylate and chloroacrylate series. The shorter side chain
polymers of these series do not show this y-peak [194], which is supposed
to arise from the independent flexibility of the longer alkyl components. This
can be illustrated with respect to the ethyl and propyl esters by reference to
Figure 3.18. For the ethyl ester, the conformational changes involve very
little 'volume of action', whereas there are a number of possibilities for the
propyl ester, involving rotation about bonds A, Band C. The activation
energies are of the order of 21 ± 4 kJ mol-I, similar to the energy barriers
for rotation about C-C bonds in simple paraffins. Some recent molecular
mechanics calculations [198] for the y-relaxation in poly(propyl metha-
crylate) give an activation energy of 16,4 kJ mol- 1 for hindered rotation
around the OCH 2-CH 2 bond of the propyl group, in reasonable agree-
ment with the measured value, suggesting that the attribution of this
relaxation mechanism is indeed correct.
The y-relaxation is also evident in poly(cycIohexyl methacrylate) at about
- 80°C at 1 Hz [184]. In this case the molecular motion is believed to be
the flipping of the saturated six-member ring, and involves an activation
Secondary relaxations 133

1l.A lLB

Figure 3.18 Schematic illustration of conformational changes possible in ethyl (I)


and propyl (lIA and liB) esters. The main chain carbon atom is denoted em. Different
conformations can be achieved by rotation about (-0 and (-( bonds, as indicated
by the arrows. (Reproduced from reference 193, with permission.)

energy from dynamic mechanical analysis of 47.3 kJ mol-I, in good agree-


ment with the NMR value of 47.7 kJ mol- 1 for the chair-chair transition of
the cyc10hexyl ring. Other cyc10alkyl side group methacrylates [199] show
y-peaks at temperatures decreasing in the order six-ring ( - 79°q, eight-ring
( -111 0q, seven-ring (,.... - 1700q and five-ring ( < - 190"q, with activa-
tion energies decreasing in the same order. The implication is that the
six-member ring is the least mobile and the five-member and, to a slightly
lesser extent, the seven-member ring are the most mobile. Also, the width of
the relaxation increases in the same order, implying that the complexity
of the relaxation is least for cyc10hexyl methacrylate.
A further loss peak at even lower temperatures, the (i-relaxation, is evident
from Heijboer's data [184] on poly(cyc1ohexyl methacrylate), though only
the high temperature tail is seen. Unlike the y-relaxation, this cannot be
attributed to a chair-chair transformation, as it is also seen in 1,I-dichloro-
cyc10hexane for which both conformers are identical. Heijboer [199] sug-
gests that it may be due to rotational oscillation of the whole cyc10hexyl
nng.

3.3.4 POLYSTYRENE
The early work of Illers and Jenckel [200,201] identified, in addition to the
main ex-relaxation glass transition in polystyrene, at least three further
secondary relaxations. These have subsequently been confirmed by many
134 Relaxation processes and physical aging

other workers, for example Yano and Wada [202,203] who find the
f3-relaxation at 77°C, the y-relaxation at -143°C and the b-relaxation at
- 218°C, all at a frequency of 10 Hz. There has been considerable debate
about the molecular origins of these relaxations, though phenyl group
motions have nearly always been involved in one way or another.
In one of the earliest studies, Sinnott [204] attributed the b-relaxation to
motion of the phenyl groups, later suggested by Yano and Wada [202,203]
to be located in regions where syndiotactic diads inserted between isotactic
sequences, or vice versa, introduced lattice defects. It has also been suggested
[205,206] that this relaxation is due not to complete rotation of the phenyl
group but to a wagging motion between two energy minima. Confirmation
of this idea came from the conformational energy calculations of Reich and
Eisenberg [207], who compared the activation energy found experimentally,
c. 7.5 kJ mol- 1, with a computed value for complete phenyl ring rotation
of c. 29.3 kJ mol- 1, and concluded that complete phenyl ring rotation
cannot be the origin of the b-relaxation.
The y-relaxation was originally assigned by Illers and Jenckel [201] to
torsional motion of methylene sequences resulting from occasional head-to-
head, tail-to-tail polymerization, thus introducing weak points in the main
chain which is normally head-to-tail polymerized. The implication is that
the strength of the y-relaxation should depend on polymerization condi-
tions, and support for this was provided by data [208] showing that the
mechanical loss values were consistently higher for atactic than for isotactic
polystyrene, as would be expected. This was disputed by Yano and Wada
[202] on the basis of dynamic mechanical analysis of anionically poly-
merized polystyrene, for which none of the heterojunctions postulated by
Illers and Jenckel are anticipated. Furthermore, the absence of the y-
relaxation in dielectric spectroscopy suggested that phenyl group rotation
was involved, supported later by the results of Reich and Eisenberg [207]
and by Shimizu et al. [209]. The activation energy of the y-relaxation was
37.7 kJ mol- \ in reasonably close agreement with the calculated energy for
phenyl group rotation of 33.5 kJ mol- 1 [207]. Similar molecular conforma-
tion calculations by Tonelli and coworkers [210,211], however, led them to
the conclusion that the large steric barrier to complete rotation of the
phenyl group meant that it was restricted to ± 20° about positions where
the plane of the phenyl ring bisects the backbone valence angle at the
asymmetric carbon atom to which it is attached. In fact, they showed that
complete rotation can only occur in the unlikely situation when the back-
bone adopts sterically unfavourable conformations, for which an activation
energy of 37.7~41.8 kJ mol- 1 again applies. They conclude for the y-relax-
ation, as indeed also for the b-relaxation, that phenyl group motions other
than complete rotation are involved. More recent results [212,213] support
this view, and emphasize the role of phenyl~backbone interactions.
Secondary relaxations 135

The fJ-relaxation is known to be resolved only at low frequencies, merging


into the IX-relaxation at frequencies higher than about 40 Hz [201]. The
nature of the molecular motions involved in this secondary relaxation still
remain unclear. Illers and lenckel [200] associated it with motion of phenyl
groups in parts of the main chain where there was less steric hindrance, and
the NMR data of Vol'kenshtein et al. [214] was considered to favour this
interpretation. Yano and Wada [202], on the other hand, consider that
rotational vibrations of main chain segments, so-called local mode relax-
ations, more accurately account for the narrowing of NMR absorption
spectra. The question of phenyl ring rotation and its effect on NMR data
has been considered more recently in some detail [215,216].
The role of phenyl groups in glassy polystyrene has also been studied by
X-ray scattering [217]. Strong interphenyl correlations, both intermolecular
and intramolecular, are suggested, and a model in which phenyl groups from
neighbouring molecules associate in stacks in microsegregated regions has
been proposed to account for the observed scattering peak at 0.75 A-1.
Computer simulations of polystyrene glassy structure [218], showing the
packing of phenyl rings at distances of the order of 5 A, with a tendency to
align at right angles to each other, are able to predict to a reasonable extent
the radial distribution functions for X-ray scattering, but truncation effects
do not permit the representation of the 0.75 A-1 peak.

3.3.5 EFFECT OF AGING ON SECONDARY RELAXATIONS


The main IX-relaxation glass transition has been shown in section 3.2 to be
influenced substantially by aging, with the changes monitored macroscopi-
cally by dilatometry or calorimetry, and microscopically by a variety of
techniques. One (and by far the most widely used) interpretation is in terms
of a reduction of free volume, and hence a reduction in molecular mobility
on aging. One might anticipate that this reduction in mobility could have
an effect on the secondary relaxations, and in particular on the fJ-relax-
ations, which typically involve either local mode relaxation of segments of
the main chain or side group rotation about bonds to the main chain. Such
considerations would be especially relevant for the latter case, where the
molecular motions are sometimes considered to be affected by interactions
with the main chain, for example in some methacrylate polymers [197].
A common problem in studying the effect of aging on the fJ-relaxation is
the close proximity, and often even merging, of the IX- and fJ-relaxations in
some amorphous polymers. For example, in dynamic mechanical analysis of
polystyrene the fJ-relaxation often appears as just a shoulder on the low
temperature side of the IX-relaxation peak, and the same is true for
poly(methyl methacrylate) and for bisphenol A polycarbonate; in contrast,
the two relaxations are much more widely separated, at dynamic mechanical
136 Relaxation processes and physical aging

frequencies, for poly(vinyl chloride) and poly(vinyl acetate). For those


polymers for which the ()(- and f3-relaxations are close, the influence of the
one on the other must be carefully considered.
The early torsion pendulum studies of Golden, Hammant and Hazell
[219] on polycarbonate showed a f3-peak at about 80°C which reduced in
magnitude and became sharper after annealing at 132°C for 24 h. Similar
dynamic mechanical results of Allen et al. [220] associated the loss of the
f3-peak with embrittlement, while a dielectric study [221] also showed a
decrease in the f3-peak magnitude on annealing, but with no change in its
location on the temperature or frequency axis.
This last observation now appears to be a rather common, but not
universal, result. Thus, in extensive dielectric relaxation studies of a wide
range of organic and other glassy systems, Johari [222-224] has found that
the effect of isothermal annealing, or of reducing the cooling rate, is a
reduction in the magnitude of the f3-relaxation without changing its location
on the temperature or frequency scale. The interpretation is that the
f3-relaxation in general in glassy systems involves a non-cooperative molecu-
lar motion which is hindered by its environment, an 'island of mobility',
which itself is mobile only through cooperative molecular motions. Thus the
fundamental molecular motions of the f3-relaxation are coupled to the more
global structural state, which is viewed as an inhomogeneous structure with
regions of high molecular density separated by loosely packed regions.
Johari argues that the f3-relaxation strength, which reflects the number of
loosely packed sites in the glassy structure, would decrease on annealing.
Diaz-Calleja et al. [225J likewise observed, by dynamic mechanical analysis
of poly(methyl methacrylate), a reduction in the height but not in the
relative position of the f3-peak, and discussed this in terms of a reduction in
the population of 'defective zones' with no change in barrier height. The idea
of defects being introduced by quenching of polycarbonate was also con-
sidered by Yee and Smith [187J as the origin of their f3-relaxation, which
was reduced in amplitude on annealing. There are numerous other examples
of reports of a reduction in the magnitude of the f3-relaxation with no
change in time scale: for example, the dielectric and thermally stimulated
depolarization studies of Guerdoux and Marchal [226], the creep response
of various amorphous and semicrystalline polymers investigated by Read
and coworkers [227,228], and the dynamic mechanical analysis of
poly(alkyl methacrylate)s [229].
As mentioned earlier, however, there are exceptions to these observed
effects of aging on the f3-relaxation, and there are also different interpre-
tations. Interesting exceptions are seen in the dilatometric data of
Goldbach and Rehage [9, 16J and of Greiner and Schwarzl [19], these two
groups finding opposite behaviours in different amorphous polymers. The
Secondary relaxations 137

p-transition is manifest as a change in the slope of the volume-temperature


curve, or as a step change in the thermal expansion coefficient. For
polystyrene, Goldbach and Rehage [9,16] found that, curiously, the p-
transition moved to lower temperatures as the cooling rate increased, in
other words in the sense contrary to the effect of cooling rate on the glass
transition. This observation on the basis of dilatometric data was confirmed
by dynamic mechanical analysis and was rationalized as follows. The
motion of phenyl groups at 1p requires a certain volume, so that rapidly
cooled specimens, which have a higher specific volume than slowly cooled
specimens, will freeze-in the phenyl group motion at lower temperatures.
Greiner and Schwarzl [19], on the other hand, found for polystyrene,
poly(vinyl chloride), poly(methyl methacrylate) and polycarbonate that the
temperature of the secondary p-relaxation shifted to lower temperatures
with decreasing cooling rate. The results for polystyrene do not show a very
well defined Tp, and therefore the dependence of 1p on cooling rate was not
shown in their analysis. The results for poly(vinyl chloride), poly(methyl
methacrylate) and polycarbonate were more precise in this respect, however,
and were compared with the mechanical loss data of Heijboer [230] which
showed the same cooling rate dependence. No explanation was offered by
Greiner and Schwarzl for the different dependence on cooling rate observed
by Goldbach and Rehage [9,16].
Even when there is agreement on the apparent effect of aging on the
p-relaxation, the interpretation can be quite different. Thus, in their investi-
gation of aging and the p-relaxation in poly(methyl methacrylate), Muzeau
and Perez and coworkers [231-233] observe an apparent reduction in the
strength of the p-peak, but consider this only to be the influence of aging
on the low temperature tail of the IX-relaxation, which has the effect of
distorting the appearance of the J1-relaxation. This interpretation accords
with the long-held view of Struik [18], summarized in a study of volume
relaxation and secondary transitions in a wide range of amorphous poly-
mers [234], that thermal history does not affect the p-relaxation but does
affect the onset of the IX-relaxation.
The difficulty in interpreting the effects of aging on the p-relaxation, and
indeed on the IX-relaxation also, often lies in the thermal history used in the
measurements themselves. Thus, for example, dilatometric studies are
usually made at various cooling rates, or isothermally following aT-jump.
Dynamic mechanical analysis, on the other hand, is usually made on heating
at constant rate after previously subjecting the sample to a certain aging
treatment; but the process of heating is in itself a thermal history which must
be taken into consideration. In this respect there have been several recent
reports of an additional relaxation peak, denoted p', in poly(methyl metha-
crylate) when a quenched and annealed sample is studied by dynamic
138 Relaxation processes and physical aging

mechanical analysis [225,231-233,235,236]. These are very reminiscent of


the so-called 'sub-~' peaks seen in differential scanning calorimetry of
glasses, both organic and polymeric, which have been quenched and then
annealed at a relatively low temperature with respect to ~ [10,237,238].
These sub- ~ peaks result from the interaction of the quench-and-anneal
thermal history with the distribution of relaxation times when the constant
heating rate of the DSC scan is applied. Similarly, one might anticipate
additional peaks in tan £5 when the usual constant heating rate of dynamic
mechanical analysis is applied, and such peaks should therefore be inter-
preted with some care. Because the heating scan is continuously changing
both the fictive temperature and the thermal history of the glass, this
interpretation may be difficult. Ideally, a better experimental procedure for
investigating the aging dependence of damping peaks, either mechanical or
dielectric, would be to follow the isothermal changes in, for example, tan £5
or the complex modulus, after a quench from above Tg • The compilation of
such data over a range of temperatures and frequencies would then provide
the damping curves for equal values of aging time; unfortunately, such
experiments have rarely been reported.

3.4 PHYSICAL AGING AND MECHANICAL PROPERTIES

3.4.1 INTRODUCTION
In earlier sections it has been shown how physical aging (or structural
relaxation) in the glass transition region can be described macroscopically
in terms of the volume or enthalpy relaxation behaviour, or microscopically
by techniques such as small angle X-ray scattering, positron annihilation
and other spectroscopic methods. The usual isothermal relaxation behav-
iour, in which the non-equilibrium glass, characterized by excess thermo-
dynamic quantities, gradually approaches an equilibrium state, is broadly
interpreted in terms of a lengthening time scale for molecular motion as the
molecular mobility is reduced. It is to be expected, therefore, that any
property of the glass which depends on molecular mobility would be subject
to physical aging, which would be manifest generally as a lengthening of the
time scale. In particular, the viscoelastic response of the glass in creep or
stress relaxation would be expected to shift to longer times, while the
dynamic mechanical response would shift to lower frequencies. In broad
terms, this is just what is observed in practice, as is amply demonstrated by
the extensive work of Struik [18]. An example of the effect of physical aging
on the creep response of poly(vinyl chloride) is shown in Figure 3.19, where
the major effect is a horizontal shift of the creep curves to longer times.
Since the earliest observations of McLoughlin and Tobolsky [239], who
found that stress relaxation in poly(methyl methacrylate) occurred much
Physical aging and mechanical properties 139

aging lime Ie ,days


lensile creep 1 1

compliance, TO- °m</ N 0..03·0.1-0.3-1-3 -10.-30. -10.0 --JOD-IDDD


1
5

1 / /- I ~~ / I / 11 / i
I // j
/1}; J
t /" / #
/1 ~/
0 -

o ./ I> l}/
/'7
I /0

0/
./'
0/ /
• ,,/ Il/ '7
.~~i /
/ t.

/
/

~
..!
v~ /
0
0
L..-
/
,.,1
L_
.-,;
I.
~ "" /); ,./J' /v ~/ 0 / . / %
0 •
. .v(
o "
----"/I_Il/ / ' 7 / /~
'7 /
/ov',...._ / 0 /p".,.J! I 2%
.--._0....-0 ...--__ </10.'"
,,-"I .-1....-1,- '7 ____ '7 -~ /~____ 0 ...... 0,"-;::f-
i--'~ i f "
0 """ ;-
1- -v_v 0
¥:::~-Y::ci::: ~~;::~.~
-c~~
,~ --.. -
---+ cTep lime I, sec

10'

Figure 3.19 Low-strain tensile creep curves for rigid poly(vinyl chloride) quenched
from 90 0 ( (about 100 ( above Tg) to 40°(, and aged at 40 0 ( for a period of 3 years.
The aging times are indicated against each curve. The arrow (almost horizontal)
indicates the direction of shifting to obtain optimum superposition. (Reproduced
from reference 18, with permission.)

less rapidly for a slowly cooled glassy sample than for a quenched sample,
the aging of mechanical properties has most commonly been attributed to
volume relaxation. This simple explanation, however, has not always been
reconcilable with experimental data; nor has the simple horizontal shifting
of the viscoelastic response, exemplified by the data in Figure 3.19, always
resulted in good superposition. Some of the implications of these and other
detailed aspects of the physical aging of mechanical properties are now
considered.

3.4.2 AGING OF THE VISCOELASTIC RESPONSE


The creep response of a viscoelastic material to a constant stress applied at
time t = 0 may be described by the time-dependent creep compliance [23]:

f +OO
J(t) = Ju + IlJ _ 00 <p(ln r)[1 - exp( - t/r)]d In r (3.28)

where Ju is the unrelaxed compliance, IlJ is the relaxation strength, equal


140 Relaxation processes and physical aging

to the difference between the relaxed compliance J R and the unrelaxed


compliance J v , and ¢(In 1") is the normalized distribution of creep relaxation
times. If the effect of aging were simply reflected in a change in time scale,
then for a thermorheologically simple system the relaxation time distribu-
tion for the aged sample, ¢e(ln 1"), would be related to that of the unaged
sample, ¢eo(ln 1"), by the equation [240J
(3.29)
where a e represents the aging time shift factor, in analogy with the shift
factor aT used in time-temperature superposition [23]. Adopting the
convention that superscript e denotes an aged sample (elapsed time te) and
superscript eo denotes an unaged sample (elapsed time t~), following Chai
and McCrum [240J, the aged and unaged compliances are then related
simply by
Je(t) = J"o(t/a e) (3.30)
The aged creep curve is therefore simply shifted along the log creep time axis
by an amount log a e relative to the unaged curve (refer to Figure 3.20,
dashed line).
As has been discussed in an earlier section, however, aging may have an
effect on the secondary p-relaxation in glassy polymers, which in respect of
creep would have a corresponding effect on the unrelaxed compliance in the

------- _/
Log time

Figure 3.20 Schematic illustration of the possible effects of physical aging on the
creep compliance. Curve A (full line), unaged sample; curve 8 (dashed line), aged
sample, aging affects time scale only; curve C (dash-dotted line), aged sample, aging
affects time scale, unrelaxed compliance and relaxation strength.
Physical aging and mechanical properties 141

IX-relaxation region. This has been particularly well demonstrated by the


creep and aging studies of Read [241] on poly(methyl methacrylate),
poly(vinyl chloride), polycarbonate and poly(butylene terephthalate). In
addition, there is the possibility that the relaxation strength I1J may be
affected by aging. The effect of aging would then no longer be a simple
horizontal shift along the logarithmic time scale, but would now involve
a horizontal shift together with a change in magnitude and location on the
vertical axis (refer to Figure 3.20, dash-dotted line). Under these circum-
stances, the procedure of superposition of aged and unaged creep curves
becomes questionable, though frequently this is done using both horizontal
and vertical shifts. However, a thorough analysis of the situation has been
made by Chai and McCrum [240] making use of additional shift parameters
to define the effects of aging time on the relaxation strength I1J, and on the
unrelaxed compliance J u :
(3.31)
J~ = cJti' (3.32)
These authors identify four different methods of superposition.
• Method 1. Horizontal shifting alone of J(t) data, assuming be = Ce = 1.
• Method 2. Horizontal and vertical shifting of J(t) data, assuming be = 1,
ce "# 1.
• Method 3. Horizontal shifting alone of log J(t) data, assuming be =
Ce = 1.

• Method 4. Horizontal and vertical shifting of log J(t) data, assuming be =


ce "# 1.
Only one of these (method 4) allows for the possibility of be #- 1, but it is
restricted by the requirement that be = Ceo Likewise, to investigate the
possibility that Ce #- 1 requires the further assumption either that be = 1
(method 2) or that be = Ce (method 4), again limiting the use of these
superposition procedures. Chai and McCrum showed, however, that it is
possible to eliminate one of these parameters by attempting to superpose
creep rate data rather than creep compliance data. Thus, they showed that
the time derivatives of J(t) in the aged and unaged conditions are related by
log je(t) = log jeo(t/a e) - log a e + log be (3.33)
Hence the log je(t) and log jeo(t) curves should superpose by means of a
horizontal shift of log a e and a vertical shift of (log a e - log be)·
Strangely, besides the work of McCrum and coworkers, this rational
approach to the problem of allowing for possible effects of aging on the
relaxed and unrelaxed compliances has rarely been used, though there are
142 Relaxation processes and physical aging

some examples of its application [e.g. 242]. One of the difficulties is the
additional imprecision in the data as a result of the differentiation, and
another is that the lack of any marked features, for example distinctive
curvature, in the differentiated creep curves makes their superposition
somewhat sUbjective. Nevertheless, a valid analysis of the effects of aging on
creep behaviour should consider the possible contributions of the limiting
compliances, and available approaches are either that of McCrum or that
due to Read [241], whereby allowance is made for the contribution of the
p-relaxation. If neither of these approaches is adopted, it should be borne in
mind that an assumption is being made about the effects of the limiting
compliances, for example that be = 1 if horizontal and vertical shifting of the
J(t) curves is used to obtain superposition. The values obtained for the
double logarithmic shift rate, f.1., to be discussed below, should then be
interpreted accordingly.
The alternative technique of dynamic mechanical analysis or of dynamic
mechanical thermal analysis (DMT A) may also be employed in the study of
the evolution of mechanical properties during physical aging. In the early
work of Kovacs, Stratton and Ferry [243], a torsion pendulum was used to
follow the changes in storage modulus G' and loss modulus G" of poly(vinyl
acetate) during isothermal aging in the ~ region. Whilst G' curves super-
posed reasonably well, G" data did not, suggestive of a departure from the
usual assumption of thermo rheological simplicity. The approach to equilib-
rium as measured by the dynamic mechanical properties, however, appeared
to occur on approximately the same time scale as did the volume relaxation
measured dilatometrically on the same sample.
More recently, though, aging studies have used DMTA rather than a
torsion pendulum for following the evolution of dynamic properties. Since
the most common mode of operation of DMT A is that of constant heating
rate, a procedure often adopted to study aging by DMT A is to anneal the
sample for a chosen time at a given temperature below ~ and then to scan
the aged sample by DMT A through the temperature range of interest. This
unfortunately introduces a rather complex thermal history which makes
interpretation of the data somewhat ambiguous; for example, it is well
known that such thermal histories in DSC studies can give rise to 'sub- ~
peaks' [10,237] and 'upper peaks' [62] in addition to the usual annealing
features.
In contrast, the isothermal changes in dynamic response following a
quench are easier to interpret and to compare with enthalpy and volume
relaxation. For example, for poly(methyl methacrylate) annealed both just
below [244] and far below ~ [225] after quenching from above ~, the
storage modulus E' increases and the loss modulus E" decreases monotoni-
cally during isothermal aging, as is also found for some polystyrene blends
Physical aging and mechanical properties 143

[245]. More complex thermal histories can give an isothermal response in


which memory effects appear in the evolution of dynamic mechanical
properties [225, 245, 246] similar to those reported by Kovacs in volume
relaxation [8] and by Struik in creep [18], confirming an apparently close
relationship between the evolution of dynamic mechanical properties and
the structural state of the glass.

3.4.3 PHYSICAL AGING TIME SCALES


The changes in mechanical properties, for example the creep compliance,
during phyical aging are presumed to be related to the structural changes
occurring within the material. These structural changes are identified most
commonly by means of either enthalpy or volume relaxation. The question
of which, if either, of these bulk structural properties is related to the
changes in mechanical properties, and in what way, is one for which an
unequivocal answer has yet to be found. In particular, the question of the
time scales required for the evolution of these various properties during
physical aging remains distinctly controversial. For example, Struik
[18,247], on the one hand, advocates that, on aging, the mechanical creep
behaviour shifts along the time scale by an amount determined uniquely by
the momentary value of the relative excess volume, (j (equation 3.1), and that
this applies generally to glassy polymers and for both simple and complex
thermal histories. On the other hand, there are other results, to be discussed
below, which suggest that the situation is not so simple.
A convenient way to quantify the rate of physical aging, as measured by
changes in the mechanical properties, is by means of the double logarithmic
shift rate p. suggested by Struik [18]:

d loga.
p.= (3.34)
d log t.

Here it is supposed that adequate superposition of the mechanical response,


for example the creep curves, can be achieved by a horizontal shift a., which
mayor may not be accompanied by a vertical shift and/or a change in
relaxation strength. Struik has shown that p. is very close to unity for a wide
variety of glassy polymers in a more or less wide range of temperature below
the glass transition, decreasing rapidly to zero at ~ itself.
In fact, for isothermal aging at a temperature sufficiently close to ~, it
should be possible to follow the evolution of the mechanical properties
until an equilibrium structural state is achieved. Under these conditions, the
shift rate p. will decrease as a function of aging time from a value which is
usually close to unity towards zero as equilibrium is approached. So also
144 Relaxation processes and physical aging

should it be possible to follow, either dilatometrically or calorimetrically, the


approach to equilibrium of the volume or enthalpy, respectively, of the same
glassy polymer under the same conditions, and in particular at the same
temperature. In this way, direct comparison of volume, enthalpy and the
evolution of the mechanical properties can be compared.
Surprisingly, such studies are relatively rare, and those results that are
available in the literature are not always in agreement. The torsion pendu-
lum studies of Kovacs et al. [243] suggest that, for poly(vinyl acetate) in the
1'g region, the viscoelastic properties approach limiting values, in other
words an equilibrium state, after elapsed times similar in magnitude to those
required for volume equilibration following down T-jumps. Their data for
up T-jumps are not strictly comparable for viscoelastic and volumetric
evolutions, however, as slightly different temperatures were used (39°C for
tan~, 40°C for volume); nevertheless, the fact that tan ~ at 39°C reaches
equilibrium following an up-jump at approximately 0.38 decades longer
time than does the volume at 40°C, which is reasonably consistent with an
activation energy of c. 790 kJ mol- 1 found from their other data, would
appear to confirm the equivalence of volumetric and viscoelastic time scales
for aging in poly(vinyl acetate). Similar conclusions in respect of creep
deformation have also been reached for polystyrene [242].
On the other hand, Perez et ai. [244] compare the aging of volume,
enthalpy and dynamic mechanical properties (tan~) of poly(methyl metha-
crylate), and find that the time scales for the evolution of these properties
increases in the order mechanical < volume < enthalpy. From a compari-
son of her own enthalpy relaxation data on polystyrene with the volume
relaxation data of Kovacs under approximately the same conditions, Petrie
[248] also reaches the conclusion that volume relaxation occurs more
rapidly than enthalpy relaxation. Similarly, Sasabe and Moynihan [68] find
for poly(vinyl acetate) the same result as above for the relative time
scales for the evolution of volume and enthalpy, namely that the volume
relaxation time is shorter, by a factor of about two, than the corresponding
enthalpy relaxation time. Additionally, they find that the mechanical shear
relaxation (and dielectric relaxation) evolves, at 1'g, much more rapidly than
does either the volume or the enthalpy, in qualitative agreement with the
results of Perez et ai. [244], but with time scale differences of the order of
two decades between volumetric and mechanical response, substantially
greater than that observed by Perez and coworkers.
These results contrast both with a comparison of changes in creep
behaviour with enthalpy relaxation in polystyrene [249] for which the
enthalpy attained equilibrium more rapidly than the creep response, and
also with a recent study of the aging of dynamic mechanical properties
of poly(vinyl acetate) in the 1'g region [250,251] which shows, for
Physical aging and mechanical properties 145

down T-jumps, the mechanical response evolving more slowly towards


equilibrium than the volumetric response. It was suggested that the rate of
physical aging measured by the evolution of the dynamic mechanical
properties may be influenced by the process of 'rejuvenation' [18], by which
the application of a mechanical strain modifies the structural state of the
glass; for example, in these down-jump experiments the free volume would
be increased by the mechanical strain, resulting in a further departure from
the equilibrium state, and hence a lengthening of the time scale for the aging
of the mechanical response. Under these circumstances, the aging rate
determined mechanically would be anticipated to be reduced, and indeed the
double logarithmic shift rate J1 was found [250,251] to be as low as 0.3 at
an aging temperature of 35°C, increasing to approximately 0.5 at 32SC.
These results suggest that the effect of rejuvenation could seriously
complicate the interpretation of physical aging time scales determined
mechanically, and this possibility should always be borne in mind, though
the process of rejuvenation is not universally accepted. In this respect, some
interesting experiments have recently been performed by McKenna and
coworkers [252, 253], using a torsional dilatometer to measure simulta-
neously the volumetric and viscoelastic response of epoxy resin samples
subjected to carefully controlled thermomechanical histories, including
both down-jumps and up-jumps in temperature. These authors find that, for
down-jumps, the mechanical response attains an equilibrium before the
volume, whereas for up-jumps the opposite is true. Moreover, the torsional
mechanical deformations do not affect the volume recovery after aT-jump,
as illustrated by the results reproduced in Figure 3.21. Here it can be seen
that, despite significant magnitudes of excess volume excursions induced by
the torsional strains at equal intervals (multiples of two) of logarithmic
aging time, the underlying volume recovery kinetics is unaffected by the
mechanical history. The structure of the glassy state would seem to be
decoupled in some way from the mechanical stress field.
This area of physical aging appears to be one in which there remains a
significant number of unanswered questions. Even in the linear viscoelastic
region, the interaction of mechanical stimulus and structural (volumetric or
enthalpic) state is open to debate in respect of rejuvenation. Likewise, time
scales for the relaxation of different properties (mechanical, structural) do
not appear to follow a systematic pattern, though differences in materials
and relative aging temperatures no doubt play an important role. Ultimately
the interpretation of physical aging of mechanical properties by means of
free volume concepts seems difficult to sustain if the volumetric state is
indeed decoupled from the mechanical stress; similar problems with the free
volume concept were noted in sections 3.2.5 and 3.2.6 above in respect of
small angle X-ray scattering studies and of the effect of combined pressure
146 Relaxation processes and physical aging

IllT= - 5.3 ·ei


,., 0 .50 1'>'=0.05 I
o
x

,IIIIII
a
>
:::::::, 0.25
>
I
......>
0 .00

- 0.25 L..-.........................""'-___.....................t.____-'-'"...........L -........................,uJ

10 2

Figure 3.21 Volume response of epoxy resin sample in the torsional dilatometer
when subjected to a quench from equilibrium at 40.8°C to an aging temperature of
35.soC (Tg = 42.4°C by DSC). The volume excursions are caused by torsions applied
at equal logarithmic intervals of aging time, with a torsional strain of 0.05 and a
duration of one-tenth of the aging time. (Reproduced from reference 253, with
permission .)

and temperature histories. The impact of physical aging on higher strain


properties, for example non-linear viscoelasticity and yield behaviour, which
is clearly of considerable interest, would benefit from a better understanding
of the effects in the linear viscoelastic region. It is to be hoped that the
significant amount of work currently being undertaken in this area will at
least begin to answer some of the unresolved questions discussed here. The
influence of physical aging on yield and post-yield deformation is discussed
in Chapters 4 and 5.

REFERENCES
1. Tammann, G. (1925/ 1926) Glastech. Ber., 3, 73 - 87.
2. Tammann, G. (1930) Der Glaszustand, L. Voss, Leipzig.
3. Simon, F. (1930) Ergeb. exact Naturwiss., 9,222- 74.
4. Simon, F. (1931) Z. anorg. allgern. Chern., 203, 220-7.
5. Kauzmann, W. (1948) Chern. Rev., 43, 219- 56.
6. Davies, R.O. and Jones, G.O. (1953) Proc. Roy. Soc. London A, 217, 26- 42.
7. Davies, R.O. and Jones, G.O. (1953) Adv. Phys., 2, 370- 410.
References 147

8. Kovacs, A.1. (1963) Fortschr. Hochpolym. Forsch., 3, 394-507.


9. Rehage, G. and Borchard, W. (1973) in The Physics of Glassy Polymers (ed.
R.N. Haward), Applied Science Publishers, London, Ch. 1, pp. 54-107.
10. Hodge, I.M. (1994) J. Non-Cryst. Solids, 169, 211-66.
11. Ngai, K.L. and Wright, G.B. (eds) (1984) Relaxations in Complex Systems.
Proceedings of a Workshop held at Virginia Polytechnic Institute and State
University, Blacksburg, Virginia, USA in July 1983, National Technical Infor-
mation Service, US Department of Commerce, Springfield, VA.
12. Dorfmiiller, Th. and Williams, G. (eds) (1987) Molecular Dynamics and Relax-
ation Phenomena in Glasses. Proceedings of a Workshop held at Zentrum fiir
interdisziplinare Forschung, Universitat Bielefeld, Germany in November 1985.
Lecture Notes in Physics, 277, Springer-Verlag, Berlin.
13. Ngai, K.L. and Wright, G.B. (eds) (1991) Relaxations in Complex Systems.
Proceedings of International Discussion Meeting held in Aghia Pelaghia,
Heraklion, Crete, Greece in June 1990. J. Non-Cryst. Solids, 131-133,1-1285.
14. Ngai, K.L. and Wright, G.B. (eds) (1994) Relaxations in Complex Systems 2.
Proceedings of 2nd International Discussion Meeting held in Alicante, Spain in
June/July 1993. J. Non-Cryst. Solids, 172-174, 1-1457.
15. Heydemann, P. and Guicking, H.D. (1963) Koll. Z. Z. Polym., 193,16-25.
16. Goldbach, G. and Rehage, G. (1967) Koll. Z. Z. Polym., 216/217, 56-63.
17. Wittmann, J.c. and Kovacs, A.1. (1969) J. Polym. Sci. C, 16,4443-52.
18. Struik, L.C.E. (1978) Physical Aging in Amorphous Polymers and Other Mater-
ials, Elsevier, Amsterdam.
19. Greiner, R. and Schwarzl, F.R. (1984) Rheo!. Acta, 23, 378-95.
20. Struik, L.C.E. (1987) Polymer, 28, 1869-75.
21. Lee, H.H.D. and McGarry, F.1. (1990) J. Macromol. Sci. B: Phys., 29,185-202.
22. Nakayama, M., Hanaya, M., Hatate, A. and Oguni, M. (1994) J. Non-Cryst.
Solids, 172-174, 1252-61.
23. McCrum, N.G., Read, B.E. and Williams, G. (1967) Anelastic and Dielectric
Effects in Polymeric Solids, John Wiley, New York.
24. Hutchinson, J.M. and Kovacs, A.1. (1976) J. Polym. Sci., Polym. Phys. Edn, 14,
1575-90.
25. Kohlrausch, F. (1866) Annalen der Physik und Chemie, 128, 1-20, 207-227,
399-419.
26. Williams, G. and Watts, D.C. (1970) Trans. Faraday Soc., 66, 80-5.
27. Kovacs, A.J., Aklonis, J.1., Hutchinson, J.M. and Ramos, A.R. (1979) J. Polym.
Sci., Polym. Phys. Edn, 17, 1097-162.
28. Tool, A.Q. (1946) J. Am. Ceram. Soc., 29, 240-53.
29. Narayanaswamy, O.S. (1971) J. Am.::eram. Soc., 54, 491-8.
30. Moynihan, C.T., Easteal, A.J., DeBolt, M.A. and Tucker, J. (1976) J. Am.
Ceram. Soc., 59, 12-6.
31. Hutchinson, J.M. and Kovacs, A.1. (1977) in The Structure of Non-Crystalline
Materials (ed. P.H. Gaskell), Taylor and Francis, London, pp. 167-72.
32. Ngai, K.L. (1979) Comments on Solid State Physics, 9, 127-39.
33. Ngai, K.L. (1980) Comments on Solid State Physics, 9,141-55.
34. Robertson, R.E., Simha, R. and Curro, J.J. (1984) Macromolecules, 17, 911-9.
35. Ngai, K.L., Rendell, R.W., Rajagopal, A.K. and Teitler, S. (1986) Ann. NY Acad.
Sci., 484, 150-80.
36. Hodge, I.M. and O'Reilly, J.M. (1992) Polymer, 33, 4883.
37. Cortes, P. (1994) Study of the structural relaxation of linear polyesters by
thermal analysis, PhD Thesis, Universidad Politecnica de Catalufia, Barcelona,
Spain (in Spanish).
148 Relaxation processes and physical aging

38. Gibbs, lH. and DiMarzio, E.A. (1958) J. Chern. Phys., 28, 373-83.
39. Adam, G. and Gibbs, lH. (1965) J. Chern. Phys., 43, 139-46.
40. Mathot, V.B.F. (1984) Polymer, 25, 579-99. .
41. Hodge, I.M. (1987) Macromolecules, 20, 2897-908.
42. Kovacs, A.J., Hutchinson, J.M. and Aklonis, J.J. (1977) in The Structure of
Non-Crystalline Materials (ed. P.H. Gaskell), Taylor and Francis, London,
pp.153-63.
43. Williams, M.L., Landel, R.F. and Ferry, J.D. (1955) J. Am. Chern. Soc., 77,
3701-7.
44. Doolittle, A.K. (1951) J. Appl. Phys., 22,1471-5.
45. Cohen, M.H. and Turnbull, D. (1959) J. Chern. Phys., 31, 1164-9.
46. Turnbull, D. and Cohen, M.H. (1961) J. Chern. Phys., 34,120-5.
47. Kovacs, AJ. and Hutchinson, 1M. (1979) J. Polym. Sci., Polym. Phys. Edn, 17,
2031-58.
48. Hutchinson, 1M. (1987) Lecture Notes in Physics, 277, 172-87.
49. Hutchinson, 1M. and Ruddy, M. (1989) Macromol. Chern., Macromol. Symp.,
27,319-30.
50. O'Reilly, J.M., Tribone, J.J. and Greener, J. (1984) Bull. Am. Phys. Soc., 29-30,
326.
51. Chow, T.S. and Prest, W.M. Jr (1982) J. Appl. Phys., 53, 6568-73.
52. Oudhuis, A.A.e.M. and ten Brinke, G. (1992) Macromolecules, 25, 698-702.
53. O'Reilly, 1M. (1987) Crit. Rev. Solid State Mater. Sci., 13,259-77.
54. Gomez Ribelles, J.L., Ribes Greus, A. and Diaz Calleja, R. (1990) Polymer, 31,
223-30.
55. Moynihan, e.T., Crichton, S.N. and Opalka, S.M. (1991) J. Non-Cryst. Solids,
131-133,420-34.
56. Rekhson, S. and Ducroux, loP. (1992) in The Physics of Non-Crystalline Solids
(eds L. Pye, W.e. La Course and HJ. Stevens), Taylor and Francis, London,
pp.315-20.
57. Ducroux, l-P., Rekhson, S.M. and Merat, F.L. (1994) J. Non-Cryst. Solids,
172-174,541-53.
58. Hodge, I.M. and Huvard, G.S. (1983) Macromolecules, 16, 371-5.
59. Privalko, V.P., Demchenko, S.S. and Lipatov, Yu. S. (1986) Macromolecules, 19,
901-4.
60. Prest, W.M., Roberts, FJ. and Hodge, I.M. (1983) Proc. 12th NATAS Confer-
ence (ed. J.C. Buck), Williamsburg, VA, p.119.
61. Hutchinson, J.M. and Ruddy, M. (1988) J. Polym. Sci., Polym. Phys. Edn, 26,
2341-66.
62. Hutchinson, J.M. and Ruddy, M. (1990) J. Polym. Sci., Polym. Phys. Edn, 28,
2127-63.
63. Tribone, J.J., O'Reilly, 1M. and Greener, 1 (1986) Macromolecules, 19, 1732-9.
64. Mijovic, 1, Nicolais, L., D'Amore, A. and Kenny, 1M. (1994) Polym. Eng. Sci.,
34,381-9.
65. Hutchinson, J.M., Ingram, M.D. and Pappin, AJ., submitted to Macro-
molecules.
66. Hodge, I.M. and Berens, A.R. (1982) Macromolecules, 15, 762-70.
67. Pappin, A.J., Hutchinson, J.M. and Ingram, M.D. (1992) Macromolecules, 25,
1084-89.
68. Sasabe, H. and Moynihan, e.T. (1978) J. Polymer Sci., Polymer Phys. Edn, 16,
1447-57.
69. Montserrat, S., Cortes, P., Pappin, A.J., Quah, K.H. and Hutchinson, J.M.
(1994) J. Non-Cryst. Solids, 172-174, 1017-22.
References 149

70. Montserrat, S., unpublished data.


71. Hutchinson, J.M., McCarthy, D., Montserrat, S. and Cortes. P. (1996) J.
Polymer Sci., Polymer Phys. Edn, 34, 229-39.
72. Cortes, P., Montserrat, S. and Hutchinson, J.M. (1995) Proc. Workshop on
Non-Equilibrium Phenomena in Supercooled Fluids. Glasse" and Amorphous
Materials, Pisa, Italy, 25-29 September.
73. Billmeyer, F.W. Jr (1971) Textbook of Polymer Science, 2nd edn, Wiley
Interscience, New York.
74. Hodge, I.M. (1991) J. Non-Cryst. Solids, 131-133,435-41.
75. Angell, e.A. (1984) in Relaxations in Complex Systems (eds K.L. Ngai and G.B.
Wright), National Technical Information Serice, US Department of Commerce,
Springfield, VA, pp. 3-11.
76. Bohmer, R, Ngai, K.L., Angell, e.A. and Plazek, D.1. (1993) J. Chem. Phys., 99,
4201-9.
77. Plazek, D.1. and Ngai, K.L. (1991) Macromolecules, 24, 1222--4,
78. Hodge, I.M. (1983) Macromolecules, 16, 898-902.
79. Ingram, M.D., Hutchinson, J.M. and Pappin, A.1. (1991) Ph.vs. Chem. Glasses,
32,121-8.
80. Wendorff, J.H. and Fischer, EW. (1973) Kolloid Z. Z. Polym., 251, 876-83.
81. Curro, J. J. and Roe, R-J. (1984) Polymer, 25, 1424-30.
82. Miiller, J. and Wendorff, J.H. (1988) J. Polym. Sci., Polym. Lett. Edn, 26, 421-7.
83. Roe, R.-J. and Curro, 1.1. (1983) Macromolecules, 16, 428-34.
84. Song, H.-H. and Roe, R.-J. (1987) Macromolecules, 20, 2723-32.
85. Haward, RN. (1973) in The Physics of Glassy Polymers (ed. R.N. Haward),
Applied Science Publishers, London, pp. 1-53.
86. Lamarre, L. and Sung, e.S.P. (1983) Macromolecules, 16,1729--36.
87. Yu, W.-e., Sung, e.S.P. and Robertson, RE. (1988) Macromolecules, 21,
355-64.
88. Sung, e.S.P., Gould, I.R. and Turro, N.J. (1984) Macromolecules, 17, 1447-51.
89. Victor, J.G. and Torkelson, J.M. (1987) Macromolecules, 20,2241-50.
90. Victor, J.G. and Torkelson, J.M. (1988) Macromolecules, 21, 3490-7.
91. Royal, J.S., Victor, J.G. and Torkelson, J.M. (1992) Macromolecules, 25,
729-34.
92. Royal, J.S. and Torkelson, J.M. (1992) Macromolecules, 25, 4792-6.
93. Hutchinson, J.M. (1995) Prog. Polym. Sci., 20, 703-60.
94. Loutfy, R.O. (1981) Macromolecules, 14,270-5.
95. Meyer, E.F., Jamieson, A.M., Simha, R. et al. (1990) Polymer, 31, 243-7.
96. Royal, J.S. and Torkelson, J.M. (1992) Macromolecules, 25,1705-10.
97. Royal, J.S. and Torkelson, J.M. (1993) Macromolecules, 26,5331-5.
98. Hasan, O.A., Boyce, M.e., Li, X.S. and Berko, S. (1993) J. Polym. Sci., Polym.
Phys. Edn, 31, 185-97.
99. Wang, y.Y., Nakanishi, N., Jean, Y.e. and Sandreczki, T.e. (1990) J. Polym.
Sci., Polym. Phys. Edn, 28,1431-41.
100. Jean, Y.e. and Deng, Q. (1992) J. Polym. Sci., Polym. Phys. Edn, 30, 1359-64.
101. Deng, Q., Sundar, e.s. and Jean, Y.e. (1992) J. Phys. Chem., 96, 492-5.
102. Deng, Q. and Jean, Y.e. (1993) Macromolecules, 26,30-4.
103. Kobayashi, Y., Zheng, W., Meyer, E. F. et al. (1988) Proceedings of 8th ICPA,
Gent, Belgium, World Scientific Publishing, pp. 812-4.
104. Kobayashi, Y., Zheng, W., Meyer, E. F. et al. (1989) Macromolecules, 22,
2302-6.
105. McGervey, J.D., Panigrahi, N., Simha, R. and Jamieson, A.M. (1985) Proc. 7th
ICP A, New Delhi, India, World Scientific Publishing, pp. 690-1.
150 Relaxation processes and physical aging

106. Hill, AJ., Jones, P.L., Lind, lH. and Pearsall, G.W. (1988) J. Polym. Sci. A:
Polym. Chern., 26, 1541-52.
107. Kluin, J.E., Yu, Z., Vleeshouwers, S. et al. (1992) Macromolecules, 25,5089-93.
108. Liu, L.B., Gidley, D. and Vee, A.F. (1992) J. Polym. Sci., Polym. Phys. Edn, 30,
231-8.
109. O'Connor, PJ., Kochner, C.W., Landes, B.G. et al. (1992) Polym. Preprints, 33,
308-9.
110. Kluin, lE., Yu, Z., Vleeshouwers, S. et al. (1993) Macromolecules, 26, 1853-61.
111. Li, X.S. and Boyce, M.C. (1993) J. Polym. Sci., Polym. Phys. Edn, 31, 869-73.
112. Brandt, W. and Wilkenfield, l (1975) Phys. Rev., 812, 2579-87.
113. Rehage, G. and Goldbach, G. (1965) Kolloid Z. Z. Polym., 206, 166-8.
114. Rehage, G. and Goldbach, G. (1966) Ber. Bunsenges. Phys. Chern., 70, 1144-8.
115. Goldbach, G. and Rehage, G. (1967) J. Polymer Sci. C, 16,2289-98.
116. Goldbach, G. and Rehage, G. (1967), Rheol. Acta, 6, 30-53.
117. Tribone, lJ., O'Reilly, lM. and Greener, l (1989) J. Polym. Sci., Polym. Phys.
Edn, 27, 837-57.
118. McKinney, lE. and Goldstein, M. (1974) J. Res. Nat. Bur. Std, 78A, 331-53.
119. Hirai, N. and Eyring, H. (1959) J. Polym. Sci., 37, 51-70.
120. Shishkin, N. (1960) Soviet Phys. Solid State, 2, 322-8.
121. Hellwege, K.-H., Knappe, W. and Lehmann, P. (1962) Kolloid Z. Z. Polym.,
183, 11 0-20.
122. Bianchi, U. (1965) J. Phys. Chern., 69, 1497-504.
123. Karasz, F.E., Bair, H.E. and O'Reilly, lM. (1965) J. Phys. Chern., 69, 2657-67.
124. Abu-Isa, I. and Dole, M. (1965) J. Phys. Chern., 69, 2668-75.
125. Gee, G. (1966) Polymer, 7, 177-91.
126. Breuer, H. and Rehage, G. (1967) Kolloid Z. Z. Polym., 216-217, 159-79.
127. Ichihara, S., Komatsu, A., Tsujita, Y. et al. (1971) Polym. J., 2, 530-4.
128. Quach, A. and Simha, R (1971) Macromolecules, 4, 268-70.
129. Billinghurst, P.R and Tabor, D. (1971) Polymer, 12, 101-17.
130. Quach, A. and Simha, R (1971) J. Appl. Phys., 42, 4592-606.
131. Quach, A. and Simha, R (1972) J. Phys. Chern., 76, 416-21.
132. Sasabe, H. and Saito, S. (1972) Polym. J., 3, 749-55.
133. Weitz, A. and Wunderlich, B. (1974) J. Polym. Sci., Polym. Phys. Edn, 12,
2473-91.
134. Oels, H.-J. and Rehage, G. (1977) Macromolecules, 10, 1036-43.
135. Stevens, J.R, Oakley, RW., Chau, K.W. and Hunt, J.L. (1986) J. Chern. Phys.,
84, 1006-14.
136. O'Reilly, lM. (1962) J. Polym. Sci., 57, 429-44.
137. Zosel, A. (1964) Kolloid Z. Z. Polym., 199, 113-25.
138. Bianchi, u., Turturro, A. and Basile, G. (1967) J. Phys. Chern., 71, 3555-8.
139. Sasabe, H. and Saito, S. (1972) Polym. J., 3, 631-9.
140. Parry, E.J. and Tabor, D. (1973) Polymer, 14, 623-7.
141. McKinney, lE. and Simha, R. (1977) J. Res. Nat. Bur. Std, 81A, 283-97.
142. Naoki, M. and Owada, A. (1984) Polymer, 25, 75-83.
143. McKinney, lE. and Belcher, H.V. (1963) J. Res. Nat. Bur. Std, 67A, 43-53.
144. Parry, EJ. and Tabor, D. (1973) Polymer, 14, 628-31.
145. McKinney, lE. and Simha, R (1974) Macromolecules, 7, 894-901.
146. Quach, A., Wilson, P.S. and Simha, R. (1974) J. Macromol. Sci. B: Phys., 9,
533-50.
147. Olabisi, O. and Simha, R. (1975) Macromolecules, 8,206-10.
148. Skorodumov, V.F. and Godovskii, Yu, K. (1993) Vysokomol. Soed. Ser. B, 35,
214-26.
References 151

149. Ichihara, S., Komatsu, A and Hata, T. (1971) Polym. J., 2, 650--5.
150. Zoller, P. and Bolli, P. (1980) J. Macromol. Sci. B: Phys., 18, 549-62.
151. Fakhreddine, Y.A. and Zoller, P. (1991) J. Polym. Sci., Polym. Phys. Edn, 29,
1141-6.
152. Zoller, P. (1978) J. Polym. Sci., Polym. Phys. Edn, 16, 1261-75.
153. Maeda, Y., Karasz, F.E., MacKnight, W.1. and Vukovic, R. (1986) J. Polym.
Sci., Polym. Phys. Edn, 24, 2345-57.
154. Parry, E.J. and Tabor, D. (1973) Polymer, 14, 617-22.
155. Yoon, H.N., Pae, K.D. and Sauer, J.A. (1976) J. Polym. Sci., Polym. Phys. Edn,
14, 1611-27.
156. Zoller, P., Kehl, T.A., Starkweather, H.w. Jr and Jones, G.A. (1989) J. Polym.
Sci., Polym. Phys. Edn, 27, 993-1007.
157. Pae, K.D., Tang, c.-L. and Shin, E.-S. (1984) J. Appl. Phys., 56, 2426-32.
158. Singh, H. and Nolle, A.W. (1959) J. Appl. Phys., 30, 337-41.
159. Nolle, AW. and Billings, J.J. (1959) J. Chem. Phys., 30, 84-90.
160. Anderson, J.E., Davis, D.D. and Slichter, W.P. (1969) Macromolecules, 2,
166-9.
161. Payne, A.R. (1958) in Rheology of Elastomers (eds P. Mason and N. Wookey),
Pergamon, London, p. 86. .
162. McKinney, J.E., Belcher, H.V. and Marvin, R.S. (1960) Trans. Soc. Rheol., 4,
347-62.
163. Sasuga, T. and Takehisa, M. (1977) J. Macromol. Sci. B: Phys., 13, 215-29.
164. Dalal, E.N. and Philips, P.1. (1983) Macromolecules, 16, 890-7.
165. Pae, K.D. and Questad, D.L. (1983) J. Polym. Sci., Polym. Phys. Edn, 21,
1195-203.
166. Pappin, A.J., Ingram, M.D., Hutchinson, J.M. et al. (1995) Phys. Chem. Glasses,
36,164-71.
167. Hutchinson, J.M., Ingram, M.D. and Robertson, A.H.1. (1992) Phil. Mag., 66,
449-61.
168. Price, c., Williams, R.c. and Ayerst, R.c. (1972) in Amorphous Materials (eds
R.N. Douglas and B. Ellis), Wiley-Interscience, London, pp.117-24.
169. Yourtree, J.B. and Cooper, S.L. (1974) J. Appl. Polym. Sci., 18,897-912.
170. Brown, I.G., Wetton, R.E., Richardson, M.1. and Savill, N.G. (1978) Polymer,
19,659-63.
171. Kogowski, G.J. and Filisko, F.E. (1986) Macromolecules, 19, 828-33.
172. Price, C. (1975) Polymer, 16, 585-9.
173. Bree, H.w., Heijboer, J., Struik, L.C.E. and Tak, A.G.M. (1974) J. Polym. Sci.,
Polym. Phys. Edn, 12, 1857-64.
174. Allen, G., Ayerst, R.C., Cleveland, J.R. et al. (1968) J. Polym. Sci. C, 23, 127-9.
175. Ichihara, S., Komatsu, A and Hata, T. (1971) Polym. J., 2, 644-9.
176. Kimmel, R.M. and Uhlmann, D.R. (1970) J. Appl. Phys., 41, 2917-27.
177. Kimmel, R.M. and Uhlmann, D.R. (1971) J. Appl. Phys., 42, 4917-25.
178. Prest, W.M. Jr and Roberts, F.1. Jr (1981) Ann. NY Acad. Sci., 371,67-86.
179. Prest, W.M. Jr, O'Reilly, J.M., Roberts, F.J. Jr and Mosher, R.A. (1981) Polym.
Eng. Sci., 21, 1181-7.
180. Mackenzie, J.D. (1964) J. Am. Ceram. Soc., 47, 76-80.
181. O'Reilly, J.M. and Mosher, R.A. (1980) J. Appl. Phys., 51, 5137-9.
182. O'Reilly, J.M. and Mosher, R.A. (1981) J. Polym. Sci., Polym. Phys. Edn, 19,
1187-98.
183. Destruel, P., Ai, B. and Hoang-The-Giam (1984) J. Appl. Phys., 55,2726-32.
184. Heijboer, J. (1977) Intern. J. Polym. Mat., 6,11-37.
185. Illers, K.H. and Breuer, J. (1961) Koll. Z., 176, 110-19.
152 Relaxation processes and physical aging

186. Jho, J.Y. and Vee, A.F. (1991) Macromolecules, 24,1905-13.


187. Vee, A.F. and Smith, S.A. (1981) Macromolecules, 14, 54-64.
188. Schaefer, 1., Stejskal, E.O., Perchak, D. et al. (1985) Macromolecules, 18,
368-73.
189. Vee, A.F. (1977) Polymer Eng. Sci., 17,213-19.
190. Floudas, G., Higgins, 1.S., Meier, G., Kremer, F. and Fischer, E.W. (1993)
Macromolecules, 26, 1676-82.
191. Schartel, B. and Wendorff, 1.H. (1995) Polymer, 36, 899-904.
192. Jones, A.A. (1985) Macromolecules, 18, 902-6.
193. Hoff, E.A.W., Robinson, D.W. and Willbourn, A.H. (1955) J. Polym. Sci., 18,
161-76.
194. Heijboer, 1. (1969) Br. Polym. J., 1, 3~ 14.
195. Williams, G. (1966) Trans. Faraday Soc., 62, 2091.
196. Dionisio, M.S., Mouro-Ramos, 1.1. and Williams, G. (1994) Polymer, 35,
1705-13.
197. Diaz-Calleja, R., Devine, I., Gargallo, L. and Radic, D. (1994) Polymer, 35,
151-6.
198. Heijboer, 1., Baas, 1.M.A., van de Graaf, B. and Hoefragel, M.A. (1992) Polymer,
33, 1359-62.
199. Heijboer, 1. (1968) J. Polym. Sci. C, 16, 3413-22.
200. Illers, K.H. and Jenckel, E. (1958) Rheol. Acta, 1, 322.
201. Illers, K.H. and Jenckel, E. (1959) J. Polym. Sci., 41, 528-31.
202. Yano, O. and Wada, Y. (1971) J. Polym. Sci., Polym. Phys. Edn, 9, 669-86.
203. Yano, O. and Wada, Y. (1974) J. Polym. Sci., Polym. Phys. Edn, 12, 665-83.
204. Sinnott, K.M. (1962) SPE Trans., 2, 65.
205. McCammon, R.D., Saba, R.G. and Work, R.N. (1969) J. Polym. Sci., Polym.
Phys. Edn, 7,1721-33.
206. Irvine, 1.D. and Work, R.N. (1972) J. Polym. Sci., Polym. Phys. Edn, 11, 175-91.
207. Reich, S. and Eisenberg, A. (1972) J. Polym. Sci., Polym. Phys. Edn, 10,
1397-400.
208. Wall, R.A., Sauer, J.A. and Woodward, A.E. (1959) J. Polym. Sci., Polym. Phys.
Edn, 35, 281-4.
209. Shimizu, K., Yano, O. and Wada, Y. (1973) Polym. J., 5,107-9.
210. Abe, Y., Tonelli, A.E. and Flory, P.J. (1970) Macromolecules, 3, 294-303.
211. Tonelli, A.E. (1973) Macromolecules, 6,682-3.
212. Hiigele, P.e. and Beck, L. (1977) Macromolecules, 10, 213-15.
213. Tanabe, Y. (1985) J. Polym. Sci., Polym. Phys. Edn, 23, 601-6.
214. Vol'kenshtein, M.V., Kol'tsov, A.I. and Khachaturov, A.S. (1965) Vysokomol.
Soedin., 7, 296.
215. Alexandrovich, P.S., Karasz, F.E. and MacKnight, W.J. (1980) Polymer, 21,
488-94.
216. Kulik, A.S. and Prins, K.O. (1993) Polymer, 34, 4635-41.
217. Mitchell, G.R. and Windle, A.H. (1984) Polymer, 25, 906-20.
218. Khare, R., Paulitis, M.E. and Lustig, S.R. (1993) Macromolecules, 26, 7203-9.
219. Golden, J.H., Hammant, B.L. and Hazell, E.A. (1967) J. Appl. Polym. Sci., 11,
1571-9.
220. Allen, G., Morley, D.e.W. and Williams, T. (1973) J. Mater. Sci., 8, 1449-52.
221. Watts, D.C. and Perry, E.P. (1978) Polymer, 19, 248-54.
222. Johari, G.P. and Goldstein, M. (1970) J. Chem. Phys., 53, 2372-88.
223. Johari, G.P. (1973) J. Chem. Phys., 58, 1766-70.
224. Johari, G.P. (1982) J. Chem. Phys., 77, 4619-26.
References 153

225. Diaz-Calleja, R, Ribes-Greus, A. and Gomez-Ribelles, J.L. (1989) Polymer, 30,


1433-8.
226. Guerdoux, L. and Marchal, E. (1981) Polymer, 22, 1199-204.
227. Dean, G.D., Read, B.E. and Small, G.D. (1988) Plast. Rubb. Compo Proc. Appl.,
9, 173-9.
228. Read, B.E., Tomlins, P.E. and Dean, G.D. (1990) Polymer, 31, 1204-15.
229. Beiner, M., Garwe, F., Schroter, K. and Donth, E. (1994) Polymer, 35, 4127-32.
230. Heijboer, J. (1978) in Molecular Basis of Transitions and Relaxations (ed. DJ.
Meier), Gordon and Breach, London, p. 75.
231. Perez, J., Muzeau, E. and Cavaille, J.Y. (1992) Plast. Rubb. Compo Proc. Appl.,
18,139-48.
232. Muzeau, E., Cavaille, J.Y., Vassoille, R. et al. (1992) Macromolecules, 25,
5108-10.
233. Muzeau, E., Vigier, B., Vassoille, R. and Perez, 1. (1995) Polymer, 36, 611-20.
234. Struik, L.CE. (1987) Polymer, 28, 57-68.
235. Muzeau, E. and lohari, G.P. (1990) Chem. Phys., 149, 173-83.
236. Muzeau, E., Perez, 1. and lohari, G.P. (1991) Macromolecules, 24, 4713-23.
237. Ruddy, M. and Hutchinson, I.M. (1988) Polym. Commun., 29, 132-4.
238. Pappin, AJ., Hutchinson, J.M. and Ingram, M.D. (1994) J. Non-Cryst. Solids,
172-174,584-91.
239. McLoughlin, I.R and Tobolsky, A.V. (1951) J. Polym. Sci., 7, 658.
240. Chai, CK. and McCrum, N.G. (1980) Polymer, 21, 706-12.
241. Read, B.E. (1991) J. Non-Cryst. Solids, 131-133,408-19.
242. Hutchinson, I.M. and Kriesten, U. (1993) in Macromolecules 1992 (ed. J.
Kahovec), VSP International Publishers, Zeist, The Netherlands, pp. 45-54.
243. Kovacs, A.l., Stratton, RA. and Ferry, J.D. (1963) J. Phys. Chem., 67, 152-61.
244. Perez, J., Cavaille, 1.Y., Diaz-Calleja, R. et al. (1991) Makromol. Chem., 192,
2141-61.
245. Cavaille, 1.Y., Etienne, S., Perez, J., et al. (1986) Polymer, 27, 686-92.
246. Haidar, B. and Smith, T.L. (1990) Macromolecules, 23, 3710-12.
247. Struik, L.CE. (1988) Polymer, 29, 1347-53.
248. Petrie, S.E.B. (1972) J. Polym. Sci., Polym. Phys. Edn, 10, 1255-72.
249. Roe, R.-J. and Millman, G.M. (1983) Polym. Eng. Sci., 23, 318-22.
250. Hutchinson, I.M., Singh, J., Rychwalski, RW. et al. (1995) Proceedings of 1st
International Conference on Mechanics of Time-Dependent Afaterials (eds I.
Emri and G. Knauss), Ljubljana, Society for Experimental Mechanics, Inc.,
Connecticut, USA, pp. 67 - 72.
251. Delin, M., Rychwalski, RW., Kubat, 1. et al. (1996) Polym. Eng. Sci., 36,
2955-67.
252. McKenna, G.B. (1994) J. Non-Cryst. Solids, 172-174, 756·-64.
253. McKenna, G.B., Leterrier, Y. and Schultheisz, CR. (1995) Polym. Eng. Sci., 35,
403-10.
Yield processes
in glassy polymers 4
B. Crist

4.1 INTRODUCTION
Virtually all solid polymers - amorphous polymers below the glass transi-
tion temperature ~ or crystalline polymers below the melting temperature
Tm - undergo a permanent shape change when subjected to a stress of
sufficient magnitude. This chapter focuses on the onset of irreversible
deformations in glassy polymers, i.e. the transition from elastic to plastic
behaviour. Reversible elastic and linear viscoelastic deformations, which
result from small stresses with corresponding strains e < 0.01, were treated
in Chapter 3. Chapter 5 addresses large irreversible strains and associated
phenomena that develop after the transition.
The onset of plastic deformation is termed yielding, which reflects the
observation that mechanical compliance increases abruptly at a certain
point. Figure 4.1 from recent work of Hasan and Boyce (1993) illustrates
many mechanical features of glassy polymers, exemplified here by atactic
polystyrene (PS), a common amorphous thermoplastic generally thought to
be 'brittle' at room temperature. The uniaxial compression test is described
in section 4.2.2; for now be reminded that strain is a fractional length
change, and the deformation force is expressed as stress (MPa =
106 N m - 2). First note that PS can be compressed to 74% of the original
length without fracture, and that the vast majority of the deformation
remains after unloading. Glassy PS is clearly ductile in compression. Here
we concentrate on yield, which is associated with the maximum in load or
stress. Strains imposed below the stress maximum are recoverable (elastic or
viscoelastic), while plastic strain is seen to develop at and beyond the stress
maximum. Another characteristic of glassy polymers is history dependence;
curves in Figure 4.1 are for samples either quenched or slowly cooled from

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
156 Yield processes in glassy polymers

100 .-r---r---..,----r---r--"""T'""---r---.

50

0.0 0.1 0.2 0.3


TRUE STRAIN

Figure 4.1 Uniaxial compression of polystyrene at 23°e, strain rate Et =


-1 x 10- 3/s. The upper curve is for an annealed (physically aged) glass, while the
lower curve is for a glass quenched from the melt. Response to unloading is shown
after different imposed strains; plastic or irreversible deformation is seen after the
stress maximum. (Reproduced from Hasan and Boyce, 1993, with permission.)

the liquid state ('annealed'). Yield stress and the subsequent stress drop are
clearly altered by sample history, while the flow stress at large strains is not.
Yielding is of tremendous technological importance, defining upper limits
of service stress in load-bearing applications or the conditions required for
shaping parts during manufacturing. The scientific basis of yielding is well
understood for crystalline materials in which planes of atoms slide over one
another to new equilibrium positions as sketched in Figure 4.2. Such
irreversible deformation can occur at moderate stresses (below 25 MPa) in
metals because certain crystallographic defects, i.e. dislocations, facilitate
rearrangement from the undeformed state in Figure 4.2a to the deformed
state Figure 4.2c (Kelly and Macmillen, 1986; Weertman and Weertman,
1992). Atomic and particularly polymeric glasses are also capable of
undergoing substantial plastic deformation. Since these materials are by
Introduction 157

--
(01 '1= 0 - (bl 't= 0.5 lei '1= to
Figure 4.2 Model for elementary shear displacement; rows of circles represent
planes of atoms or polymer chains. Undeformed (a) and deformed (c) states are
stable, while (b) represents the unstable intermediate state.

definition positionally disorderd solids lacking a well defined equilibrium


state, the validity of the elementary shear displacement model in Figure 4.2
may be questioned.
Indeed, the fundamental nature of yielding and subsequent plastic defor-
mation in glasses is a subject of ongoing study. A more general picture is
that applied stress converts the structure from one metastable state to
another, which persists after the stress is removed. Structural rearrangement
requires that interatomic or intermolecular bonding in the unstrained state
is overcome and replaced by roughly equivalent bonding in the deformed
state. Challenges to implementing this concept are manifold, primarily
because the structure of a glass is difficult to describe with either experimen-
tal or theoretical approaches. Argon (1993) further emphasizes that non-
uniformity of local structure is central to plasticity in glasses, the idea being
that some regions are less stable than others under macroscopic stress.
These 'critical' regions are not revealed, unfortunately, by average experi-
mental features such as mass density or radial distribution functions.
Unique to polymer glasses are the multiple ways by which the structure
may react to an applied force. Molecular response may be by (a) stretching
or (b) bending covalent bonds in the chain; (c) by rotation about bonds in
the chain backbone; or (d,e) by displacement of neighbouring chain seg-
ments. These deformation modes are sketched in Figure 4.3. Covalent bond
distortion (a,b) and internal rotation (c) dominate the elastic deformation of
highly oriented fibers and isotropic elastomers respectively. For glassy
polymers one expects intersegmental displacements (d,e) to be important.
The potential energy change on increasing the separation between units
in Figure 4.3d determines the elastic moduli E and G (Haward, 1973;
Struik, 1991). Local shear displacements are likely candidates for plastic
strain, which is generated at higher stress. The sketch in Figure 4.3e attempts
to illustrate the establishment of new intersegmental bonding after a shear
step.
158 Yield processes in glassy polymers

tr~

,,~
I&'"
0
"tr' W~
\
tr~) 0

(a) (b) (c)

~
\
~ ~~
(d)
'\ (e) ~
Figure 4.3 Deformation modes of a polymer in response to force with direction
indicated by arrow: (a) covalent bond stretch; (b) covalent bond angle distortion;
(c) intramolecular rotation with fixed bond angle; (d) segment-segment separation;
(e) segment-segment shear displacement. Dashed circles indicate initial atom
positions for intramolecular displacements (a)-(c).

This discussion is limited to distortional plasticity, i.e. flow processes in


glassy polymers occurring at essentially constant volume. The important
subject of crazing, or dilatational plasticity where volume of the deforming
body increases, is covered in Chapter 6. Polymer yielding has been written
about extensively; perhaps the best early summary is the comprehensive
chapter by Bowden (1973) in the first edition of this book. Other cover-
ages by Brown (1986), Matsuoka (1986, 1992), Argon (1993), Ward and
Hadley (1993) and Crist (1993) are recommended. The present treatment is
intended to be largely self-sufficient and emphasizes recent experimental and
Mechanical testing and definitions 159

analytical developments that contribute to our understanding of the yield


process in glassy polymers.

4.2 MECHANICAL TESTING AND DEFINITIONS


Glassy polymers considered here are initially isotropic; macroscopic struc-
ture and properties, including mechanical properties, are independent of
direction. Elastic behaviour is observed for small loads, and two indepen-
dent elastic constants serve to describe the relation between stress and
strain, both of which are second-rank tensors. Stress fields may be uniaxial
or multiaxial. We note at this point that yielding does not occur under
purely hydrostatic stress; loading geometries of lower symmetry are required
to achieve plastic deformation. In all experiments considered here the
deformation (strain) is imposed and the resulting force (stress) is measured
as a dependent variable.

4.2.1 UNIAXIAL TENSION


Uniaxial tensile deformation is employed most frequently for evaluating
mechanical properties. Specimens such as those shown in Figure 4.4a, with
circular or rectangular gauge sections of initial length 10 and cross-sectional
area A o, are clamped at the larger gripping portions and extended at a
constant rate. The stretching force F is recorded as a function of the
increasing length 1 of the gauge section. Within the uniform gauge section
the (true) uniaxial stress is

(J = FjA (4.1)

where A ~ Ao is the instantaneous cross-sectional area. During elastic


extension along y, the transverse (x, z) dimensions and A decrease. Results
are often presented in terms of engineering or nominal stress (In = FjAo,
which is based on the undeformed gauge area Ao. Tensile deformation in the
stretching (y) direction is expressed as the engineering or nominal strain:

(4.2)

An alternative measure of uniaxial deformation is the 'true' or logarithmic


strain:

(4.3)
160 Yield processes in glassy polymers

t t

t y

(a) (b)

t
~x z

(c) (d)

Figure 4.4 Specimens for the study of yield in polymers. Uniaxial tension (a); uniaxial
compression (b); plane strain compression (c); simple shear (d).

Here we have introduced the axial stretch ratio A = 1/10 , Either definition of
strain is satisfactory, although logarithmic strain is sometimes preferred for
describing large deformations, the subject of Chapter 5.
The initial ratio of stress to strain defines the Young's modulus:

E = 1l· mdO"- (4.4)


' .... 0 dB
Either nominal or true stress and strain may be used to evaluate E, as all
values converge at low strain. Measurement of transverse strain, for example
that in the x direction, provides a second elastic constant, the Poisson ratio:
-Bx
v= __ (4.5)
By

From E and v one can determine the shear modulus G and the bulk
Mechanical testing and definitions 161

modulus K:
E
G=---
2(1 + v)
(4.6)
K= E
3(1 - 2v)
Typical values for glassy polymers near room temperature are E", 3 GPa
(l GPa = 109 N/m2) and v'" 0.35, hence G '" 1.1 GPa and K", 3 GPa
(Gilmour, Trainor and Haward, 1979).
While elastic properties are of interest in themselves, and are also signifi-
cant in certain models of plastic deformation (section 4.7), this chapter is
concerned with yielding. An example of tensile behaviour of the polycarbon-
ate of bisphenol A (PC) is given in Figure 4.5, where the difference between
true stress (lower solid line) and nominal stress (lower dashed line) is
evident. True stress was calculated from load and undeformed area for a
Poisson ratio v", 0.4 (Imai and Brown, 1976). The tensile yield point is
defined by the maximum true stress O"y = 210 MPa and the corresponding
strain By = 0.011. One can be certain that unloading from the yield point
will result in an irreversible or plastic strain Bp > O. The yield point is
sometimes based on other criteria (Bowden, 1973; Ward and Hadley, 1993),
particularly when there is no maximum in the load-displacement record.

300
__ .. :a.
280
260 ,,;' ,," Compression

240 /
,
/
220 I
I

200
,,
I

0180 ,,
~ 160
~ 140
in 120
100
80
60

Figure 4.5 Stress-nominal strain curves for PC at T = -196°C in uniaxial tension


and compression; strain rate is en = ± 1.67 x 10- 4 /s. Nominal stress O"n is given by
the dashed lines, and true stress 0" by the solid lines. (Reproduced from Imai and
Brown, 1976, with permission.)
162 Yield processes in glassy polymers

Unless specified otherwise in this chapter, the yield point is the maximum
in the true stress-strain curve for uniaxial tension or other stress fields.
Two complications are frequently encountered in tensile deformation at
or near the yield point. First, the sample may craze or fracture in a
macroscopically brittle manner, as discussed in Chapter 6 and Chapter 7.
At low temperatures these responses may preclude the observation of
yielding. It is not generally appreciated that crazing under tension is
enhanced by the gases (e.g. N 2) used to establish low temperatures (Imai
and Brown, 1976). Second, in the absence of crazing and fracture virtually
all glassy polymers deform inhomogeneously beyond the tensile yield point.
The resulting strain localization (neck formation) sketched in Figure 4.6
complicates the measurement of material properties because the stress state
is difficult to define and the strain and strain rates vary appreciably over the
gauge section. A schematic nominal stress-nominal strain diagram is shown
in Figure 4.7. Neck formation coincides with the growth and coalescence of

t t t t
'~ ~, .....
-
I'
"

I
f
(al (bl (el (d)

Figure 4.6 Simplified schematic of neck formation under tension; more than two
shear bands are usually present. (a) Nucleation of the first shear band inclined at 45 0
to the loading direction. (b) Growth of the first shear zone and nucleation of the
second band. also at 45 0 to the load axis. (c) Growth of the second shear zone
creates a nearly symmetric local region of plastic deformation. the neck. (d)
Propagation of the stabilized neck. (Reproduced from Stokes and Bushko. 1995. with
permission.)
Mechanical testing and definitions 163

Figure 4.7 Schematic nominal stress-nominal strain (O"n-8n) record of a conventional


tensile test when the sample necks. Localized shear bands (Figure 4.6a,b) form just
before or at the load maximum. The neck evolves during the load drop, after which
the stable neck propagates along the gauge section. Stress then rises when the neck
grows into the wider gripping portions.

inclined shear zones as illustrated in Figure 4.6a -c; conversion from a single
macroscopic shear band (Figure 4.6a) to a collection of shear zones with
nearly axial symmetry (Figure 4.6c) is rapid and generally occurs at the peak
load or nominal yield stress. In conventional tensile experiments the
extrinsic yield point is defined by the maximum nominal stress and the
corresponding nominal strain. The neck stabilizes at the point of the shallow
load minimum in Figure 4.7, after which deformation proceeds at a
relatively constant force by propagation of the neck shoulders along the
gauge section and some creep-like response of material within the neck.
Stress later rises when the neck advances into the wider clamping regions.
The relation between tensile stress-strain behaviour and neck formation
is treated further in Chapter 5. It is clear, however, that information on local
sample dimensions and displacements are needed for evaluation of local
stress and strain. Here one must distinguish between microscopic and
macroscopic size scales. Some type of microscopy is required to follow
deformation within shear bands with one dimension as small as 1 J.lm (e.g.
Figure 4.17a below). Macroscopic strain inhomogeneities over dimensions
of 0.1 mm and larger can be accommodated as described here and in section
5.2. G'Sell and Jonas (1979) and G'Sell et al. (1992) employ mechanical or
optical sensing of deformation within a volume element having an effective
gauge length 10 < 0.5 mm to measure local (macroscopic) tensile stress
and strain, even while the neck is being formed. Strain rates are also
distinctly non-uniform at this stage, and plastic deformation behaviour is
164 Yield processes in glassy polymers

120.-----,------r-----,------r-----~

100

~
..
~

co

..
.~
:::
.....

150
oL-__~__==±===~=====c~~
o 0.2 0.4 0.6 0.8 1.0
Effective strain

Figure 4.8 Tensile true stress-logarithmic strain (0'-8 t ) behaviour of PC at constant


local strain rate et = 1 x 10- 4 S-1 at various temperatures. Yield stress and yield
strain increase as temperature is lowered. (Reproduced from G'Sell et al., 1992 with
permission.)

rate sensitive as discussed below. Local strain changes may be combined


with microprocessor control of the stretching apparatus to maintain a
constant logarithmic strain rate (dct/dt = at) in the small volume element
being monitored. This refinement, although implemented rarely, is required
to observe intrinsic flow behaviour under well defined tensile conditions at
and beyond the yield point.
An example is shown in Figure 4.8 for PC deformed in tension at various
temperatures with a constant local strain rate et = 1 x 10 - 4 S - 1. Effective
stress is the true local tensile stress, equation 4.1, with a small correction for
triaxiality that accompanies neck formation, and 'effective strain' is the local
logarithmic strain in equation 4.3. At any temperature below Tg = 145°C
glassy PC yields with intrinsic strain softening (do/dc t < 0), i.e. a drop in
true stress beyond the yield point. At sti11larger strains there is conspicuous
strain hardening (do/dc t > 0) which serves to stabilize the neck. It is shown
with the Considere construction in Chapter 5 that the load maximum in a
conventional uniaxial tensile test, performed with a constant nominal strain
rate en' provides accurate measurements of intrinsic yield stress (Jy and yield
strain cY ' provided the stress is corrected for uniform reduction in area A as
in Figure 4.5. Characterization of deformation beyond the yield point
requires special instrumentation to measure intrinsic response in the usual
case when strain is localized in a neck. Meaningful tensile flow behaviour
can be obtained only with constant local strain rate experiments, few of
which have been reported.
Mechanical testing and defi nitions 165

4.2.2 UNIAXIAL COMPRESSION


Crazing is circumvented and strain localization is minimized with uniaxial
compression in which a cylindrical, or sometimes prismatic, specimen is
loaded on the two flat basal surfaces, as sketched in Figure 4.4b. Stress and
strain are defined with equations 4.1-4.3. While both quantities are nega-
tive, results are usually presented as absolute values as in Figures 4.1 and
4.5. Crazing cannot result directly from the applied compressive stress, but
crazes are occasionally seen, oriented normal to the loading direction, at the
intersection of shear bands comparable to those sketched in Figure 4.6b
(Wu and Li, 1976). The same post-yield strain softening that enhances
necking in tension causes the local transverse area A to increase in
compression, lowering the local true stress (J = F 1A and stabilizing the
deformed region against additional strain. With careful attention to fric-
tional end effects, alignment and sample geometry it is possible to achieve
macroscopically uniform deformation far beyond the yield point. Strain rate
control is straightforward under this condition. Boyce, Arruda and Jaya-
chandran (1994) have reported uniform compression of cylindrical PC
specimens at room temperature to 1 '" 0.4/0 , or a compressive uniaxial strain
l:t = -0.9. Results in Figure 4.9 show strain softening after the yield point
followed by strain hardening, qualitatively similar to tensile response in
Figure 4.8. Note in Figure 4.5 that the magnitude of (Jy is larger in
compression than tension because of the pressure dependence of the yield
stress, a topic discussed in section 4.3.3.

4.2.3 PLANE STRAIN COMPRESSION


Another type of mechanical testing applies a compressive force along y while
constraining one transverse direction (z in Figure 4.4c); the sample is
compressed in the y-direction and erpands in the x-direction. Data are
presented as stress versus strain in the y-direction. As in uniaxial compres-
sion, deformation tends to be macroscopically uniform and crazing is
suppressed. The initial slope gives EI(1 - v2 ) '" 1.1E, and the yield point is
defined by the maximum (true) stress and corresponding strain. Figure
4.9 from Boyce, Arruda and Jayachandran (1994) illustrates that plane
strain yielding occurs at a greater stress than for uniaxial compression,
again because of pressure effects resulting from the dimensional constraint.
Bowden (1973) points out that plane strain compression is equivalent to
pure shear with a finite hydrostatic pressure for plastic flow; this correspon-
dence does not apply for elastic deformation before yield.

4.2.4 SIMPLE SHEAR


The sample in Figure 4.4d is gripped by the large sections and mutually
displaced in the ± y directions. A shear strain y = Aylx o is imposed on the
166 Yield processes in glassy polymers

200
d
c.. I
6 I
I
I

~ I
...
en'"
I
I
/
/
~
100 ,
/

~ /
,I
.....
I'
/" -- ....... , ( 4IJ ....
,I
j ............. .
uniaxial data

,
I plane strain data
I
;
o ~~~~~~~~~~~~~~~

0.0 0.5 1.0 1.5


True Strain
Figure 4.9 Plane strain compression (----) and uniaxial compression (.... ) of P( at
25°(, strain rate St = -1.0 x 10- 2 s. Plots are of true stress versus logarithmic strain
(0'-8 t on 0' - In A). Yield stress is higher and strain softening is less pronounced in
plane strain compression. (Reproduced from Boyce, Arruda and Jayachandran, 1994,
with permission.)

uniform gauge section of area Ao = YoZo and width x o, giving rise to a shear
stress L = F / Ao. Tests with this geometry have been extended to large strains
by G'Sell et al. (1983). Typical results for PC at different temperatures for a
shear rate dy/dt = Y = 3 X 10- 3 s-1 are illustrated in Figure 4.10 (G'Sell
and Gopez, 1985). Applied shear stress L and overall shear yare based on
gauge section dimensions as described above. Strain is uniform up to the
stress maximum that defines the yield point. There follows a stress drop
(strain softening) beyond yield, but intrinsic flow properties are distorted by
strain localization in a single narrow shear band that grows first in the
y-direction and then spreads in the x-direction. The magnitude of the stress
drop is correct, but local strain in the shear band is larger than the applied
strain y. This zone of high shear strain encompasses the gauge section and
deformation is once more uniform, e.g. for y > 1 at T = 23°C. Local strain
YI within the shear band can be measured to establish more accurate
post-yield flow behaviour (Grenet and G'Sell, 1990). With video monitoring
of local deformation, the experiment can be performed at a constant local
strain rate YI (Aboulfaraj et aI., 1994). Relatively few studies are done in
Mechanical testing and definitions 167

~ 80
o
a.
~ 70

I-' 60
en
~ 50
0:
l-
en
40
0:
<C
~ 30
en

...J
a.
a.
<C

0.5 1.0 1.5 20 2.5


OVERALL SHEAR) r
Figure 4.10 Simple shear stress-strain (-r-y) behaviour of PC at nominal strain rate
y = 3 X 10- 3 S-l at various temperatures. Yield stress and yield strain increase as
temperature is lowered. (Reproduced from G'Sell and Gopez, 1985, with per-
mission.)

simple shear, and the number with constant local strain rate, required for
meaningful post-yield flow behaviour, is smaller still. Elastic shear modulus
G is the ratio ,jy at small strains. The shear stress field has no hydrostatic
(pressure) component so deformation occurs at constant volume. Crazing
and brittle fracture are suppressed.

4.2.5 OTHER GEOMETRIES


Torsion-twist experiments with cylindrical tubes can be used to determine
the shear response of a glassy polymer (Wu and Turner, 1975), although it
is not feasible to evaluate stress, and strain y when deformation becomes
inhomogeneous beyond the yield point. Various states of biaxial stress may
be achieved by combining pressure within the cylinder and tension or
compression along the cylinder axis (Raghava, Caddell and Yeh, 1973). An
additional variation is hydrostatic pressure to establish a triaxial stress field
in combination with applied shear (Wu and Turner, 1975) or uniaxial
tension or compression (Spitzig and Richmond, 1979).
168 Yield processes in glassy polymers

In concluding this section it should be repeated that yield stress and yield
strain can be determined with acceptable accuracy from the usual plots of
nominal stress versus nominal strain, as illustrated in Figure 4.5. Further-
more, the presence or absence of post-yield strain softening or strain
hardening can be inferred from uniaxial compression, plane strain compres-
sion or simple shear recorded in engineering formats, although actual
softening or hardening rates may be distorted. Tensile deformation is a
different matter in the typical instance when a neck is formed. Here
engineering stress-strain curves give virtually no information on post-yield
flow characteristics, which can be established only with methods described
in sections 4.2.1 and 5.2.

4.3 YIELD PHENOMENA IN GLASSY POLYMERS


In this section we review how yield is affected by temperature, strain rate,
hydrostatic pressure and sample history. Emphasis is placed on the yield
point and flow behaviour immediately thereafter, which indicates the
presence or absence of strain softening. Where possible, the discussion is
based on true stress-strain curves obtained with well defined strain rates.

4.3.1 EFFECT OF TEMPERATURE

Temperature profoundly influences the yield stress, as exemplified by PC in


tension (Figure 4.8 from G'Sell et al., 1992) or simple shear (Figure 4.10
from G'Sell and Gopez, 1985). For any glassy polymer the critical reference
temperature is I'g. At temperatures at or above the glass transition (I'g =
145°C for PC), the material behaves like a viscoelastic liquid, deforming at
a small, nearly constant stress. A distinct yield point with strain softening is
seen when the glassy state is established at T = 135°C, and the yield stress
(Jy or 't"y more than doubles on cooling to room temperature. Yield strain By

likewise rises over the same interval. These trends are continued at much
lower temperatures, as seen in Figure 4.10 for shear and Figure 4.5 for
tension.
Shear behaviour in Figure 4.10 is qualitatively different near the glass
transition temperature. The initial slope (shear modulus G) at T = 142°C is
smaller and very temperature dependent, and yield occurs with an imper-
ceptible stress drop. These responses are more characteristic of viscoelastic
liquid deformation. Even in the 'glassy' regime at 135°C one puzzling feature
is the yield stress 't"y = 18 MPa, which is nearly equal to the tensile yield
stress (Jy = 20 MPa in Figure 4.8. At and below 120°C the ratio (Jy/'t"y '" 2, a
ratio more characteristic of solid deformation.
Yield phenomena in glassy polymers 169

The inverse relation between (Jy or Oy and temperature points to the


kinetic nature of yielding, which can be thought of as a type of reaction that
proceeds more readily at high T. However, readers should be advised that
much of the change of yield stress with T derives from the temperature
dependence of the elastic modulus, as discussed in section 4.5.2. At tempera-
tures more than c. 25°C below ~, yield in glassy polymers occurs with
intrinsic strain softening. The characteristic stress drop may not be seen for
deformation close to the glass transition range, suggesting a change from
solid to liquid-like behaviour.

4.3.2 EFFECT OF STRAIN RATE


A larger imposed strain rate S or y invariably results in a greater yield stress.
Viscoelastic heating may cause a temperature rise (Haward, 1994) and thus
reduce flow stress at very large rates, but results presented here are thought
to be free from this unwanted effect. A good example of rate dependence of
yield is provided by the work of G'Sell et al. (1992) for PC deformed in
tension at a series of constant (local) strain rates St. Effective stress and
strain in Figure 4.11 are the same as in Figure 4.8. Over the 20-fold
range of strain rate the yield stress rises from 69 MPa to 74 MPa and ey
changes from 0.04 to 0.06. The curves are displaced vertically by a constant
amount in the flow region, i.e. the shape of the stress-strain curve beyond
yield is unchanged. Here we are specifically concerned with yield and strain

120

100
~
:t:
80
:::
~
·zu~
~
::. 40

20

0
0 0.2 0.4 0.6 0.8 1.0
Effective strain
Figure 4.11 Strain rate dependence of tensile stress-strain behaviour (0"-6 t ) for PC
at 25°C. (Reproduced from G'Seil et aI., 1992, with permission.)
170 Yield processes in glassy polymers

60~---r----T---~----T----'----~

50
Etff (5-')
,.2_'0- 2
.
~ 30
,',.2-'0"
<...2_'0'4
",-2'10'5
'"
co
>
~ 20
~
UJ
10

°O~--~--~~--~----~--~----~
0.1 0.2 0.3 0.4 0.5 0.6
Effective strain

Figure 4.12 Strain rate dependence of tensile stress-strain behaviour (O'-e t ) for PC
at 125°C. Note the lack of strain softening at the lowest strain rate. (Reproduced
from G'Sell et al., 1992, with permission.)

softening; Hasan et ai. (1993) similarly find that strain softening is indepen-
dent of rate in uniaxial compression of poly(methyl methacrylate) (PMMA)
at room temperature.
Quite different effects may be seen as the temperature approaches ~.
Mention was made above of peculiar features of shear deformation for PC
near 145°C. At 125°C the rate dependence of O'y is much larger and strain
softening is variable, as shown in Figure 4.12. Most noticeable is the
reduction and eventual disappearance of the stress drop at low rates. While
O'y is altered substantially, flow stress at larger strains e ~ 0.2 is affected
much less, resembling low temperature behaviour in Figure 4.12. The same
effect is seen for PS deformed in plane strain compression at 80°C (Bowden
and Raha, 1970).
It is found that yield stress at one temperature has a weak power-law
dependence on strain rate, O'y oc sn' The coefficient n = d InO'y/d Ins is a
convenient way to summarize rate behaviour. Data in Figure 4.11 give
n = 0.035 for PC at room temperature; Matsuoka (1986) reports that
n '" 0.05 is typical of glassy polymers well below ~. It is apparent from
inspection of Figures 4.11 and 4.12 that n is temperature dependent,
particularly in the range of Tg• Rate sensitivity is also expressed as the nearly
equivalent ratio dO'y/d Ins (Figure 5.12); this quotient relates to the Eyring
model for rate dependence in section 4.5.2.
Yield phenomena in glassy polymers 171

4.3.3 EFFECT OF PRESSURE


Hydrostatic pressure P increases yield stress in a manner somewhat similar
to lower temperature or higher strain rate. The earliest experiments did not
always account for possible effects of the pressure-transmitting fluid on the
mechanical properties of the polymer (Pae and Bhateja, 1975). Among the
most reliable work on polymer glasses is that of Spitzig and Richmond
(1979) for Pc. Local dimensions were not measured in the pressure cell, so
results for uniaxial tension and compression at 25°C are presented in
engineering formats in Figure 4.13. The extrinsic yield point in tension
moves to higher (Jy and Ey with increasing pressure, and the load drop (not
a true stress drop) is less conspicuous at higher P. Changes in the

UO",,~~-r~'-'-~~~ 3110
210
340
200
190
azo
180 3DO
170 2110
180 :.:::E 2tIO
150
140 Ili
III
240

130 ...
a: 220
en
120 200
!E
110 Ii:
III
180
100 III
z 160
90 c;
z
III 140
80

70 120
80 100
50
80
40
60
30
20 40

10 20

o 0.1 0.2 0.3 0 .• 0.5 0.1 0.7 D.• 0.9 1.0 1.1 1.2 0 0.06 0.10 0.15 0.20 0.25 O.X
ENGINEERING STRAIN
ENGINEERING STRAIN

(a) (b)

Figure 4.13 Pressure dependence of (a) tensile and (b) uniaxial compressive defor-
mation of PC at 25°(, strain rate En = ±7 X 10- 3 s-1, presented as engineering
stress-engineering strain curves (O'n-en). Load or nominal stress drops are less
pronounced or absent at large pressures. (Reproduced from Spitzig and Richmond,
1979, with permission.)
172 Yield processes in glassy polymers

compressive response were even more pronounced. The increase in modulus


E with pressure is clearly seen with the expanded strain scale. For
P ~ 276 MPa there is no load drop in compression, although slope changes
indicate that yield occurs at a stress that is larger at high P. Yield stress
evaluated from the 1% offset strain was found to be the same linear function
of P for both tension and compression; linearity is also obtained for extrinsic
a y in tension from Figure 4.13a. As with temperature, a sizeable fraction of
the change is related to elastic modulus. Tensile yield stress a y is increased
by a factor of 3.5, and Young's modulus E increases by a factor of 2.3 over
the pressure range from 0.1 MPa (1 atm) to 1104 MPa.
Spitzig and Richmond (1979) calculated true stress in an approximate
manner and report no true stress maximum for either tension or compres-
sion when P ~ 552 MPa. Hence raising the hydrostatic pressure from
0.1 MPa (1 atm) to c. 103 MPa renders the yield process, which occurs at
greater uniaxial stress magnitude, more gradual in both tension and
compression. Less abrupt yielding at high pressures is exhibited by a
number of other glassy polymers described by Pae and Bhateja (1975).

4.3.4 CRITERIA FOR YIELDING


The stress state on the solid varies for the mechanical testing schemes
described in section 4.2. One notes for instance in Figure 4.5 that the yield
stress ay is greater in uniaxial compression than in uniaxial tension.
Conditions for shear yielding in glassy polymers (crazing is not considered
here) are best summarized by the pressure-modified von Mises criterion:

Here a l' a 2 and a 3 are the principal (normal) stresses, T~ is the yield stress
in pure shear and under zero pressure, P = - (a 1 + a 2 + a 3)/3 is the
hydrostatic pressure (negative of the mean stress anJ, and Ji is the dimen-
sionless pressure coefficient. The quantity T~ is not a material constant, but
depends on temperature, strain rate and, as shown below, thermomechanical
history of the glass. Equation 4.7 corresponds to yield occurring when the
elastic shear strain energy in the stressed body achieves a critical value,
which depends on pressure P. At higher P this energy is larger and yield
occurs at a greater shear stress Ty = T~ + JiP.
It is useful to evaluate equation 4.7 for uniaxial stress where a 1 = a y,
a 2 = a 3 = 0 and P = - ay/3, which gives the yield stress magnitude:

(4.8)
Yield phenomena in glassy polymers 173

- 2.0

-1.0

-2.0

Figure 4.14 Locus of biaxial and uniaxial yield stresses for PVC (., x), PC (_), PS
(T) and PMMA ( +). Yield stresses have been normalized by tensile uy at the same
temperature and strain rate. Line is calculated for the representative pressure
coefficient J1 = 0.23. (After Raghava, Caddell and Yeh, 1973, with permission.)

Here the plus sign is for tension and the minus sign for compression.
Tension/compression experiments in Figure 4.5 give t~ = 133 MPa and
Jl. = 0.15 for PC at -196°C.
The yield surface represented by equation 4.7 is a tapered cylinder centred
on the hydrostatic line u 1 = (J 2 = U 3; circular sections normal to this axis
are smaller for greater mean stress Urn = (u 1 + u 2 + u 3)/3. This surface is
intersected by the (J 1(J 2 plane along an elliptical line corresponding to yield
at constant (J 3; data are often plotted on the biaxial (J 1 (J 2 plane for (J 3 = 0
(neglecting atmospheric pressure). For instance, (Jy for uniaxial tension and
compression are along + (J 1«(J 2) and - (J 1( - (J 2) respectively, and shear
yielding is for (J 1 = - (J 2' Figure 4.14 presents data for a number of glassy
polymers normalized by the tensile yield stress measured for the same
temperature and strain rate. Experimental points measured or compiled by
Raghava, Caddell and Yeh (1973) conform well to equation 4.7 with
Jl. = 0.23. Similar values of Jl. = 0.10-0.25 were reported by Bowden and
Jukes (1972). The tension/compression experiments of Bauwens-Crowet,
Bauwens and Homes (1972) on PC indicate that Jl. is independent of strain
rate, but may increase at low temperatures.
The von Mises yield criterion is additionally used to convert between
uniaxial stress (J and shear stress t. The von Mises equivalent uniaxial
174 Yield processes in glassy polymers

stress is
(4.9)
The equivalent linear (tensile or compressive) strain is given by
Seq = y/3 1 / 2 (4.10)
Since the shear yield stress "C y = "C~ + JIP in equation 4.7 is a linear function
of pressure, the same is true of the uniaxial yield stress (Jy defined with
equation 4.9. The vast majority of experiments show that tensile and
compressive yield stresses increase linearly with P (Raghava, Caddell and
Yeh, 1973; Spitzig and Richmond, 1979).

4.3.5 CHEMICAL MICROSTRUCTURE


Results thus far have been for glassy thermoplastics that furthermore are
homopolymers. Aside from possible time effects during physical aging
(section 5.3.2), there is no evidence that yield is sensitive to molecular
weight, provided the glass is composed of well-entangled high molecu-
lar weight chains and the glass transition temperature has reached its
asymptotic limit. Data for random and block copolymers are meager.
Should copolymerization change Tg appreciably, one expects yield at fixed
temperature and strain rate to reflect proximity to the glass transition.
Gloaguen, Escaig and Lefebvre (1995) considered a set of random
copolymers of styrene and methyl methacrylate for which ~ is fairly
insensitive to chemical composition. Compressive yield characteristics at
temperatures between -100°C to + 70°C varied regularly between limits set
by the respective homopolymers. Some aspects of yield and deformation of
block copolymers and two-phase blends have been reviewed by Crist (1993).
Crosslinked network polymers are exemplifed by epoxies. Early work by
Lohse et al. (1969) showed that increased crosslink density causes more
pronounced strain hardening in tensile experiments c. 25°C below ~. This
observation is consistent with models for large strain behaviour treated in
Chapter 5. Of more importance here are yield and immediate post-yield
flow, for which reports are mixed, perhaps because cure conditions affect the
results. Yamini and Young (1980) varied the amount of alkyl amine curing
agent by 100% and evaluated glassy epoxies under uniaxial compression
between - 60°C and ~ '" 100°C. While ~ increased as expected with
crosslinking, both room temperature Young's modulus E and compressive
yield stress (Jy were reduced by c.30%. Stress-strain curves were not
reported, but compressive strain was said to be homogeneous (no shear
bands) with strain softening past the yield point. Yield stress had conven-
tional dependencies on strain rate and temperature.
Yield phenomena in glassy polymers 175

Different results were obtained by Oleinik (1986) for tensile properties of


epoxies in which the amount of amine crosslinking agent was varied, again
by a factor of two. In this case Young's modulus E increases slightly (c. 10%)
with amine content, and the yield stress at 25°C is constant. Kozey and
Kumar (1994) find that increased crosslinking, accomplished with cure
temperature or time, eventually leads to compressive yield at room tempera-
ture with no strain softening. It is possible that highly crosslinked epoxies
experience strain hardening at stretches so small that strain softening is
suppressed.
Regardless of subtle effects that may arise from different crosslinking
conditions, yielding in network glasses is generally similar to that in glasses
composed of linear chains (Yamini and Young, 1980; Oleinik, 1986). The
elastic modulus, yield stress and strain softening rate under different stress
fields are comparable to values in uncrosslinked materials. Two special
characteristics, increased strain hardening at large plastic strains (Lohse et
al., 1969) and the lack of crazing in tension (Oleinik, 1986) are direct
consequences of network structure.

4.3.6 THERMAL AND MECHANICAL HISTORY


Polymer glasses are not thermodynamic substances with a unique relation
between pressure, temperature and volume. Physical aging is the term
applied to the slow approach of state functions toward equilibrium liquid
values below the laboratory glass transition temperature Tg• These effects, as
well as elastic and viscoelastic properties measured at small strains, are
discussed in Chapter 3. Physical aging has a number of consequences on
yielding as well, the most obvious being the increase in yield stress
illustrated in Figure 4.1. For PS tested in uniaxial compression, aging while
the 'annealed' glass has been cooled slowly from above ~ increases the
magnitude of (Jy by 20 MPa (nearly 30%), compared to a glass quenched
from the liquid state. Two other features are apparent in Figure 4.1; strain
softening beyond the yield point is enhanced, and flow curves for quenched
and aged samples merge at large strains. Additional discussion of these
effects may be found in Chapter 5.
Another way to modify substantially the yield properties of a polymer
glass is to subject it to previous deformation cycles. Quite spectacular
effects may be observed when mechanical conditioning involves strains well
beyond the yield point, as in Figure 4.15 for simple shear of PC at room
temperature (G'Sell et al., 1989). The first deformation along path (a) is
conventional, with growth of a localized shear band and strain softening
beyond the yield point. Reversed loading along the path (b) reverses
all plastic and elastic strains, the sample returning to its undeformed
176 Yield processes in glassy polymers

50
~ 40
:::E
V! 30
V!
w
II: 20
l-
V!
10
II:
« 0
w
:J:
V!
-10
0
w -20
:J
11.
« -30
11.

-4~.25 0 .25 .5 .75 1.0 1.25


APPLIED SHEAR

Figure 4.15 Large strain cycling (a,b) substantially reduces yield stress T:y and
eliminates strain softening (c) for PC deformed in simple shear at room temperature.
See text for explanations. (Reproduced from G'Sel1 et al., 1989, with permission.)

dimensions. Note that the reverse deformation occurs at an appreciably


smaller stress magnitude. This strong Bauschinger effect (Bowden, 1973) is
associated with entropic back stresses generated at large plastic strains, as
discussed in Chapter 5. It is important to note that the mechanically cycled
glass at zero strain and stress is isotropic by birefringence and X-ray
diffraction. However, the deformation cycle has created 'strain energy' that
can be observed by differential scanning calorimetry as discussed in section
4.4.3. The second loading of PC is along the singularly different path (c):
yield stress is reduced by more than 50% and there is no evidence whatever
of strain softening. Deformation is macroscopically uniform in the gauge
section, and curves from the first and second loadings merge at high strains.
Network glasses show the same effect; yield stress in a physically aged epoxy
is reduced from 't y = 30 MPa to 10 MPa by a single shear cycle which also
eliminates strain softening (Aboulfaraj et ai., 1994). It has long been known
that this sort of mechanical softening or 'rejuvenation' occurs when condi-
tioning deformation is extended past the yield point (Bowden, 1973); steady
state flow must be established to modify the nature of the glass.

4.4 RELATED STUDIES OF YIELDING

4.4.1 STRAIN UNIFORMITY


Plastic deformation invariably starts at some point of stress concentration
or perhaps in a region where local structure is more susceptible to yielding.
Related studies of yielding 177

The degree of strain non-uniformity varies from polymer to polymer, and


further can be modified by temperature, strain rate and history. It was
shown in section 4.3 that polymer glasses exhibit strain softening under most
conditions. Hence it is likely that strain is localized in those parts of the
sample where yield first occurs. One consequence is exacerbated necking in
tension (although this plastic instability may occur in the absence of
intrinsic strain softening; see Chapter 5), kinking in uniaxial compression or
banding in simple shear that complicates interpretation of load displace-
ment records. More importantly, the intrinsic response of the material
during yield and post-yield flow may be distorted, thereby muddling our
understanding of yield processes. Consider a macroscopically homogeneous
deformation as in the compression tests of PS in Figure 4.1, where the yield
strain is cy ~ - 0.06. Does this signify that all of the material has a strain of
0.06 and subsequently flows (homogeneously) with decreasing stress? Or is
the macroscopic strain achieved by much larger local deformations in for
instance 10% of the material, and the observed stress drop derives from
subsequent conversion of un yielded regions to a state of high plastic strain?
The distinction between these possibilities cannot be made from mechanical
measurements.
We first consider PMMA as representative of those glassy polymers with
relatively homogeneous strains. PMMA that had been deformed slightly
past the yield point at 22°C in plane strain compression (just beyond the
stress maximum in a plot similar to Figure 4.9) was unloaded, sectioned,
polished and viewed in a polarizing microscope (Bowden, 1973) is shown in
Figure 4.16. Plastic strain is evident in the light birefringent shear zones,
about 0.5 mm thick, inclined at 45° to the compression axis, the direction of
maximum resolved shear stress. Bowden and Raha (1970) showed that these
zones form just before the yield point, then widen as the true stress rises to
(Jy and subsequently falls, with the birefringence pattern (local strain)

becoming uniform at the true stress minimum. An important observation


(Bowden, 1973) is that the plastic strain within the diffuse shear zones is only
a few percent above that in the 'matrix', implying that strain is nearly
uniform during yield and subsequent flow. Matsushige, Radcliffe and Baer
(1976) similarly found that tensile yield in PMMA correlates with the
growth of diffuse shear zones across a small part of the gauge section.
Very different behaviour is seen in slowly cooled polystyrene (PS) after
compressing past the yield strain cy = - 0.06. Other conditions were the
same as for PMMA, but plastic strain in PS is seen in Figure 4.17a to be
localized in dozens of well defined shear bands (not diffuse shear zones) of
thickness ~ 1 J.lm. Analysis by Bowden and Raha (1970) showed that these
bands first nucleate at a strain I: = - 0.04. They increase in number and in
length, but not width, through the yield point. Plastic strain within the shear
178 Yield processes in glassy polymers

Figure 4.16 Diffuse shear zones in PMMA deformed at room temperature just past
the plane strain compression yield point and unloaded. (Reproduced from Bowden,
1973, with permission.)

bands is Yl '" 2 and that in other regions is essentially zero; yield in this case
is extremely inhomogeneous at the microscopic level. Quenched PS, on the
other hand, deforms at room temperature with diffuse shear zones that more
closely resemble those in PMMA (Figure 4.17b).
A definitive study has also been carried out on simple shear deformation
of PC at room temperature (G'Sell and Gopez, 1985). One well defined
shear band of thickness '" 0.1 mm, much larger than the features seen in PS,
is nucleated at the yield point. Local shear strain within the band is uniform
at Yl '" 0.6, and remains nearly constant as the band grows first in length
and then in width as described in section 4.2.4.
The situation with other polymers is uncertain. Bowden (1973) reports
that local strain Yl within shear bands in glasses deformed in plane strain
compression at room temperature increases in the order PVC, epoxy and
poly(ethylene terephthalate) (PET), although no values are given. Yamini
and Young (1980) note the absence of shear bands in epoxies subjected to
uniaxial compression, while Oleinik (1986) observes diffuse shear zones in
tensile deformation of similar network glasses. Kozey and Kumar (1994)
find shear bands c. 0.1 mm thick in epoxies that yield in compression with
a stress drop, but see no such features in more highly crosslinked networks
that deform without strain softening.
There is a tendency for yielding to proceed homogeneously (diffuse shear
zones) at high temperatures and low strain rates, while shear banding is
favoured by low temperatures and high strain rates. Bowden and Raha
(1970) found this transition occurs near 70°C for slowly cooled PS at strain
Related studies of yielding 179

Figure 4.17 (a) Intense shear bands in slowly cooled PS deformed at room
temperature just past the plane strain yield point and unloaded. (Reproduced from
Bowden, 1973, with permission of John Wiley & Sons.) (b) Diffuse shear bands in
quenched PS deformed in the same manner. (Reproduced from Haward, 1974, with
permission.)

rates ,...,1O - 4 s - 1 . Wu and Li (1976) confirmed the switch from shear zones
to shear bands with increasing strain rate in notched PS under uniaxial
compression. Similar but less well documented transitions have been found
in PMMA (Bowden and Raha, 1970) and PC (Walker, 1980). Returning to
PS, Wu and Li (1976) observed with transmission electron microscopy that
diffuse shear zones are not homogeneous, but are composed of a dense
network of very thin ( ,..., 0.1 11m) slip bands. No estimate of local shear strain
within the fine slip bands was given, and we are unaware of other reports
of this type of structure.
Strain localization is obviously enhanced by intrinsic strain softening rate,
i.e. the slope magnitude -do/ de beyond the yield point in stress- strain plots
180 Yield processes in glassy polymers

when the deformation is macroscopically homogeneous. A more subtle effect


derives from the positive rate dependence of the flow stress beyond the yield
point, which stabilizes a locally yielded element against further deformation.
Bowden and Raha (1970) concluded that the ratio of softening rate to rate
dependence (-da/de)/(da/dlne) was indicative of shear band formation. A
large ratio (pronounced strain softening and/or small rate dependence) was
found to correlate with intense shear bands, and a small ratio with diffuse
shear zones. Argon and Bessonov (1977a) presented a more quantitative
treatment that predicts strain magnitudes in both shear bands and the
surrounding matrix. Such appraisals, however, require knowledge of intrin-
sic properties that cannot be measured directly when deformation is·
inhomogeneous, i.e. dominated by shear bands. Prediction or analysis of
intense shear bands must be extrapolated from relatively homogeneous
conditions when strong strain localization is absent.
The macroscopic strain softening rate that accompanies strain localiz-
ation is affected by nucleation and growth of shear bands as well as intrinsic
flow characteristics. Hence a large softening rate -da/de does not necessar-
ily correlate with shear banding during yield. Indeed, Bowden and Raha
(1970) observe about ten times more strain softening in PS deformed at
75°C (homogeneous, diffuse shear zones) than at 21°C (heterogeneous,
intense shear bands). An obvious corollary is that it is incorrect to interpret
macroscopic stress-strain behaviour in terms of a homogeneous mechanism
when shear bands are important.
Thus yielding in uniaxial tension, uniaxial compression or plane strain
compression develops by shear on planes '" 45° to the principal stress
direction. This deformation is relatively homogeneous throughout the
stressed material when yielding occurs with diffuse shear zones, typically at
high temperatures and low strain rates or with quenched glasses. Low
temperature, high strain rate and physical aging favour yielding in which
plastic deformation is localized in thin shear bands where strains are large
(YI> 1). Strain gradients associated with shear bands cause the macroscopic
true stress-true strain curves to distort the intrinsic flow properties of the
polymer. In particular, the strain softening rate -da/de beyond the yield
point will reflect nucleation and growth of shear bands in addition to local
relations between stress and strain. G'Sell· (1986) therefore argues that
simple shear is the method of choice for evaluating intrinsic stress-strain
behaviour, because the stress field generates only one well defined region of
strain localization that can be monitored.

4.4.2 MACROSCOPIC AND FREE VOLUME


It is well established that plastic deformation of polymer glasses occurs at
nearly constant volume, but with a small contraction ~ V /Vo '" - 2 x 10 - 3
Related studies of yielding 181

0
-0.1 o LOADING
o UNLOADING
-0.2
-0.3

-0.•
/If.
-0.5
i>
<3 -0.8

-0.7
-0.•
-0.9
-1.0
-1.1
0 10 20 30 40 50 eo 70 eo
TRUE STRESS, Mh

Figure 4.18 Volume change as a function of uniaxial compressive stress for PC at


room temperature; circles are for applied compressive strain, squares for strain
reversal or unloading. There is an abupt densification of c 0.1 % during yield and
strain softening at the highest stresses. (Reproduced from Spitzig and Richmond,
1979, with permission.)

after yield (Brady and Yeh, 1971; Bowden, 1973). Here we present data for
PC that are thought to be representative. Powers and Caddell (1972)
followed volume changes by measuring local axial and transverse logarith-
mic strains during tensile elongation at 25°C, strain ratei::, = 7.6 x 10- 6 S - 1.
Neck formation corresponding to extrinsic yield occurred at ey = 0.047
under these conditions. The volume increased to a maximum (~0.4%) at
et = 0.03 and then decreased by 0.1 % before the yield point. Spitzig and
Richmond (1979) observed volume changes by dilatometry during uniaxial
compression at a strain rate of -7 x 1O- 4 s- 1 . Data in Figure 4.18 show
elastic densification up to the yield stress G y = - 80 MPa, after which there
is an abrupt volume decrease of about 0.1 % during strain softening that is
preserved during and after unloading. Note that this volume change with
compressive yield is of the same magnitude as that reported in tension.
There is a strong historic association between yield and free volume,
which is amplified in sections 4.6 and 4.8. In recent years positron annihila-
tion lifetime spectroscopy (PALS) has been used to probe free volume in
liquid, glassy and semicrystalline polymers. The technique relies on rela-
tively slow 'pick-off' decay of o-positronium, an electron-positron particle,
in 'holes' with dimensions of a nanometer or less. Experimental quantities
182 Yield processes in glassy polymers

are the decay time r 3 and the normalized intensity 13 of o-positronium. The
usual analysis has r 3 indicate the effective hole radius and 13 the number
density of such holes. There are three reports of PALS studies of plastic
deformation of glassy polymers, two on PC under load with increasing step
strains, while the investigation of PMMA was done (unloaded) after
different plastic strains. Ruan et al. (1992) find for PC that the number of
free volume sites (I3) is unchanged by tensile strain, while average hole size
(r3) increases with strain and becomes constant near the yield point. Similar
results were obtained by Xie et al. (1995), although a subsequent decrease
in total free volume was seen beyond By. Uniaxial compression caused both
the number (I3) and the size (r 3) to decrease, leading to less free volume
under compressive load. In all cases the relative changes in free volume were
about five times larger than relative changes in macroscopic volume.
PMMA was investigated by PALS before and after uniaxial compression
(Hasan et al., 1993). In the unloaded state the number of free volume sites
(I3) is unchanged by total applied strain from Bt = 0 to Bt = - 0.4, although
size (r 3)' and hence total free volume, increases. This finding is at odds with
the decrease in free volume reported for PC under compressive stress. It is
unclear if this difference reflects strain amplitudes, stress state, specific
polymer effects or some unknown aspects of PALS. Hasan et al. (1993)
further find that free volume (r3) in unstrained PMMA is reduced by
physical aging, but is independent of thermal history after deformation well
past the yield point. These effects are associated with the increase of (jy and
the independence of large strain flow stress after aging. Such behaviour is
illustrated in Figure 4.1 and is developed in a constitutive model in section
4.6.3.

4.4.3 CALORIMETRY
Mechanical work is done on the solid body during plastic deformation,
some being released as heat, while the remainder is stored as internal energy.
The change in internal energy AU can be measured more or less directly by
calorimetry during deformation, or inferred from differential scanning calor-
imetry (DSC) on samples before and after the imposition of plastic strain.
Oleynik (1989, 1991) and his colleagues have been the most active in the
study of deformation calorimetry on polymer glasses. Their experiments
were done under uniaxial compression with simultaneous recording of heat
flow Q from the sample and nominal stress and strain. A specimen is
compressed to a predetermined total strain Bdef and unloaded at the same
strain rate Bn. Irreversible work done on the material, i.e. the area enclosed
by one loading-unloading cycle in the inset of Figure 4.19, is designated by
Adef and liberated heat, measured by the calorimeter throughout one cycle,
Related studies of yielding 183

~
..-..
Q>
3
::::>
10 <]
'Qj
"C

0
'Qj
"C
5 ~

30

Figure 4.19 Deformation calorimetry results for PS during uniaxial compression at


30°C, En = -10- 3 S 1 (see inset). Curve 1 is mechanical work A de! in loading-
unloading cycles. Curve 2 is the heat Ode! liberated during the cycles. Curve 3 is flU
from equation 4.11, which becomes constant at large strains. (Reproduced from
Oleynik, 1989, with permission.)

is Qdef' The change in internal energy associated with deformation to a


specific strain Bdcf is
(4.11)
The process is repeated for a different 8 def , etc.; results for PS are shown in
Figure 4.19. The most significant feature is the rise of internal energy flU to
a nearly constant value at strain magnitude greater than 0.25. One also
notes that the inflection point for flU versus Edef is near the yield strain of
0.12. The same sort of increase in flU to a plateau, the height of which
depends on the specific polymer, was observed in epoxies, PMMA, PVC and
amorphous PET. Similar behaviour of PC is shown in Figure 5.19.
Related effects are seen by DSC on glassy polymers that have been
plastically deformed and unloaded. In such experiments it is feasible to vary
temperature, strain rate, stress state, etc. far beyond limits imposed by
deformation calorimetry. Oleynik (1989, 1991) has reported work of this
sort, but Figure 4.20 is from Hasan and Boyce (1993) who define better the
relation between heat effects and mechanical behaviour. We present the
example of quenched PS, examined after uniaxial compression at room
temperature, for which the yield strain is By = -0.06 (Figure 4.1). A
decrement in heat capacity is present between 70°C and the calorimetric
~ = 105 C when deformation approaches or exceeds the yield point.
C
184 Yield processes in glassy polymers

0.20
Quenched

0.15

~
~
_ 0.10
.~

0.05

50 100 150
Temperature (·C)

Figure 4.20 DSe of quenched PS after uniaxial compression at room temperature


to different total (logarithmic) strains; see Figure 4.1. Deformation energy is liberated
between about 70 0 e and ~. (Reproduced from Hasan and Boyce, 1993, with
permission.)

This is interpreted as the release of enthalpic energy created by yielding and


post-yield flow. The liberated 'strain energy' I1H is given by the area between
scans of deformed and undeformed samples; I1H = 2.0-5.5 J g -1 parallels
I1U (Figure 4.19) in both strain dependence and size. Additional DSC
experiments emphasizing high strain behaviour are discussed in section 5.4.

4.4.4 OTHER STUDIES OF DEFORMAnON


One obvious mechanism for straining a polymer glass involves stretching or
uncoiling the chain-like molecules, analogous to deformation of elastomeric
networks above ~. The stress rise at large strains (Figures 4.8-4.13) is nicely
associated with this process, as discussed in Chapter 5. Quantitative infor-
mation on chain orientation in deformed glasses is not easy to obtain. One
method uses IR spectroscopy to monitor the relative population of back-
bone bonds in gauche and trans rotational states. Magonov, Vainilovitch
and Sheiko (1991) thus concluded that room temperature straining up to
the tensile yield point of PS and amorphous PET involves no changes in
rotational states. Post-yield strain, on the other hand, does increase the
number of extended trans conformers. A similar study of PS at ~ = 105°C
provided very interesting results (Theodorou, Jasse and Monnerie, 1985).
High temperature tensile yield and flow occurred without a load drop,
The nature of yielding in glassy polymers 185

presumably without necking. In this case the number of extended trans bond
sequences decreased abruptly by c. 10% at the yield point then increased
with post-yield flow. The transient increase in high energy gauche con-
formers at yield bears directly on the Robertson model discussed in section
4.7.1.
Small angle neutron scattering (SANS) can measure the distortion of
initially isotropic polymer chains, and thus provides information on defor-
mation at the molecular level. Lefebvre et al. (1985) examined PS and
PMMA after uniaxial compression of notched samples, where plastic strain
is concentrated in a macroscopic zone having y ~ 1. Results were seen to
parallel strain uniformity (section 4.4.1). When PS yields with well defined
microscopic shear bands, in this case at T < 34°C, coil strain was nearly
equal to macroscopic strain. For PS at higher temperature and PMMA,
where yield occurs with diffuse shear zones, the polymer chains remained
essentially isotropic. This interesting correlation between strain localization
and molecular response appears to break down in tensile deformation.
Dettenmaier et al. (1986) similarly examined PMMA after uniaxial drawing
to plastic strain Ep ~ 1 at temperatures between 60°C and ]~ = 125°C. In all
cases the chain deformation was c. 90% of the macroscopic strain, although
tensile yield is with diffuse shear zones (Matsushige, Radcliffe and Baer,
1976). It should be emphasized that SANS studies were on samples strained
well past the yield point; this method is unlikely to provide any information
on chain deformation near Dy •
An indirect technique relies on recovery of original dimensions (thermal
reversion) while heating to and above ~ at a constant rate (Oleynik, 1989,
1991). Strain recovery first occurs in the temperature range below ~ where
release of stored deformation energy causes the heat effect in DSC (Figure
4.20). Only for samples compressed well past yield and having plastic strains
Ep ~ 0.1 was a second, high temperature recovery seen at T ~. ~. From that
work it was concluded that the first increment of plastic strain, that
associated with strain softening, involves unspecified local displacements
that increase the internal energy. This process saturates and further strain
is achieved by uncoiling the polymer chains. The latter mechanism leads to
strain recovery only after heating above ~.

4.5 THE NATURE OF YIELDING IN GLASSY POLYMERS


Attempts to understand the fundamental manner by which glassy polymers
yield fall into two categories, one based on viscous flow and the other
borrowing heavily from plasticity in metals. The earliest models proposed
that yield occurs by liquid-like flow when adiabatic heating raises the local
temperature to ~, or when pre-yield (tensile) strain increases the free
186 Yield processes in glassy polymers

volume sufficiently to shift ~ to the deformation temperature. Shortcomings


of these concepts have been reviewed many times (Bowden, 1973; Struik,
1991; Oleynik, 1991; Crist, 1993). Suffice it to restate that, while local heat-
ing or volume (free volume) changes may facilitate yield and flow under
certain conditions, glassy polymers exhibit considerable ductility when
there is no increase in thermal energy or macroscopic volume. Figure 4.1,
showing compressive yield of PS at room temperature, provides an excellent
example. Equally convincing is the compressive yield of PC at T = -196°C
in Figure 4.5.
Because glasses are effectively supercooled liquids, it is natural to think of
deformation mechanisms in terms of those molecular processes - slip and
uncoiling of chains - that occur above ~. Aside from the question of how
such large scale diffusive motion might be activated, liquid-like deformation
is generally inconsistent with intrinsic strain softening and resultant strain
localization. Indeed, microscopic shear bands (Figure 4.l7a) are suggestive
of similar features in metals where plastic strain is controlled by mobile
dislocations. If one accepts the possibility that yielding in polymer glasses is
governed by molecular scale plastic events that are spatially correlated, as
with a dislocation mechanism, the growth or multiplication of dislocations
is a natural way to account for strain softening. This concept is amplified in
section 4.7.3.

4.5.1 THE ULTIMATE SHEAR STRENGTH


An estimate of the stress required to effect yield in a polymer glass, in the
absence of thermal activation of flow or any specific defect mechanism, can
be based on local shear displacements similar to those sketched in Figure
4.2. For defect-free crystalline solid the stress required to cause planes of
atoms to slide past one another is r* '" G/15 (Kelly and Macmillen, 1986).
This would correspond to macroscopically homogeneous yield in simple
shear at T = 0 K, where the shear modulus G is also measured at 0 K.
Indeed, Buchdahl (1958) noted nearly four decades ago that the ratio of
strength to modulus in polymer glasses at conventional temperatures is
about 0.04, which certainly approaches r* /G '" 1/15 = 0.067.
Glassy polymers are composed of entangled, covalent chains with static
Gaussian conformations (Wignall, 1987). The response of such a system to
stress is usually confined to changes in internal rotation angles (Figure 4.3c)
and intermolecular positions (Figure 4.3d,e), with covalent bonding remain-
ing fixed. Brown (1983) considered various ways a chain may rearrange
under a shear stress to establish permanent or plastic strain in a glass. He
concluded that intersegmental shear, similar to the displacement in Figure
4.2, is the dominant molecular response, and calculated the ultimate shear
The nature of yielding in glassy polymers 187

strength ,* '" GI13 for segments interacting by a Lennard-Jones potential.


Argon (1973) took the alternate view that molecular stram results from
coordinated rotation about pairs of C-C backbone bonds. In the glassy
state this conformational change is opposed by intermolecular forces, as
with the sliding displacements favoured by Brown. The kink pair model is
developed more fully in section 4.7.2; for now, we use the result ,* '"
G/8.
These two estimates of theoretical yield stress bracket the likely motions of
chains in a glass, and give similar values of ,*.
The important conclusion is that shear yielding in glassy polymers occurs
by rearrangements that overcome, and then re-establish, intermolecular
bonding that is reflected in the magnitude of G. This idea IS supported by
data for PC in Figure 4.21, where the normalized shear strength 'y/G
is seen to approach 0.065 at T = 0 K. Other glasses have similar yield
properties; the average for PC, PMMA, PS and PET is Ty = Gill when
extrapolated to T = 0 K (Brown, 1983), in reasonable agreement with
theoretical estimates of ,*
outlined in the preceding paragraph. The solid
line in Figure 4.21 is an empirical fit to the data, from which the open point
at T = 416 K (143°C) is excluded. Yield at temperatures close to ~ acquires
characteristics of liquid deformation, indicated by the abrupt drop in 'y/G.
Rere we are concerned with solid behaviour at lower temperatures.
It has been established that yield in glassy polymers is opposed by inter-
molecular interactions represented by the shear modulus G. Experimental

0.08

-
0.06
(!)
••
>-
l-'
0.04


0.02

0.00
0 100 200 300 400
T (K)
Figure 4.21 Normalized yield stress ryfG as a function of temperature for Pc. Circles
are from r/T) from Figure 4.10 and G(T) from G'Sell (1986). The low temperature
square is from Figure 4.5 and equation 4.8. The line is an empirical fit to all but the
highest temperature point.
188 Yield processes in glassy polymers

yield stress 't y is reasonably close, typically within a factor of two, to the
calculated ultimate strength 't* '" 0.1 G. It should be appreciated that
polymer glasses are extraordinarily strong on a relative basis. A high
strength steel with (1y = 400 MPa has 'ty/G = 3 x 10- 3 , easily an order of
magnitude below the reduced yield strength for PC in Figure 4.21. Absolute
strength can be impressive as well; PMMA yields in uniaxial compression
at (1y = -410 MPa at T = -120°C (Bowden, 1973; Haussy et ai., 1980).
Temperature and strain rate have moderate but important effects on
reduced strength 'ty/G that are considered next.

4.5.2 THERMAL ACTIVArION OF YIELD


Notice that 'ty/G decreases by about 50% between T = 200 K and T =
400 K in Figure 4.21, while 't y itself is reduced by twice as much (Figure
4.10). The difference is obviously the temperature dependence of G, an
important subject that is beyond the scope of this discussion. The drop in
'ty/G at T> 0 K is often attributed to thermal energy assisting the stress-
driven change from the undeformed state to the deformed state. Other
experiments show that the strain rate dependence of 't y is influenced by
temperature. These observations are accommodated by Eyring's rate theory
for breaking and reforming (intersegmental) bonds under the combined
influence of stress 't and thermal energy kT, where k is the Boltzmann
constant. A convenient form for expressing the relation between yield
occurring at an applied stress 't = 'ty and macroscopic strain rate y is
(Krausz and Eyring, 1975):

(4. 12a)

Here B is a fundamental rate factor, which may have an inverse temperature


dependence, and AUi and AU: are stress-dependent activation barriers for
displacements parallel (forward) and antiparallel (back) to the applied
stress. With the usual assumptions that each activation barrier is a linear
function of stress, and that AU: > AUi, the back reaction is ignored and
equation 4.12a reduces to

(4.12b)

AHa is the enthalpy of activation in the limit of zero applied stress and Va is
the 'activation volume' for flow of a unit of unspecified size that rearranges
cooperatively during yield. The quantity (AHa - va't y) = Aui is the effective
activation barrier that must be overcome by thermal energy as segments
Constitutive analyses 189

move in the direction of stress Ty ' While Va is called the activation volume,
its physical significance is best conveyed as a factor in the energy term VaTy.
At a constant strain rate, equation 4.12b gives for the yield stress

* kT .
Tv = T - -In(BIY) (4.13)
. Va

The athermal yield stress (at T = 0 K or at the limiting strain rate y = "8) is:
T* = AHa (4.14)
Va

The rate dependence of Ty is used to evaluate va:

dTy kT
(4.15)
din y va
With Va established, the temperature dependence •
and absolute value of T y
give AHa (and hence T*) and the rate factor B.
Analogous expressions for tensile deformation can be obtained with the
von Mises equivalent parameters in equations 4.9 and 4.10. The Eyring
treatment does not invoke specific deformation mechanisms, but certain
features are worth emphasizing. Equation 4.13 nicely captures, at least
qualitatively, the negative temperature dependence and positive rate de-
pendence of T y • Note also that the strain rate dependence (equation 4.15) is
greater at high T Athermal strength T* is proportional to AHa' the barrier
for intersegmental rearrangements, establishing an implicit link between
yield stress and shear modulus G.
The Eyring concept that plastic strain rate y is determined by a combina-
tion of applied stress T = Ty and thermal energy kT is at the heart of many
treatments of yielding in glassy polymers. Different applications are con-
sidered in the following sections.

4.6 CONSTITUTIVE ANALYSES

4.6.1 EYRING MODEL


Experimental rate and temperature dependence of Ty can be expressed as the
parameters Va and AHa in equation 4.12b. One issue, of course, is confor-
mance of data to the Eyring formalism. Figure 5.12 illustrates that do)d lnl':
is constant for PVC at room temperature, leading to Va = 3.1 nm 3 from
equation 4.15. This magnitude of Va corresponds to 40 repeat units within
the chain, but there is no justification for such a literal interpretation
(Bowden, 1973). In Figure 5.12 the history dependence of tensile yield stress
190 Yield processes in glassy polymers

ay at constant strain rate and temperature



indicates that activation enthalpy
I1Ha grows, or perhaps the rate factor B diminishes, during physical aging.
Difficulties often arise when experiments are conducted over appreciable
ranges of strain rate and temperature, as shown for example in the studies
of Brady and Yeh (1971). Apparent I1Ha rose from 160 to 430kJmol- 1
when the deformation temperature of PS passed above 75°C. Although such
temperature anomalies were not observed for PMMA, both va and I1Ha
more than doubled when the average strain rate was decreased by only one
order of magnitude. More complex variations on the basic kinetic model
may be invoked to account for such behaviour (Haussy et ai., 1980; Nanzai,
1993), but these treatments lack general utility. So while the Eyring model
provides analytical relations between yield stress, strain rate and tempera-
ture, the phenomenological parameters Va' I1Ha and B have no a priori
significance and provide little insight into the nature of yield or plastic
deformation. Strain softening, which almost always accompanies yield, is
not addressed by the Eyring model.

4.6.2 VISCOELASTIC MODELS


Certain aspects of deformation and yield may be accounted for with a
simple linear viscoelastic analysis. Consider a tensile experiment for which
one can write

aCt) = Jedt'
0
de
E(t') dt' (4.16)

Letting the strain rate 8 be constant and the relaxation modulus be E(t) =
Eo exp( - t/cJ, where <; is the relaxation time, equation 4.16 gives

(4.17)

The stress-strain curve has a slope Eo at small strains (e « 8<;). The tangent
modulus decreases conspicuously near the 'yield strain' ey = 8<;, beyond
which there is viscous (unrecoverable) flow at a constant stress a y = E 0 8<;.
Although capturing the transformation from elastic to plastic stretch, this
approach has a number of shortcomings. Both yield stress and yield strain
are predicted to be linearly proportional to strain rate 8, while a much
weaker dependence is observed experimentally (Figures 4.11 and 4.12).
Equally serious is the lack of a stress drop or strain softening.
Viscoelastic models can be modified to mimic experimental behaviour
more closely. It is well known that glassy polymers have mechanical
Constitutive analyses 191

responses that require a broad distribution of relaxation times g(~). A


convenient way to represent the effect of such a distribution is the Kohl-
rausch~ Williams~ Watts or stretched exponential function for which the
relaxation modulus becomes (Matsuoka, 1992)
E(t) = Eo exp[ -(t/~YJ (4.18)
Here ~c is the characteristic relaxation time of the system at temperature T;
E(t = ~c) = Eoexp -1 for any value of [3. The width of the distribution
function g(~) is implied by 0 < [3 ~ 1. A single relaxation time ~ = ~c is
associated with [3 = 1, for which E(t) decays exponentially. Increasing the
width of the distribution decreases [3, and E(t) drops first more rapidly, then
more slowly, than for a single ~. During tensile deformation each element
with a particular ~ deforms at the imposed rate eand generates a local stress
given by equation 4.17; the total stress is the sum of local stresses. Matsuoka
(1986) has shown that the yield stress in a multiple relaxation time, linear
viscoelastic solid is (Jy oc en, where n ~ [3. Recall from section 4.3.2 that the
modest rate dependence of (J y was characterized by n ~ 0.05, which is indeed
about the same as [3 obtained from stress relaxation well below ~.
Furthermore, the distribution g(~) narrows ([3 increases) as T approaches ~,
leading to larger rate effects as seen in Figure 4.12. A modified model
invokes plastic, as opposed to linear viscoelastic, flow when the local stress
on a relaxation element equals an athermal strength (J* (Matsuoka, 1992).
The dependence of (Jy on strain rate is the same as described above.
Linear viscoelasticity thus provides a phenomenological framework
for modelling stress~strain behaviour of glasses up to and past the
yield point. Initial modulus is accounted for by Eo, yield point by ~c and
rate dependence by [3. Two additional parameters are needed for tempera-
ture dependence expressed as ~c = ~o exp( -AHa/kT). This approach may be
contrasted to the Eyring model (sections 4.5.2 and 4.6.1) that describes the
magnitude of (Jy and temperature and rate effects through variables B, AHa
and Va' It should be apparent that activation enthalpies AHa have the same
meaning and that ~o ~ B- 1 • The physics of rate dependence is different,
coming from stress-modified kinetic barriers (v;ry) in the Eyring model and
distribution width ([3) in linear viscoelasticity.
Neither approach can account for the drop in stress that accompanies
yield in polymer glasses. Strain softening requires that at least one variable
depends on stress, strain and/or time in such a manner that flow continues
at a stress less than (Jy. If, for example, Va in equation 4.13 were made to
increase with strain ]i, the flow stress in the Eyring formulation would
decrease. We are unaware of any treatment of this sort, but equivalent ideas
are behind non-linear viscoelastic models. An interesting analysis was made
192 Yield processes in glassy polymers

by Knauss and Emri (1987) for uniaxial tensile deformation in which the
relaxation modulus E(t), when written as equation 4.18, has the character-
istic time ~e reduced by stress and time itself. Tensile stress, expressed as
equation 4.16 for a constant strain rate e, increases the macroscopic volume
which, in a delayed fashion, enhances mobility (reduces ~e) in a manner
borrowed from free volume concepts operating above ~. One subtlety is the
stress-volume response function, i.e. the bulk compliance 11K, evolves with
time from a low glassy value to a higher liquid-like value with a time
constant ~' which itself decreases as free volume is generated by the
'transition'. When these concepts are incorporated in an otherwise linear
viscoelastic model just a few degrees below ~, tensile yield occurs with a
stress drop (intrinsic strain softening) that is more pronounced at higher
strain rates, just as seen in Figure 4.12. Other calculations showed that
hydrostatic pressure P caused yield at larger (Jy and with less strain
softening, as suggested by experiments in Figure 4.13. In this model the yield
point derives from the time-delayed autocatalytic increase in volume and
associated mobility. Such a mechanism is inapplicable to compression or
shear in which macroscopic volume decreases or is unchanged. Strain
softening under those conditions requires another explanation.
A less mechanistic treatment of a similar type was performed by Rendell
et al. (1987). These authors use the stretched exponential form of the
relaxation modulus E(t) in equation 4.18, but with different physical
interpretation. Macroscopic non-exponential decay is ascribed to a single
relaxation process with elementary time constant ~. The relaxation rate is
lis at very short times, then slows when t exceeds a critical time 1/we
because of cooperative 'coupling' of motions that occurs at large time and
length scales. The result is a relation between characteristic time ~e and the
elementary time s:
(4.19)

The expression for tensile stress (J(B) obtained with equations 4.16 and 4.18
is of course identical to the case where f3 reflects a distribution of relaxation
times, but here the microscopic stress is uniform. Rendell et ai. (1987) further
allowed f3 to increase linearly with strain in a manner suggested by visco-
elastic response experiments. This modification reduces the characteristic
time ~e' making the material more compliant and results in a stress
maximum followed by strain softening during calculated tensile deformation
at constant rate e. A similar effect was obtained when ~ was lowered by the
application of stress. No physical reasons were presented for the increase in
f3 or reduction in ~, but these sources of non-linear response indeed result
in yield followed by a stress drop.
Constitutive analyses 193

4.6.3 STRAIN SOFTENING


The non-linear viscoelastic treatments above have stress or strain modify
the system homogeneously to account for strain softening. An alternative
approach is purely mechanical; if some local regions are assumed to undergo
plastic deformation, strain gradients between regions of high and low shear
generate elastic misfit stresses that affect subsequent deformation of the
solid. The relation to strain softening is in the directions of the misfit forces,
some of which assist and some of which oppose the macroscopic stress 1:.
This has been elucidated with a simplified two-dimensional system by
Bulatov and Argon (1994) and Argon et al. (1995), who employ an
elastic-plastic model in which initial order (packing uniformity) can be
varied between crystalline and glassy. There is no intrinsic strain softening
in the local material response. Each element may undergo an elementary
shear transformation Yo = 2 with a probability or rate determined by the
Eyring model (equation 4.l2b), where local stress on the element is estab-
lished by the degree of disorder in the initial structure and by elastic forces
created when deforming the model at a constant tensile strain rate. Yield
starts with isolated plastic events and continues with strain tending to
concentrate in shear bands oriented at 45° to the loading direction. This
strain localization is accompanied by a drop in overall plastic resistance or
flow stress to a steady state value. In more ordered (better packed)
structures the yield stress is higher and both strain softening and strain
localization are enhanced, just as seen with physically aged polymer glasses
(Figure 4.1 and Figure 4.17a). The most disordered or 'glassy' structure has
many isolated, highly stressed elements that contribute to the plastic strain
at low system stress, and strain localization and stress drop are suppressed.
Finally, the elastic strain energy associated with misfit stresses rises most
rapidly near the yield strain, then levels off at a constant value, exactly as
seen with deformation calorimetry (Figure 4.19).
Hence virtually all phenomena associated with strain softening in glassy
polymers are recovered by the Bulatov-Argon model. There is no 'structure'
change associated with free volume, dislocation density, chain orientation,
etc., only elastic misfit stresses generated by shear transformation of elemen-
tary units that facilitate local yielding in nearby regions. It is easy to imagine
that shear cycling increases disorder and associated internal stresses suffi-
ciently to reduce !y and suppress strain softening. The effect of hydrostatic
pressure is more difficult to understand. One would expect the compressed,
closer packed structure to be more ordered and hence display more strain
softening, whereas the opposite effect is observed experimentally.
Hasan and Boyce (1995) have developed a model of yield in polymer
glasses that focuses on the strain energy, rather than directionality, of misfit
194 Yield processes in glassy polymers

forces arising from local plastic events. This is an elastic-plastic analysis


where the plastic strain rate Yp is expressed in terms of strain-dependent
parameters that cause the flow stress t to decrease after yield. Plastic strain
rate is described by the usual Eyring formalism with two initial variations;
'back reactions' or strain steps against the direction of t are retained, and a
distribution of activation energies g(I1H a), reflecting microstructural irregu-
larity within the glass, is added. The plastic strain rate analogous to
equation 4.l2a now becomes

Here the rate factor B and forward activation barrier l1ui = I1Ha - Vat are
the same as in the Eyring model. The back reaction has activation barrier
I1U~ = I1Ha + Vat that is effectively reduced by local elastic strain energy S
developed when an elementary region has undergone a plastic event. In
contrast to the Bulatov-Argon treatment, this is a homogeneous deforma-
tion model without explicit coupling of local misfit stress to neighbouring
regions.
The total strain rate has elastic and plastic components:

(4.21)

In a constant strain rate experiment the response is elastic until a sensible


number of local plastic events occur in those regions with lowest activation
energies I1Ha. The slope of the stress-strain curve then decreases, but much
of this 'plastic' strain is recoverable. If the strain is reversed (the sample
unloaded) many of the plastic events also reverse because local strain energy
S helps overcome the modest activation barrier I1U~ for the back reaction.
To this point the model is equivalent to a linear viscoelastic solid with a
distribution of relaxation times; there is a gradual transition from recover-
able to partially unrecoverable strain, rate sensitivity of the yield stress, etc.,
but no strain softening.
Hasan and Boyce impose a strain dependence on both the distribution of
activation energies g(I1HJ and the local elastic strain energy S according to
the following concepts. With increased plastic strain the low I1Ha portion of
the distribution becomes exhausted and the strain energy S rises as the
number (or volume fraction) of transformed sites grows. Macroscopic yield
occurs when S reaches a critical value and more elastic energy cannot be
stored in the matrix. The entire system - local plastic regions and elastically
distorted matrix - undergoes a transition to a new distribution of activation
Constitutive analyses 195

100

-IZS

-
0..
:::s

o ~~~~~~----~--~~--~--~~--~
0.0 0.1 0.2 0.3
TRUE STRAIN
Figure 4.22 Experimental (points) and calculated (lines) response of PMMA loaded
and unloaded in uniaxial compression at room temperature, strain rate 8t = -1 X
10- 3 S-1. (Reproduced from Hasan and Boyce, 1995, with permission.)

energies having smaller values of !J.Ha and a reduced S. The coupling of S


and g(!J.Ha) to plastic strain fp is described with empirical equations.
The fit to experimental uniaxial compression data for PMMA is quite
good, as shown in Figure 4.22 where model response in shear was converted
to uniaxial stress and strain by the von Mises relations. This approach
accounts nicely for both strain softening and 'viscoelastic' retraction from
different imposed strains. Agreement is equally satisfactory at other strain
rates and temperatures, and furthermore describe experimental creep behav-
iour with the same model parameters.
196 Yield processes in glassy polymers

The basis for the strain-driven transformation of g(t1.H a) is PALS


experiments indicating that average free volume increases after compressive
yield (Hasan et ai., 1993). Hasan and Boyce (1995) associate regions having
larger free volume with a lower activation enthalpy t1.Ha• Strain softening
results directly from the creation of larger free volume sites, here expressed
as lower ~Ha' at strains beyond the yield strain By'" -0.1. Elastic
strain energy S accounts for the shape of the stress-strain curve before yield
and for strain recovery on unloading, but is not significant for strain
softening. There is a general relation to the Knauss-Emri model discussed
in the preceding subsection, but Hasan and Boyce work in the context of
elastic-plastic deformation of a solid, and free volume is not coupled to
macroscopic volume. Yet another mechanism for strain softening, where the
'structural' change is dislocation density instead of free volume, is presented
in section 4.7.3.

4.7 MOLECULAR MODELS


Here we describe more detailed treatments that employ molecular scale
deformation of polymer chains to account for yield. Attention is focused on
the yield stress and, where applicable, strain softening.

4.7.1 ROBERTSON MODEL


The first molecularly based analysis of yield in polymer glasses is from
Robertson (1966). This is a 'transition' model in which applied stress
increases mobility to liquid-like values associated with a fictive temperature
() > ~. Relevant energy derives from the rotational states of backbone
bonds as sketched in Figure 4.3c. The equilibrium fraction of bonds in the
high energy or 'flexed' rotational state, with excess energy t1.E, increases with
temperature in the liquid; the flexed fraction present in a glass defines the
fictive temperature () at which this fraction would be the equilibrium value.
Note that ~E is intramolecular in origin. The unstressed glass at tempera-
ture T < ~ is not in equilibrium, having a flexed bond population frozen in
at () = ~. Robertson showed that a shear stress 1: applied to a polymer glass
will cause a transient increase in the number of flexed bonds, thereby raising
the fictive temperature () above ~. The functional relation between () and 1:
involves t1.E and a volume Vi that couples stress to energy, as in the Eyring
model (equation 4.l2b). These two variables have well defined meanings and
values; ~E '" 6 kJ/mol can be measured with IR spectroscopy from the
temperature dependence of rotational states above ~ (Theodorou, Jasse and
Monerie, 1985), and Vi '" 0.15 nm 3 is the volume of a two-bond repeat unit
in the chain.
Molecular models 197

Chains in the stressed glass thus have greater internal energy that permits
flow as a Newtonian liquid at a fictive temperature () > ~. Equating the
applied stress L to the yield stress one obtains Ly = f/«())Y, which connects Ly
with strain rate y in terms of viscosity f/ at the fictive temperature (). The
viscosity-temperature relationship for () ~ ~ is given by the empirical
Williams-Landel-Ferry equation with universal constants C~ = 2.303 x
17.4 = 40.1 and C 2 = 51.6K (Williams, Landel and Ferry, 1955). The result
for shear deformation at absolute temperature T and rate .;. is

(4.22)

With fIg = 10 13 .6 Pa s for the viscosity of a glass at 1~, there are no


adjustable parameters in this approach. Robertson (1966) evaluated equa-
tion 4.22 numerically with values of f1E and v' appropriate for PS and
PMMA to define the relation between () and L y • The model is remarkably
successful within c. 80 K of ~, accounting for absolute values of Ly to better
than a factor of two, and temperature and strain rate dependence with even
greater accuracy. Argon and Bessonov (1977b) correctly point out that the
'y
Robertson model predicts that is much too large at low temperatures. It
is reasonable that the 'transition' to liquid-like flow can only be achieved in
glasses that are near ~.
It was established in section 4.5.1 that shear yield involves disruption of
intermolecular bonds. Robertson's model incorporates intermolecular bond-
ing indirectly through the glassy viscosity fIg and the WLF dependence of
viscosity on e, a combined strategy that seems to predict macroscopic yield
at temperatures in the neighbourhood of Tg• Independent support for the
Robertson model comes from the experiments of Theodorou et al. (1985)
presented in section 4.4.4. Recall that those authors found by IR spectros-
copy that the population of high energy gauche rotational isomers increased
as PS yielded in tension at T ~ ~. This is precisely the stress induced
structural change that raises () above ~ and activates liquid-like viscous
flow. However, room temperature deformation of PS and PET decreases,
not increases, the number of gauche bonds (Magonov, Vainilovitch and
Sheiko, 1991). This underscores that Robertson's model is restricted to
temperatures close to the glass transition region.
The earliest concepts of yield in polymers invoked adiabatic heating or
free volume to transform the glass to the liquid state. Robertson's model
accomplishes the same 'state change' with stress modifying the internal
energy of the polymer chains. This approach, while accounting nicely for
absolute magnitude of 'y and temperature and rate effects, does not address
strain softening. The increase in Ly after physical aging could perhaps be
198 Yield processes in glassy polymers

associated with a lower fictive temperature ein the unstrained state, but that
idea has not been pursued.

4.7.2 ARGON MODEL


A treatment that deals explicitly with intermolecular resistance to shear
yielding was developed by Argon (1973). He proposed that local shear
strains in a polymer glass may be represented by the rotation of a chain
segment of length z into the direction of principal stretch. Segment rotation
is accomplished by generating two kinks, separated by distance z along the
chain, each with a bending angle w ~ 2 radians (115°, the C-C-C bond
angle). A kink pair can be formed by twisting two nearby backbone bonds
through equal and opposite displacements. The resultant strain is propor-
tional to a2 w 2 z, where a is the radius of one chain or a cooperatively
deforming bundle of chains. Argon calculates the net energy ~u of a kink
pair of size z from elastic distortion of the surroundings having shear
modulus G and Poisson ratio v, reduced by work done by the applied stress
,. It is important to note that resistance to kink formation is purely
intermolecular; ~U ignores the internal energy of the kinked chain, i.e. ~E
in Robertson's theory. There is an extremum in ~U as the kink pair size z
increases at constant ,. The maximum value ~U*, when z = z*, is the
activation barrier for nucleating a stable kink pair, i.e. one for which the
energy falls after it grows beyond the critical size z*.

. . [-AU*('t)] (4.23)
y=Bexp kT

Again the applied stress has been equated to yield stress 'y. This is the
Eyring model (equation 4.12a) where the back reaction is neglected and
~U* = ~Ui derives from the specific energy surface for kink pair growth in
an elastic solid; ~U* is not a linear function of stress. After rearranging one
obtains for the yield stress 'y at temperature T and strain rate y:

,
y
=- G[
8
1 - - - l n (B)J%
kT
Gw 2a 3
-
y (4.24)

Here we have let v = 0.38 (Brown, 1983), which coincidentally simplifies


some constants. The Argon result closely resembles the Eyring model
written as equation 4.13; differences stem from absolute rate theory here
being applied to a particular, as opposed to a general, process. The
fundamental rate factor B is the same as in the Eyring analysis. The
exponent of ~ leads to temperature and rate dependencies that are virtually
indistinguishable from equation 4.13 when ~Ha and Va are constant
Molecular- models 199

['=
l
I ' I ' I ' I '

0.12

CD
LO
---
(.!)
008 f ••


• •
...... --.---
>- : • --1
-.!:::- 0.04 ~
J

~
O.OO~~ i I I I ~~
~
a
~

0.0 0.1 0.2 OJ 0.4 0.5


T/G (KlMPa)
Figure 4.23 Normalized shear strengths from Figure 4.21 replotted according to the
Argon kink pair nucleation model. The model predicts an intercept of 0.17 at
T = 0 K, considerably above the data.

(Crist, 1993). Athermal strength r* = G/8 is determined by the shear


modulus G, not the ratio of phenomenological parameters I'1Ha/va from the
Eyring model. The group w 2 a 3 is comparable to the activation volume Va in
determining rate dependence dry/d In y, but both wand a are directly related
to chain architecture.
s
Argon's kink pair model lends itself to the format of (rjG)6 versus T/G,
as in Figure 4.23 where data from Figure 4.21 are replotted. The intercept
at T = 0 K here gives r* = G/16, just one-half that expected from equation
4.24. It should be emphasized that the Argon analysis accommodates
implicitly the temperature dependence of G. Argon and Bessonov (1977b)
and Yamini and Young (1980) found that the kink pair nucleation model
accounts for the modest temperature dependence of ry/G observed in many
glassy polymers, provided T is sufficiently far below Tg• The two derived
parameters have reasonable values; B ~ 10 13 S - 1 and the effective chain
radius a ranges from 0.8 to 3 times a single chain radius for ill = 2.
Both Robertson's and Argon's schemes rely on stress causing internal
rotation about backbone bonds, although the two models diverge beyond
this common origin. Robertson envisions the modified rotational state
serving as a 'trigger' to activate liquid-like flow with no direct accounting of
intermolecular forces. Argon, on the other hand, couples bond rotation
(kink pair formation) to intermolecular energy calculated with continuum
elasticity theory, and has the macroscopic strain rate y result directly from
the frequency of kink pair formation. The reasonable success of the kink pair
200 Yield processes in glassy polymers

nucleation model should not be taken as proof that the mechanism is


correct in detail; the important conclusion is that resistance to plastic
deformation at the yield point is intermolecular in origin. This model
does not, however, account in any obvious way for strain softening. Argon
(1973) comments that an existing kink pair locally distorts its environ-
ment, perhaps facilitating the nucleation of kinks in nearby positions as in
the Bulatov-Argon model in section 4.6.3, but this idea has not been
implemented.

4.7.3 DISLOCA nON MODEL


Bowden and Raha (1974) start from the observation that shear strain may
be very inhomogeneous during yield; the microscopic shear bands in PS
(Figure 4.17a) are an extreme example. These features resemble slip bands
in metals, where it is known that plastic strain results from mobile
dislocations. Bowden and Raha argue that dislocations, i.e. linear regions
where the structure is 'defective' and local stress is large, may control plastic
deformation in a non-crystalline material. Although dislocations are usually
associated with crystals, the concept can indeed be extended to glasses
(Escaig, 1984; G'Sell, 1986). Thus motivated, Bowden and Raha (1974)
developed a thermally activated rate model for yield in polymer glasses
using the energetics of dislocation loops. Macroscopic plastic strain at stress
'y results from local shear strain of order 1 in many disk-shaped volume
elements of radius r. The top and bottom surfaces of the disks are mutually
displaced in the stress direction by the Burger vector b. As the implicit
height of the disk is also b, the elementary strain is Yo ~ 1. The dislocation
strength b ~ 0.4 nm reflects the average intermolecular spacing in the glass.
A shear patch with the Burger vector b is nucleated and grows in radius
r under a constant stress, = 'y. The energy cost flU is calculated by treating
the shear patch as a dislocation loop of radius r. Elastic strain energy
emanating from the dislocation (a line of length 2nr surrounding the shear
patch) is expressed in terms of dislocation strength b and shear modulus G.
Work done by the stress 'y reduces flU, which thus has a maximum flU*
when r achieves the critical value r*:

Gb 2-
flU* = - r* [(2r*)
In - - 1] (4.25)
4 ro

Here ro is the radius of the dislocation core that Bowden and Raha (1974)
set roughly equal to b to make the athermal strength of their model = ,*
G/(3 1 / 2 n), which is about 50% larger than ,*
from the Argon model. The
critical barrier flU* for nucleating a dislocation loop of radius r* is
Molecular models 201

substituted into equation 4.23. Stress dependence of AU* In equation 4.25


is based on an implicit relation between Ty and r, meaning that one cannot
write an analytical expression for the yield stress analogous to equation 4.13
or equation 4.24.
With the postulate that ro ~ b, the dislocation nucleation model has
only two parameters, band B (or the value of AU*/kT that gives the
macroscopic strain rate)' ~ 10- 3 S-I). Bowden and Raha 11974) effectively
set B ~ 10 12 S - 1 and evaluate from the experimental yield :;tress Ty at room
temperature the dislocation strengths b = 0.27-0.49 nm for five glassy
polymers. The model accounts reasonably for observed temperature and
rate dependence of yield stress in PMMA.
As with the kink pair nucleation model, macroscopic yield is attributed
to formation of isolated microscopic plastic zones that are opposed by
intermolecular forces calculated with continuum elasticity and expressed in
terms of the shear modulus G. Satisfactory agreement IS obtained with
experiments, at least over limited ranges of temperature and strain rate, for
plausible values of the molecular scale parameter b for the dislocation model
or a for the Argon model. The models differ in other ways. First, the
dislocation nucleation treatment predicts a rapid increse of Ty toward T* as
T approaches 0 K (Argon and Bessonov, 1977a; Crist, 1993). This is not seen
in experiments at or above 77 K (Figure 4.21), hence the kink pair model
may be more appropriate. However, Bowden and Raha (1974) emphasize
that strain softening is easily accounted for by dislocation loops. Yield at the
imposed strain rate is achieved by nucleation of loops with critical size
r* ~ 5b. Further deformation may occur by growth of stable loops having
r > r*, which contribute to the strain proportionally to r2. Bowden and
Raha contend that loop growth needed to maintain the strain rate y
proceeds at a stress T less than T y • Their qualitative discussion asserts that
dislocation loop growth will lead to stress drops much larger than those
observed, hence mechanisms are proposed to limit the amount of strain
softening.
The relation between dislocations and strain softening in glassy polymers
was addressed from a different perspective by G'Sell and Jonas (1981) and
GSell (1986). Here the creation and mobility of dislocation lines are
considered, as opposed to the nucleation of effectively static loops in the
Bowden - Raha model for T y ' The approach starts with the Orowan expres-
sion where the plastic strain rate is the product of dislocation density Pd'
dislocation velocity Vd and magnitude b of Burger vector:
(4.26)
Velocity Vd is an exponential function of stress T, essentially that given by
the Eyring equation 4.12b. This expression is for deformation via shear
202 Yield processes in glassy polymers

bands or shear zones with dimensions spanning the macroscopic sample, as


in Figures 4.16 or 4.17. An increase in Pd refers to multiplication of
dislocation lines running between external surfaces, with no consideration
of how growth to this size occurs. Strain softening is achieved by having the
dislocation density Pd increase (linearly) with strain y toward a value that is
established by the plateau flow stress at large strains. Plastic strain rate from
equation 4.26 is incorporated into the elastic-plastic rate expression, equa-
tion 4.21, which is evaluated numerically. The model indeed reproduces true
stress-strain curves with stress drops that decay to plateaus, as in Figure
4.1. Structure evolution, in this case dislocation density, is described by the
steady state value of Pd and a critical strain at which this is attained. The
authors make no claim that the model is correct in detail, but emphasize
that it captures strain softening and related effects seen when strain rate is
varied. History dependence of dislocation density can also account for
physical aging and softening after mechanical cycling.

4.8 MOLECULAR SIMULATIONS


With increased computing capabilities it is now possible to model the
mechanical properties of glassy polymers at the molecular level. Such
simulations can provide information on local structural changes unavailable
from direct experiment, and thus supply possibly valuable insights on
yieding and post-yield deformation. Both molecular mechanics and, less
frequently, molecular dynamics simulations have been carried out. We here
outline the methods and discuss the information derived from such studies.

4.8.1 MOLECULAR MECHANICS


Theodorou and Suter (1986) were the first to model the three-dimensional
structure and elastic properties of a glassy polymer. Their method is
considered in sufficient detail to provide a background for molecular
mechanics treatments of plastic deformation. An undeformed structure is
created by energy minimization at constant volume, in this case atactic
polypropylene (a-PP) at a density of 892 kg m - 3, characteristic of tempera-
ture T = -40°C, which is below the laboratory glass transition temperature
1'g "" -lO°e. The structure is based on a chain of 76 repeat units (152
backbone C atoms; M = 3194), segments of which occupy a cubic cell of
edge length 1.815 nm replicated with periodic boundary conditions. Energy
is reckoned with empirical threefold potentials for rotations about skeletal
C-C bonds and Lennard-Jones potentials for interactions between atoms
on different segments. Covalent bond lengths and angles are fixed at
appropriate values, and do not contribute to energy changes.
Molecular simulations 203

The unit cell contammg an undeformed structure is subjected to a


predetermined strain, and the energy is minimized within the deformed cell
by adjusting intrachain rotation angles and intersegmental separations and
orientations. For the study of elastic properties, Theodorou and Suter (1986)
applied hydrostatic strain (E I = f,2 = ( 3), pure shear strain (f,1 = -E 2 , E3 = 0)
or uniaxial strain (£1 #- 0, 1:2 = £3 = 0), where the strain magnitude was
8 1 = ±0.00l. Relevant elastic stiffness constants Cij were obtained from the
change in either potential energy U or the internal stress tensor (Jij" Model
properties expressed as Young's modulus E ~ 2.9 G Pa, shear modulus
G ~ 1.1 GPa, bulk modulus K ~ 3.3 GPa and Poisson ratio v ~ 0.36 are in
excellent agreement (c. 5-15%) with limited experimental data for a-PP at
-40°C.
This approach was extended to strains as large as 8 ~ 0.2 by Mott,
Argon and Suter (1993) to study yielding and irreversible or plastic
deformation. Two strain fields were applied incrementally to the unit cell,
uniaxial tensile strain at constant volume (Poisson ratio v = 0.5) and
pure shear, which is also at constant volume. System energy U was
minimized at each strain by adjusting rotation angles and intersegmental
separations and orientations, as in the small strain elastic study. It is
important to emphasize that temperature has no role in this static molecular
mechanics simulation; there is no thermal energy kT to assist in surmount-
ing barriers to local rearrangements. Hence the calculation corresponds to
deformation at T = 0 K, although the model structure has the density of
a-PP at T = 233 K (-40CC). Results are presented as internal stress (J as a
function of applied engineering strain En' with von Mises equivalent stress
(Jeq = d l/2 and strain Beg = 1'/3 1 / 2 for shear deformation (equations 4.9 and

4.10).
The stress-strain response in Figure 4.24 is based on internal stress
calculated for uniaxial tensile deformation (v = 0.5) and pure shear defor-
mation. Non-zero stress in the un strained structure results from force
imbalances inherent with non-uniform structures in small fixed volumes.
One sees immediately that uniaxial and shear behaviours are normalized
with the von Mises relations appropriate for isotropic elastic solids, and that
initial stiffness corresponds to elastic properties deduced for the same system
at small strains (Theodorou and Suter, 1986). What appears as 'noise' are
actually abrupt internal stress changes, usually negative, that accompany
plastic events pervading the entire simulation cell of volume (1.82 nm)3. This
is shown for a single simulation of uniaxial extension in Figure 4.25. The
vertical axis is the shear strain d}' of the small volume element associated
with each of the 152 backbone carbon atoms in the cell. Structural
rearrangement is by entirely cooperative events that occur with increasing
frequency as the system strain is increased. Not shown in Figure 4.24 is that
204 Yield processes in glassy polymers

300~-------r--------T-------~--------~

Axial Extensi~n EnSemj~

:1 J 1'\ :', ,

-
I \ I' I
200 ............. : ... /. . ...... : •.,..,\ 1rA ....;)' \/Y.v. .\ i \t t .
Elastic Lo~ding / . "\
0
a.. : / , '\1 / "J\ I : \' \
.......
~ 4:// I, ,)\ ~ ~I i 11\
f,: I" 1\ ~
<r ~ /1 ~ ";';"IfoJ'N"

100 ...... I '. : ....... ~.h~?!. ~~~~~~I~ ..... -:............ .

0.05 0.10 0.15 0.20


I:
eq

Figure 4.24 Molecular mechanics simulation of deformation of a-PP at T = 0 K.


Solid line is for shear deformation (von Mises equivalent stress and strain) and dashed
line is for uniaxial tensile deformation at constant volume. Elastic loading is the
extension of small strain behaviour. Stress jumps are related to plastic events; see
text and Figure 4.25. (Reproduced from Mott, Argon and Suter, 1993, with
permission.)

system deformation is irreversible after one of these plastic events, although


it is reversible between them.
The picture that emerges is a gradual elastic-to-plastic transition with
increasing strain, without a well defined yield point or strain softening. Flow
occurs at a nearly constant stress of '" 190 MPa for strains greater than 0.1.
Imposed system strain is accommodated by internal rotation and segmental
packing changes that increase the (elastic) energy of the cell. Part of this
energy is dissipated by cooperative rearrangements - plastic events - that
occur more frequently (at smaller strain intervals) as strain is increased.
Deformation is formally irreversible after the first plastic event, generally
seen at strain eeq;;:: 0.Q1. Mott, Argon and Suter (1993) analysed the
structural changes in detail but found no common local rearrangement,
analogous to kink or dislocation nucleation, etc., during plastic events.
Strain magnitudes associated with individual plastic events were calculated
from the size of abrupt stress changes in Figure 4.24 and elastic modulus;
the ensemble average elementary shear strain is 'Yo = 0.015. Limited studies
Molecular simulations 205

Figure 4.25 Local shear strains dy at each of 152 backbone C atoms during one
simulation of tensile deformation of a-PP. Every atom rearranges during a plastic
event. There are seven events in this plot, six occurring at Ceq> 0.1. (Reproduced
from Mott, Argon and Suter, 1993, with permission .)

with cell volumes up to six times larger gave the same results. In particular,
plastic events pervade cells with linear dimensions of --4 nm.
The same static molecular mechanics analysis has been conducted on PC
by Hutnik, Argon and Suter (1993). Glassy structures with a density of
1200 kg m - 3, characteristic of room temperature, were generated in a cubic
cell with edge length 1.84 nm. In this case the calculated elastic moduli,
E -- 5.6 GPa and G -- 2.1 GPa, were about twice the experimental values at
room temperature. Large strain deformation in pure shear gave results
comparable to those obtained for a-PP above. Internal stress first rises, with
small discontinuities, in a manner consistent with the elastic shear modulus
G. For strains Ceq ~ 0.1 the stress increases much more gradually, fluctuating
about a level of about O"e q -- 150 MPa. Abrupt stress drops accompany
plastic events involving cooperative displacements of all atoms in the cell,
as in Figure 4.25. For PC the average elementary shear strain is Yo = 0.012,
similar to that in a-PP.
Deformation of amorphous PC was also simulated, with some different
assumptions, by Fan (1995). Energy minimization included changes in
covalent bond lengths and angles, as well as internal rotation states and
206 Yield processes in glassy polymers

intersegmental separations. The undeformed state is an isotropic glass of


density p = 1150 kg m - 3, corresponding to somewhat above room tempera-
ture. Uniaxial tensile strain, with a prescribed Poisson ratio v = 0.27, was
applied to the unit cell of original edge length 1.97 nm. Hence the cell or
system volume increased by 11 % at the largest tensile strain B = 0.3. In each
of 18 simulated deformations, glassy PC manifested yield by a lower slope
da/dB and irreversible stress-strain behaviour. Individual runs lacked the
'sawtooth' character of Figure 4.24, perhaps because of volume expansion,
and were quite variable, with apparent yield or flow stress ranging from
200 MPa to 350 MPa. Most of the simulations showed strain softening
beyond the yield strain By '" 0.12, which is correctly attributed to anhar-
monicity of the Lennard-Jones functions describing intermolecular interac-
tions. Strain softening in this model is a direct consequence of dilatation.
Not surprisingly, more than 95% of the deformation energy is associated
with changes in intermolecular bonding. No specific structural changes were
seen, in accord with similar negative observations by Mott, Argon and Suter
(1993).
Molecular mechanics techniques give possible pictures of deformation of
glassy polymers in the absence of thermal activation, i.e. at T = 0 K. The
simulations of Mott, Argon and Suter (1993) and Hutnik, Argon and Suter
(1993) at constant volume show that plastic strain may accrue by intermit-
tent, cooperative relaxations of elastic energy involving hundreds of back-
bone atoms throughout regions greater than (3 nm)3. In sharp contrast to
the local shear displacement model in Figure 4.2, plastic strain evolves in
very small increments of Yo '" 0.01-0.02 over large volumes. Fan (1995)
assumed an unrealistically large dilatational component of strain (L1 V IV up
to 0.11); plastic strain accumulates steadily and, by inference uniformly,
throughout the simulation volume of (2 nm)3.
It is difficult to relate these athermal simulations to experimental results.
The two approaches give similar normalized strengths for PC; Ty/G '" /6
(Hutnik, Argon and Suter, 1993) or /5 (Fan, 1995). One might expect that
constraints inherent in static molecular modelling would lead to yield stress
larger than experimental, but these simulation values are indistinguish-
able from Ty/G = /6 observed for PC at low temperatures (Figure 4.21).
It appears that, molecular mechanics gives reasonable estimates of the
normalized stress required for plastic deformation of a polymer glass.
Hutnik, Argon and Suter (1993) further combine Yo, from the simulation of
PC, and activation volume Va' from the experimental rate dependence of Ty '
to estimate the size of a cooperatively rearranging unit. The result
Q '" (8 nm)3 is larger than the simulation cell volumes and hence is
consistent with the molecular mechanics results.
Molecular simulations 207

4.8.2 MOLECULAR DYNAMICS


Molecular dynamic simulations include thermal energy to assist the
deformation process. Brown and Clarke (1991) have modelled the mechan-
ical response of an amorphous polyethylene-like chain of 1000 united atom
CH 2 groups under uniaxial tensile stress at different temperatures. One
should appreciate that the total time of such simulations is very small,
typically less than 1 ns. The usual sorts of rotational and Lennard-lones
potentials were used. While C-C bond lengths were fixed, bending vibra-
tions about the equilibrium C-C-C bond angle of 112.8 c were allowed.
Constant temperature deformations were simulated between T = 10 K and
T = 500 K at loading rates of 0.1 MPa ps -1 and 0.5 MPa ps - 1. It is import-
ant to note that the apparent glass transition temperature is Tg '" 400 K for
the time scale of these 'experiments'. CH 2 units with average energy kT
move by Newton's equations of motion under the combined forces from
intramolecular (C-C-C angle vibration and internal rotation), intramolecu-
lar (Lennard-lones) and external (applied stress) sources. System dynamics
determine the strain of the cell, which may change shape and volume.
Internal tensile stress a is calculated as in static molecular mechanics
simulations, with the addition of momentum or 'pressure' terms. The
internal stress is not equal to the applied stress; the latter increases linearly
with time.
Results of true internal tensile stress versus nominal strain were presented
in Chapter 2. Referring to Figure 2.13, one sees clearly the transition from
liquid-like to glassy behaviour below T = 400 K '" ~. The initial slope
(modulus E) is large, yield is signalled by an abrupt slope change with strain
softening at the two lowest temperatures. Be advised that these simulations
have extremely large ('" 109 S -1) and variable strain rates, hence comparison
to experiment requires caution. In particular, the strain rate increases
rapidly beyond the yield point, possibly reducing the post-yield stress drop
or strain softening. Regardless of special aspects of the simulation, molecular
dynamics show that a y is increased at lower temperatures. Yield and flow
stress also increase with strain rate (actually applied stress rate), in quali-
tative agreement with experiment. Normalized shear strength at T = 10 K
is Ty/G = i" about four times larger than the molecular mechanics results for
PC in the previous subsection. Ty/G increases by about 10% at T = 200 K,
but recall that molecular dynamics simulations are not done at a constant
strain rate.
The discussion is aided by considering the density in Figure 2.16. At
T = 500 K the deformation occurs at essentially constant volume, as ex-
pected for a liquid. Lower temperature naturally increases the equilibrium
(un strained) density, and also the amount of dilatation during tensile
208 Yield processes in glassy polymers

deformation. The Poisson ratio drops from v '" 0.49 at T = 400 K to


v = 0.41 at 10 K. The obvious and interesting feature is that density
p '" 755 kg m - 3 (0.755 g cm - 3) is the same at large strains 8 ~ 0.4 for all
temperatures between 10 and 400 K. Recalling that Tg '" 400 K, the manifest
implication is that volume expansion during tensile deformation indeed
transforms the glass to a liquid! The structure, or at least the density, of the
deformed state is constant, and plastic flow occurs at a temperature-
dependent stress that can be treated as with the Eyring model. We point out
that strain softening, seen at the two lowest temperatures, is associated with
dilatational strains on the order of 0.1. Hence it is likely that the stress drop
results from intersegmental displacements approaching the inflection points
of the Lennard-Jones potentials, as discussed by Fan (1995).
Acknowledging limitations set by small time scales of molecular dynamics
calculations, the work of Brown and Clarke (1991) demonstrates that tensile
yield may occur by a volume-driven transition to the liquid state. In
laboratory experiments with much smaller strain rates there is no evidence
for the large dilatation required to achieve this 'transition'. Most welcome
would be a molecular dynamics study of compressive or shear deformation
in which dilatation was suppressed.

4.9 CONCLUSIONS

4.9.1 YIELD STRESS


Reduced yield strength is in the range Ty/G = 0.06-0.10 for virtually all
glassy polymers at the lowest experimental temperatures T '" 170 K (Argon
and Bessonov, 1977b; Yamini and Young, 1980). Similar values of reduced
athermal strength T* /G are obtained by extrapolation to T = 0 K; see Figure
4.21 and Brown (1983). Yielding at very low temperatures can be attributed
to plastic strain from segment reorientation (Argon, 1973) and/or shear
displacements (Brown, 1983), which are calculated occur at similar stress
levels. The important conclusion is that resistance to plastic deformation is
intermolecular in origin. Both temperature and rate dependence of Ty are
accounted for by thermal activation of these elementary processes; Argon's
modification of the Eyring approach, with explicit reference to shear
modulus G, is quite satisfactory (Figure 4.23). The decrease of Ty/G with
temperature is typically about 50% between T = 200 K and T = 400 K,
although Ty changes appreciably more because of the T-dependence of G.
We conclude that the absolute value and temperature and rate dependencies
of yield stress, expressed as Ty/G, are understood.
Conclusions 209

Physical aging increases the yield stress by 10-30%; sec for example
Figures 4.1 and 5.12. Since elastic modulus is enhanced by a similar amount
(Tant and Wilkes, 1981; Diaz-Calleja, Ribes-Greus and Gomez-Ribelles,
1989), it is likely that the important ratio ty/G or (JylE is not much affected
by thermomechanical history. Hydrostatic pressure is well known to in-
crease both yield stress and elastic stiffness. The relative contributions to
enhanced ty(P) resulting from G and other effects associated with elastic
compression have not been addressed quantitatively, although pressure may
reduce the rate factor B in kinetic expressions like equations 4.12 and 4.23
(Argon, 1973).
The preceding statements apply to glassy polymers more than c. 25°C
below ~. It is likely that yield close to the laboratory glass transition
temperature is accomplished by homogeneous viscous flow. Robertson
(1966) provides a general mechanism for increasing mobility to liquid-like
values. The same effect may be achieved by dilatational strain in tension if
free volume is linked to macroscopic volume (Knauss and Emri, 1987). Yield
measured under different stress states may be instructive in this regard. It
was pointed out in section 4.3.1 that the ratio (Jy/ty '" 1 for PC at T = 135°C
is unusually low. Dilatation leading to reduction of (Jy in tension is plausibly
explained by the Knauss- Emri mechanism, while ty in shear is of course
unaffected. In the absence of more reliable data, high temperature deforma-
tion is not considered further.

4.8.2 STRAIN SOFTENING


The drop in stress associated with plastic deformation beyond the yield
point is nearly ubiquitous, exceptions being noted close to Tg and in highly
crosslinked thermosets. A correct model for yielding should account for
strain softening in a self-consistent manner. What is known about polymer
glasses during strain softening is that the volume decreases by a small
amount (Spitzig and Richmond, 1979). extended chain conformations be-
come more prevalent (Magonov, Vainilovitch and Sheiko, 1991) and that
internal energy increases to a steady state value (Oleynik, 1989, 1991; Hasan
and Boyce, 1993). It is instructive to summarize those variables that reduce
the amount of strain softening in a polymer glass sufficiently below ~:
(1) quench cooling, (2) mechanical cycling and (3) hydrostatic pressure P.
The Bulatov-Argon model described in section 4.6.3 can account for the
first two effects in terms of elastic misfit stresses generated by local plastic
events with strain amplitude Yo '" 2. If, on the other hand, one accepts the
molecular mechanics result that elementary transform strain Yo is of the
order of 10- 2 , some structural change such as free volume increase is
210 Yield processes in glassy polymers

required to account for strain softening beyond yield. Pressure effects remain
ambiguous in detail, although it is possible that careful experiments in the
relatively neglected area of high pressure deformation will reveal fundamen-
tal information about the yield process.

REFERENCES
Aboulfaraj, M., G'Sell, c., Mangelinck, D. and McKenna, G.B. (1994) J. Non-Cryst.
Solids, 172-174, 615-21.
Argon, A.S. (1973) Phil. Mag., 28, 839-65.
Argon, A.S. (1993) in Materials Science and Technology, Vol. 6 (ed. H. Mughrabi),
VCH, New York, pp.461-508.
Argon, A.S. and Bessonov, M.I. (1977a) Polym. Eng. Sci., 17,174-82.
Argon, A.S. and Bessonov, M.I. (1977b) Phil. Mag., 35, 917-33.
Argon, A.S., Bulatov, V.V., Mott, P.H. and Suter, U.W. (1995) J. Rheol., 39,377-99.
Bauwens-Crowet, c., Bauwens, J.-c. and Homes, G.A. (1972) J. Mater. Sci., 7,
176-83.
Bowden, P.B. (1973) in The Physics of Glassy Polymers (ed. R.N. Haward), John
Wiley, New York, Ch. 5.
Bowden, P.B. and Jukes, J.A. (1972) J. Mater. Sci., 7,52-63.
Bowden, P.B. and Raha, S. (1970) Phil. Mag., 22, 463-82.
Bowden, P.B. and Raha, S. (l974) Phil. Mag., 29, 149-66.
Boyce, M.e., Arruda, E.M. and Jayachandran, R. (1994) Polym. Eng. Sci., 34,
716-25.
Brady, T.E. and Yeh, G.S.Y. (1971) J. Appl. Phys., 42, 4622-30.
Brown, D. and Clarke, J.H.R. (1991) Macromolecules, 24, 2075-82.
Brown, N. (l983) J. Mater. Sci., 18,2241-54.
Brown, N. (1986) in Failure of Plastics (eds W. Brostow and R.D. Corneliussen),
Hanser, New York, Ch. 6.
Buchdahl, R. (1958) J. Polym. Sci., 28, 239-42.
Bulatov, V.V. and Argon, A.S. (1994) Modelling Simul. Mater. Sci. Eng., 2, 203-22.
Crist, B. (1993) in Materials Science and Technology, Vol. 12 (ed. E.L. Thomas),
VCH, New York, Ch. 10.
Dettenmaier, M., Maconnachie, A., Higgins, J.S. et al. (1986) Macromolecules, 19,
773-8.
Diaz-Calleja, R., Ribes-Greus, A. and Gomez-Ribelles, J.L. (1989) Polymer, 30,
1433-8.
Escaig, B. (1984) Polym. Sci. Eng., 24, 737-49.
Fan, C.F. (1995) Macromolecules, 28, 5215-24.
Gilmour, I.W., Trainor, A. and Haward, R.N. (1979) J. Appl. Polym. Sci., 23,
3129-38.
Gloaguen, J.M., Escaig, B. and Lefebvre, J.M. (1995) J. Mater. Sci., 30,1111-16.
Grenet, J. and G'Sell, e. (1990) Polymer, 31, 205-65.
G'Sell, C. (1986) in Strength of Metals and Alloys, Vol. 3 (eds HJ. McQueen, J.P.
Bailon, J.I. Dickson et al.), Pergamon Press, New York, pp.1943-82.
G'Sell, e. and Gopez, AJ. (1985) J. Mater. Sci., 20, 3462-78.
G'Sell, e. and Jonas, J.J. (1979) J. Mater. Sci., 16, 1956-74.
G'Sell, e. and Jonas, J.J. (1981) J. Mater. Sci., 14, 583-91.
G'Sell, c., Boni, S. and Shrivastava, S. (1983) J. Mater. Sci., 18, 903-18.
G'Sell, c., Hiver, J.M., Dahoun, A. and Souahi, A. (1992) J. Mater. Sci., 27, 5031-9.
References 211

G'Sell, C, El Bari, H., Perez, 1. et al. (1989) Mater. Sci. Eng., ABO, 223-9.
Hasan, O.A. and Boyce, M.C (1993) Polymer, 34, 5085-92.
Hasan, O.A. and Boyce, M.C (1995) Polym. Sci. Eng., 35, 331-44.
Hasan, O.A., Boyce, M.C, Li, X.S. and Berko, S. (1993) J. Polym. Sci. B: Polym.
Phys., 31, 185-97.
Haussy, 1., Cavrot, J.P., Escaig, B. and Lefebvre, 1.M. (1980) J. Polym. Sci. B: Polym.
Phys. Edn., 18, 311-25.
Haward, R.N. (1973) in Physics of Glassy Polymers (ed. RN. Haward), John Wiley,
New York, Ch. 1.
Haward, RN. (1974) in Molecular Behaviour and the Development of Polymeric
Materials (eds A. Ledwith and A.M. North), Chapman & Hall, London, p. 444.
Haward, R.N. (1994) Thermochim. Acta, 247, 87-109.
Hutnik, M., Argon, A.S. and Suter, U.w. (1993) Macromolecules, 26,1097-108.
Imai, Y. and Brown, N. (1976) J. Polym. Sci., Polym. Phys. Edn., 14, 723-39.
Kelly, A. and Macmillen, N.H. (1986) Strong Solids, 3rd edn, Oxford University
Press, New York, 423 pp.
Knauss, W.A. and Emri, I. (1987) Polym. Eng. Sci., 27, 86-100.
Kozey, v.v. and Kumar, S. (1994) J. Mater. Res., 9,2717-26.
Krausz, A.S. and Eyring, H. (1975) Deformation Kinetics, Wiley-Interscience, New
York, 398 pp.
Lefebvre, 1.M., Escaig, B., Coulon, G. and Picot, C (1985) Polymer, 26,1807-13.
Lohse, F., Schmid, R, Batzer, H. and Fisch, W. (1969) Br. Polym. J., 1, 110-14.
Magonov, S.N., Vainilovitch, I.S. and Sheiko, S.S. (1991) Polym. Bull., 25, 491-8,
499-506.
Matsuoka, S. (1986) in Failure of Plastics (eds W. Brostow and RD. Corneliussen),
Hanser, New York, Ch. 3.
Matsuoka, S. (1992) Relaxation Phenomena in Polymers, Hanser, New York, Ch. 3.
Matsushige, K., Radcliffe, S.V. and Baer, E. (1976) J. Polym. Sci., Polym. Phys. Edn,
14, 703-21.
Mott, P.H., Argon, A.S. and Suter, U.W. (1993) Phil. Mag. A, 67, 931-78.
Nanzai, Y. (1993) Prog. Polym. Sci., 18,437-79.
Oleinik, E.F. (1986) Adv. Polym. Sci., 80, 49-99.
Oleynik, E. (1989) Prog. Coli. Polym. Sci., 80, 140-50.
Oleynik, E.F. (1991) in High Performance Polymers (eds E. Baer and A. Moet),
Hanser, New York, pp.79-102.
Pae, K.D. and Bhateja, S.K. (1975) J. Macromol. Sci., Rev. Macromol. Chern., C13,
1-75.
Powers, J.M. and Caddell, RM. (1972) Polym. Eng. Sci., 12, 432-6.
Raghava, R, Caddell, R.M. and Yeh, G.S.Y. t1973) J. Mater. Sci., 8, 225-32.
Rendell, R.W., Ngai, K.L., Fong, G.R. and Bankert, R.J. (1987) Polym. Eng. Sci., 27,
2-15.
Robertson, RE. (1966) J. Chern. Phys., 44, 3950-6.
Ruan, M.Y., Moaddel, H., Jamieson, A.M. et al. (1992) Macromolecules, 25, 2407 -11.
Spitzig, W.A. and Richmond, O. (1979) Polym. Eng. Sci., 19, 1129-38.
Stokes, V.K. and Bushko, W.C (1995) Polym. Eng. Sci., 35, 291-303.
Struik, L.CE. (1991) J. Non-Cryst. Solids, 131-133, 395-407.
Tant, 1.M. and Wilkes, G.L. (1981) Polym. Eng. Sci., 21,874-95.
Theodorou, D.N. and Suter, U.W. (1986) Macromolecules, 19, 139-54.
Theodorou, M., Jasse, B. and Monnerie, L. (1985) J. Polym. Sci., Polym. Phys. Edn,
23,445-50.
Walker, N. (1980) Polymer, 21, 857-9.
Ward, I.M. and Hadley, D.W. (1993) An Introduction to the Mechanical Properties
of Solid Polymers, John Wiley, New York, Ch. 11.
212 Yield processes in glassy polymers

Weertman, land Weertman, lR. (1992) Introduction to Dislocations, 2nd edn,


Oxford University Press, New York, 213 pp.
Wignall, G.D. (1987) in Encyclopedia of Polymer Science and Engineering, Vol. 10
(eds. H.F. Mark et al.), John Wiley, New York, pp.112-84.
Williams, M.L., Landel, R.F. and Ferry, J.D. (1955) J. Am. Chem. Soc., 77, 3701-7.
Wu, J.B.e. and Li, lC.M. (1976) J. Mater. Sci., 11,434-44.
Wu, W. and Turner, P.L. (1975) J. Polym. Sci., Polym. Phys. Edn, 13, 19-34.
Xie, L., Gidley, D.W., Hristov, H.A. and Yee, A.F. (1995) J. Polym. Sci. B: Polym.
Phys., 33, 77-84.
Yamini, S. and Young, R.J. (1980) J. Mater. Sci., 15, 1814-22.
The post-yield

5
deformation of glassy
polymers
M.e. Boyce and R.N. Haward

5.1 GENERAL FEATURES OF POST-YIELD DEFORMATION

5.1.1 INTRODUCTION
The science and technology associated with the mechanical properties of
thermoplastics still harbour many mysteries. There are numerous different
aspects of mechanical properties and their evaluation entails many types of
measurement, such as short term modulus and creep, yield behaviour and
large deformation after yield, toughness and fracture. In various ways all
these measurements are related to each other, but until about 1970 the
importance of large strains in determining other mechanical properties,
often measured separately, was not generally appreciated. Since that time
much progress has been made, though there are still many uncertainties
concerning the extent and manner in which different mechanical properties
are connected. It is therefore appropriate, in a review of large deformations
in glassy polymers, to begin by summarizing briefly the way in which ideas
on the interrelations between mechanical properties have developed and
why it is that large strains are of particular interest. This is also of
importance in efforts to understand the differences responsible for tough and
brittle fracture.
One early proposal suggested that polymers were brittle if the 'yield
strength' was above the 'brittle strength' so that such materials broke before
any significant energy absorbing deformations could take place. Although
this explanation was formally indisputable, it did little beyond putting into
words the properties which could easily be seen and measured. It provided
no mechanism and its fundamental inadequacy was eventually exposed by
developments in the study of crazing and brittle fracture as described in
Chapter 6. It was shown that this type of fracture was itself the product of

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
214 The post-yield deformation of glassy polymers

Figure 5.1 Melting during the fracture of polystyrene. Replica electron micrograph
of the fracture surface after a fast brittle fracture . (Reproduced from Haward and
Brough, 1969, with permission.)

large deformations which were localized within very small volumes of


material (Figure 6.1). Because of the small scale on which deformation took
place, the total amount of energy absorbed in relation to the whole test piece
or moulding was small. On the other hand, in relation to the amount of
material taking part in the deformation process, the energy absorbed was
sufficient to raise the temperature above its softening temperature as
illustrated in Figure 5.1.
Thus even the brittle fracture mechanism is inseparably associated with
large deformations. More recently this process, generally known as crazing,
which is the subject of Chapter 6, has been the subject of an intensive
research programme carried out by Kramer and coworkers (Donald and
Kramer, 1982; Kramer and Berger, 1990) who have shown that the proper-
ties of the material that takes part in the formation of the craze accords with
much that is known about large deformation behaviour.
Our present understanding of crazing and fracture requires that brittle
fracture processes are controlled by the absorption of energy at the crack
General features of post-yield deformation 215

tip and that this occurs in a relatively small volume of material where large
deformations take place. In the case of tough fracture a much larger volume
of material is affected, and large strains are generally easy to observe. In
either case the fracture process is controlled by the large deformation
process and especially by the way in which the volume of material contri-
buting to the energy sink is determined. .
Large deformation behaviour is also of importance in solid phase-forming
processes which may take place below, in or above the glass transition
temperature of the material. Processes such as cold drawing, stretching,
vacuum forming and blow moulding rely on the molecular orientation that
develops with large strain deformation. In additon to forming processes, the
large strain behaviour is also exploited in applications which require energy
absorption, such as the use of polymer shields to absorb impact through
their large strain deformation. Therefore, the study of post-yield deforma-
tion is not only a key for understanding brittle and ductile failure processes,
but also large scale forming processes and impact loading applications.

5.1.2 THE REQUIREMENT FOR A MINIMUM MOLECULAR WEIGHT


The amount of deformation which takes place in a craze, typically some
micrometers thick, does not itself contribute significantly to the overall
observable extension when a glassy polymer such as polystyrene is extended
until it breaks. However, the energy required to form and propagate the
craze provides a resistance to crack growth and gives the plastic its tensile
strength. A great deal of information on this subject is now available and
will be the subject of Chapters 6 and 7. Here we are concerned with the way
in which molecular weight affects other measurements that depend upon
large deformations located in a craze.
In Figure 5.2 the elongation at break is plotted for polystyrenes having
different types of molecular weight distribution and a range of molecular
weights (McCormick, Brower and Kim, 1959). It will be seen that at high
molecular weights the elongation rises to a maximum and levels off, whereas
at low molecular weights it falls effectively to zero. The major part of the
elongation observed in these experiments is of the reversible Hookean type
and depends on the ratio of the tensile strength to the Young's modulus.
For glassy polymers, well below their glass transition temperature the latter
is nearly always between 2 and 4 GPa. For those glassy polymers which do
not give bulk deformation in tension, differences in elongation therefore
reflect the differences in craze strength and stability in line with results
reported in the literature (Brady and Yeh, 1973; Donald and Kramer, 1982).
This decline in strength and elongation almost to zero occurs when the
molecular weight falls below what is called the entanglement molecular
216 The post-yield deformation of glassy polymers

2.-------------------------------------~

o 05 10

Figure 5.2 The effect of molecular weight on the elongation at break in tension of
polystyrene. The elongation is controlled by the tensile strength of a craze, which in
turn depends on a high level of entanglement. So the elongation of the narrow
molecular weight distribution anionic polymers • extrapolates to zero at the
entanglement molecular weight Me ~ 35000. The points marked. are for wide
distribution thermal polymers. (Reproduced from McCormick, Brower and Kim,
1959, with permission.)

weight Me' which can be determined from measurements of melt viscosity.


When derived from the plateau modulus a smaller quantity Me is obtained
where Me/Me'" 2 (Porter and Johnson, 1966; Ferry, 1980; Aharoni, 1983,
1986). Since it is known that even high molecular weight polystyrene
molecules do not span a craze (Haward, Daniels and Treloar, 1978), the
strength of a craze must depend on entanglements between polymer mol-
ecules. Below this molecular weight, generally found to be about 35000 for
polystyrene, a stable craze is not formed so that the tensile strength and
elongation then drop to zero on the scale used in Figure 5.2.
In the past it was assumed that, at these low molecular weights, large
deformations did not occur but apparently this is not the case. Applying the
methods of linear elastic fracture mechanics (Chapter 7), Kramer (1978) has
used measurements on low molecular weight polystyrene fractions (Robert-
son, 1976) to show that below Me the fracture energy follows a square root
dependence on molecular weight, and in all cases exceeds the Griffith (1920)
condition for brittle fracture where surface energy alone determines strength.
Applying the known dimension of polystyrene molecules in the glassy state,
Kramer (1978) was able to show that their probable end-to-end displace-
ments were such that the material which was deformed in the fracture
process could just be accomodated in a postulated 'primordial (isotropic)
craze'. When the matrix material was stretched out at the 'craze' tip the
molecules would then be stretched to their maximum length with their
extremities embedded in undeformed material. So even under these un-
favourable conditions, the strength of the glassy polymer was controlled by
General features of post-yield deformation 217

large deformations at the crack tip. Naturally the stresses involved in the
fracture of such low molecular weight materials are much smaller than for
high polymers, and the related large deformations cannot be observed by
conventional methods. The ensuing work described in this chapter is
therefore entirely concerned with materials having a molecular weight above
Me·
These arguments have been presented to underline the universality of
large deformations in thermoplastics including the glassy polymers. In the
past the argument has been put forward that in the glassy state plastics do
not show large deformations. If, on the contrary, as shown by the accepted
work on fracture, large deformation occurs at the crack tip with all these
materials, whenever the stress rises to a sufficient level for crack propaga-
tion, then this proposition cannot be sustained. By the same argument the
study of large deformations becomes of increasing importance.

5.1.3 CHARACTERISTIC FEATURES OF POST-YIELD DEFORMATION IN


GLASSY POLYMERS

A characteristic feature of large deformations in solid linear polymers,


including the glassy thermoplastics is that, at the end of the test, when the
stress is removed, the major part of the plastic deformation remains. Because
of this apparent stability of form, the process is sometimes described as
one of flow. However, if the temperature is raised above the glass transition
temperature, this deformation is reversed and the material returns to its
original shape. This effect has been known for a long time. It appears to have
been first reported by Gurevich and Kobeko (1940) for PVC and later by
Haward (1942) for cellulose esters and by Hoff (1952) with poly(methyl
methacrylate) (PMMA). Hoff described a remarkable case where hollow
cylinders were grossly distorted and staged an almost perfect recovery when
heated up. An example of this behaviour is shown in Figure 5.3 for a short
bar of polystyrene in which a large shear band had been formed under
compression. On heating above Tg it reverted very precisely to its original
shape (Murphy, Haward and White, 1971). Clearly, these polymers which
had apparently been permanently deformed had remembered their previous
shape and behaved as though they comprised a frozen elastic extension
rather than true flow.
Some of the more significant aspects of large deformation processes will
now be summarized .

• When a glassy polymer is subjected to an increasing stress, the consequent


strains are at first small, reversible, proportional to the stress and
independent of the time. If the stress continues to increase then various
218 The post-yield deformation of glassy polymers

Figure 5.3 The reversibility of large deformations in glassy polymers. (a) A wide
shear band formed in polystyrene under compression . (b) The same test piece after
heating for 30 min at 110°C (Tg ~ 100°C).

time-dependent small deformations are observed, followed by yielding as


described in Chapter 4. Under tension this generally leads to a maximum
in a conventional stress-strain curve.
• After yielding, only a small minority of glassy polymers deform in a
uniform manner under tension. More often they exhibit various types of
plastic instability such as shear bands (Bowden, 1973), crazing and
necking. The latter is accompanied by a fall in the nominal or engineering
stress.
• When the true stress (Chapter 4 and section 5.2.5) is measured or
estimated immediately after yield, a fall in stress is also often observed.
Such decreases in true stress are known as true strain softening, and have
been recorded under a wide range of experimental conditions: in tension,
by G'Sell (1982) for polycarbonate; in compression by Binder and Muller
(1961) and Hasan and Boyce (1993) for several materials; and in shear by
Sternstein, Ongchin and Silverman (1968). True strain softening is strong-
ly dependent of the thermal history of the material as shown by Mills
(1975) and Struik (1978) and, later, by Hasan and Boyce (1993); Hasan
et al. (1993).
• After yield, large deformations take place under the action of large or
moderate stresses. This process absorbs a great deal of energy, part of
General features of post-yield deformation 219

which is converted into heat as described in section 5.4. If this heat is not
lost to the environment, then the temperature of the plastic is raised and
thermal softening takes place. This generally increases the tendency of
plastic deformation to be localized in particular parts of the test piece
(plastic instability) and in some cases a thermal runaway can lead to
fracture (Figures 5.1 and 5.24) .
• As deformation increases further under isothermal conditions, strain
hardening is observed. This property is very sensitive to the state of
straining (i.e. biaxial or uniaxial), the prior orientation of the polymer
molecules, and to the molecular structure .
• If a tensile test is continued indefinitely, a point will be reached where the
test piece breaks. Although the process of fracture at high strains is not
well understood, it does not seem to follow a crazing mechanism; instead
characteristic 'rhomboidal' or 'diamond' cavities are formed which grow
larger until catastrophic fracture occurs (Regel, 1951; Comes and Ha-
ward, 1974; Kinloch and Young, 1983).

5.1.4 PLASTIC INSTABILITY PHENOMENA

The most well known phenomenon associated with plastic instability is


that of 'necking' in a tensile test. This term is used to describe the following
type of behaviour. When a dumbbell-shaped test piece is extended at a
constant speed, it does not deform to the same extent through the whole
parallel length. Instead, the material starts to extend at one point to form a
neck which then grows until it includes the whole length of the test piece.
During this process, the force applied to the test piece to propagate the neck
remains approximately constant and the same applies to the strain in the
necked material, generally referred to as the 'natural draw ratio'. Finally,
when the neck has extended through the parallel length of the test piece, the
stress will start to rise as strain hardening is recorded, until fracture occurs.
This process is illustrated in Figure 4.7. It will be understood that, except at
the beginning and the end of the test, the stresses and strains recorded are
average figures and do not correspond to the conditions in any particular
part of the test piece.
In the early literature there are numerous references to necking with
different glassy polymers, of which the following may be mentioned: with
poly(vinyl chloride) (PVC), Vincent (1960) and Robertson (1963); with
polycarbonate, Robertson (1963); with poly(methyl methacrylate) under
appropriate conditions, Lazurkin and Fogel'son (1951) and Rehage and
Goldbach (1967); and finally necking with poly(ethylene terephthalate) was
reported by Brown and Ward (1968). Polystyrene does not easily exhibit
necking, since crazing and brittle fracture are the normal responses to an
220 The post-yield deformation of glassy polymers

Figure 5.4 A neck formed in a cylindrical polycarbonate test piece. The small
inscribed bench marks enable true local strains to be observed . (Reproduced from
Nazarenko et ai., 1994, with permission .)

applied tension; but with oriented polystyrene, Ender and Andrews (1965)
observed necking at elevated temperatures and, under the influence of
hydrostatic pressure, yielding and necking were reported by Holliday et al.
(1964) and by Biglione, Baer and Radcliffe (1969). Clearly this list covers
most of the important types of glassy polymer. However, it may also be
noted that necking is equally pervasive with semicrystalline materials, e.g. it
has been observed with high density polyethylene by Vincent (1960) and in
nylon by Bender and Williams (1963). More recently, Gent and Madan
(1989) studied a number of polymers and concluded that necking was the
characteristic behaviour of semicrystalline thermoplastics, though they also
found that uniform extension can sometimes be observed with certain
materials at low strain rates. With amorphous materials, cellulose esters
(Haward, 1942) and some polyimides (Bessonov et ai., 1987) have been
shown to extend uniformly in a tensile test, while other polymers can be
persuaded to do so and will be discussed later. An example of necking in
tension is given in Figure 5.4.
General features of post-yield deformation 221

5.1.5 MECHANICS OF NECKING IN TENSION


The basic theory of neck formation is relatively straightforward and is
generally attributed to Considere (1888) e.g. by Nadai (1950) and Vincent
(1960), whose treatment is followed here. When a tensile test piece deforms
plastically with minimal changes in volume, then the cross-section areas A,
Ao and lengths 1,10 (where subscript 0 refers to the initial condition of the
test piece), the extension ratio ), and the plastic strain e are related as
follows:

However, if a test piece is considered to extend uniformly at a constant rate


of extension, the strain rate, related to the actual length, decreases continu-
ously. It is therefore convenient to define the true strain as In A and true
strain rates as din A/dt = (1/),) dA/dt.
It follows from the relations given above that if O"n is the nominal or
engineering stress, referred to the original cross-section of the test piece, then
the true stress 0" = O"n· )., from which it can be shown that

Then, at the point where O"n is a maximum, dO"n/dA = 0 and dO"/dA = O"/A.
This corresponds to the tangent to the true stress-strain curve from an
origin defined by A = 1, when plotted as described in Figure 5.5. Thus the
maximum of O"n and the tangent to the true stress-strain curve occur at the
same elongation. When the material has properties such that this condition
is complied with, any further increase in elongation beyond this point leads
to a fall in the nominal stress 0" n.
If the test piece were perfectly uniform in cross-section and composition
it would, in principle, be possible for uniform extension always to take place.
However, in practice this is never the case; there is always a point in the test
piece where O"n reaches the maximum first, and when this happens, there is
a decrease in the stress required to extend further at this point. Extension
therefore continues there while the stress on other parts of the test piece falls
below that required to exceed the yield point.
In discussing these phenomena, Vincent (1960) distinguished three types
of tensile behaviour based on the 0", A relationship.

(1) dO"/d), > O"/A

in this case stable elongation takes place. This occurs with vulcanized
rubbers and thermoplastic cellulose esters.

(2) dO"/dA < O"/A


222 The post-yield deformation of glassy polymers

150

ra 100
0...
::E
<Ii
(J)

~
Vi
Q)

~ 50

o 2 4 6 8 10
Draw Ratio
Figure 5.5 The Considere construction for high density polyethylene. The tangent
near A predicts that necking will take place and the second one at B that the necking
will be stable. (Reproduced from Vincent, 1960, with permission.)

120

100

G p =2.5 Y =25 MPa


80
<U
0...
:E
lZ 60
~
<i5
Q)

~ 40

20

0
0 2 3 4 5
Draw Ratio

Figure 5.6 A calculated stress- strain curve for high density polyethylene. The points
are derived indirectly from G'Sell's experiments using the Gaussian equation (G'Sell
and Jonas, 1981; Haward, 1995). All polymers following the Gaussian or Langevin
relations which show the first tangent will also show the second.
Developments in the measurement of true stresses and strains 223

at one point. This leads to necking with thinning to failure. Finally


(3) da/dA = a/A
for a second time and an starts to increase. This provides a stable neck, as
observed for linear polythene, polycarbonate and many other plastics. The
neck then extends through the test piece, and if a ductile fracture process
does not intervene, it will reach the ends of the test piece after which the
stress generally rises until fracture occurs. Vincent (1960) demonstrated the
third type of behaviour using experimental results with linear polythene as
shown in Figure 5.5. In Figure 5.6, calculated results are given for the same
type of material using methods to be described in section 5.6 (Haward,
1995).

5.2 DEVELOPMENTS IN THE MEASUREMENT OF TRUE


STRESSES AND STRAINS

5.2.1 USE OF BENCH MARKS


In order to derive a true stress-strain relation, it is necessary to measure the
strain, strain rate and stress in a particular section of the test piece within
which they are constant. As already noted, these quantities are not directly
provided in a conventional tensile test, even when very low rates of strain
are employed to minimize heating effects. In order to obtain the required
results in tension it is therefore necessary to make measurements over a
length of material where constant conditions may be assumed. Develop-
ments on these lines will now be described.
The test piece used in a conventional test may be inscribed with bench
marks and the local extensions followed photographically. Rather fine bench
marks are visible on Figure 5.4. Naturally, the closer the bench marks are
together, the more accurate are the results. This method was used by Meinel
and Peterlin (1971a,b) using different types of polyethylene, and probably
also by Vincent (1974) who worked with PVc. Meinel and Peterlin (1971a)
pointed out that there was some loss of accuracy in the measurement of
strain in the vicinity of the neck, where there were also local increases
of strain rate. However, the effect of the latter may have been mitigated by
thermal effects acting in the opposite sense. Curves of true stress against
linear extension were plotted at different temperatures.

5.2.2 UNIFORM DEFORMAnON


A number of polymers deform uniformly in a tensile test. These include the
cellulose esters (Haward, 1973) and some polyimides e.g. polyimide DPO
(formula given on Figure 5.28), as described by Bessonov et al. (1987).
224 The post-yield deformation of glassy polymers

Other polymers will extend uniformly at low strain rates under suitable
conditions. These included ultra-high molecular weight polyethylene
(Meinel and Peterlin, 1971b; Trainor, Haward and Hay, 1977) and quen-
ched PVC (Cross and Haward, 1978). Results reported with these materials
generally employed a constant rate of extension and measured linear strains,
but with uniform extension these could readily be converted to the logarith-
mic form. There will be small variations due to the change in strain rate as
the length of the test piece increases, unless a correction is introduced.
However, if necking occurs during tension, there is a rapid increase in local
strain rate at the initiation of the necking instability.
The method of directly measuring uniform deformation in tension is
limited to materials which are capable of this type of behaviour. However,
it has been found possible to extend the application of this technique by
working with materials which have already passed through the neck. When
a conventional tensile test is carried out, material will emerge from the
necking stage when it has reached the 'natural draw ratio' under the relevant
conditions. If attempts are made to extend it further, the plastic may break
first, but sometimes the elongation at break greatly exceeds the 'natural
draw ratio'. In that case it is possible to take a length of necked material,
previously inscribed with bench marks, and to extend it further under
conditions of uniform deformation. When the remaining extensibility is
larger than or comparable with the strain during necking, this procedure can
be used to measure a significant part of the true stress-strain curve. This
method has been used for low density polyethylene by Mills, Hay and
Haward (1985) and by Biddlestone et al. (1986), for poly(ether ether ketone)
(PEEK) in both the amorphous glassy and the semicrystalline forms
(section 5.6.2).

5.2.3 MEASUREMENTS BASED ON LOCAL DIAMETER OR THICKNESS


If the cross-sectional area of a test piece can be measured continuously with
a good degree of accuracy at a particular point, then the valu of A. can be
calculated from it. With a cylindrical specimen, it is sufficient to measure the
diameter D so that InA. becomes 2 InDo/D. This type of measurement
generally employs an 'hourglass' shaped test piece with a constricted waist
which does not allow the formation of a neck. An outline of the hourglass
test piece is included in Figure 5.7. This type of test piece was introduced by
G'Sell and Jonas (1979, 1981) and by Hope, Ward and Gibson (1980). The
latter used a test piece made from 1 mm PMMA sheet which was cut in
an hourglass form to give a 5 mm waist. After verifying that lateral
and transverse contractions were proportional, they arranged to measure
the transverse contraction in the waist using an accurate linear voltage
Developments in the measurement of true stresses and strains 225

Analogue-digital
load cell interface

ITTIB
Display monitors

Curve
plotter

Tensile testing Digital-analogue


machine ramp generator

Figure 5.7 General diagram of the tensile testing equipment used by G'Seil et al.
(1992). An hourglass test piece is used and extension is applied via a computer aided
video extensometer which maintains a constant true strain rate.

displacement transducer. Since PMMA does not give large deformations


under tension at normal temperatures, they used a temperature of 90°C and
also employed variable strain rates. Later a similar method was used by
Boyce and Arruda (1990) for polycarbonate at room temperature with a
cylindrical hourglass test piece. G'Sell and Jonas (1979, 1981) also used a
cylindrical test piece machined from rod and extended a series of different
polymers at room temperature. A flexible, thin steel cable was passed round
the waist of the test piece and attached to the core of a linear transducer
adjusted to give a voltage proportional to the diameter of the sample. The
voltage was fed back to a closed-loop tensile machine which delivered a
constant true strain rate at the point of measurement. Thus this equipment
was designed to meet all the requirements for true stress--strain measure-
ments. In more recent work G'Sell et al. (1992) introduced further improve-
ments in their experimental method. Their new procedure was similar in
principle to that used in earlier work, but it employed a computer-aided
video extenso meter to record the strain in the mid-plane of the hourglass
(Figure 5.7). In calculating the effective stress, a triaxiality correction factor
was also introduced corresponding to the local radius of curvature in the
sample profile.
226 The post-yield deformation of glassy polymers

5.2.4 DIFFERENT TYPES OF MECHANICAL TEST


Following developments in the measurements of true strains and stresses in
tension, there has been increased interest in other stress fields where different
problems are encountered. In compression a simple rectangular test piece
often first yields as a series of thin shear bands (Bowden, 1973; Chapter 4),
although broader shear bands can also be formed as in Figure 5.3. Very
often cylindrical test pieces are employed, as by Binder and Muller (1961)
and these are very susceptible to a phenomenon known as barrelling in
which the centre of the cylinder expands to form a barrel-like shape. This is
another type of non-uniform deformation. Recent work has minimized this
effect by the use of improved lubricants at the surfaces under compression
(Boyce and Arruda, 1990), where a uniform state of uniaxial compression is
obtained to very large strains (log strains greater than -1.2). In the work
of Boyce and Arruda, constant strain rate conditions are also maintained
throughout the deformation. A plane strain compression test has also been
employed in which the sample is constrained to expand in only one
dimension (Bowden (1973). This was applied to quenched PVC by Cross
and Haward (1978), who used the technique described by Williams (1967).
More recently Arruda and Boyce (1993a) have introduced an improved
plane strain compression device which more accurately constrains the
deformation of the test piece so that a globally uniform state of plane strain
compression is maintained throughout the specimen to very large strains.
G'Sell and Gopez (1985) have reported measurements in simple shear in
which it was difficult to produce homogeneous simple shear throughout the
gauge section, as deformation easily became localized in bands along the
length of the test piece. However G'Sell and Gopez (1985; Figure 4.10) claim
to have produced a state of uniform shear in the later stage of their tests
but, at the very largest strains, draw-in of material from the grip at the ends
of the specimen and edge cracking (Boyce, Arruda and Jayachandran, 1994)
interferes with the accuracy of the stress-strain measurement.

5.2.5 TRUE STRESS-STRAIN CURVES


As a result of experiments carried out using the methods described in the
preceding section, a considerable number of true stress-strain curves have
now been published both for glassy and crystalline polymers. Although they
differ, both in accuracy and whether or not a constant rate of extension or
a constant true strain rate was employed, the results are remarkable more
for their similarities than their differences. They will be illustrated here by
the tension results of G'Sell (1982), using a constant true strain rate and by
Developments in the measurement of true stresses and strains 227

til
a.
;;.
oo
oo
w
a::
l-
(/)
w 100
::J
a::
I-

Figure 5.8 True stress-strain curves as reported by G'Sell (1982). PA 6 and 66,
polyamides; PC, polycarbonate; PVC, poly(vinyl chloride); PP, polypropylene; HDPE
and LDPE, high and low density polyethylene respectively; and PTFE, polytetra-
fl uoroethylene.

those of Hope, Ward and Gibson (1980) using a constant rate of extension,
shown in Figures 5.8 and 5.9 respectively.
Examination of the curves shows that they comprise of two and
sometimes three stages. At the beginning there is a steep rise in stress,
accompanied by a small strain. Although, when studied in detail, this part
of the curve shows a number of complex effects, it can conveniently be
represented by a reversible Young's modulus in the present context. As the
stress level is raised, yielding occurs and continues until the curve, either
immediately or ultimately, feeds into a prolonged final stage where it sweeps
upwards with a continuously increasing slope. This feature, whereby the
stress increases with the strain at an accelerating rate, is known as strain
hardening. However, with high molecular weight PMMA and with polycar-
bonate (PC), at the beginning of the large strain process, before strain
228 The post-yield deformation of glassy polymers

100 100
/ I
/ I
(a) (b) I I .
/ I I
80 80 I I I
/ I /
(J / (J / I /
/ / I
/
(MPa) (MPa)
, / /
60 60 i'l,. /
~ /~ec../ . . .
. " /? ~
4.1/ ~<",e;-
/ /.s;, ~e~
40
/ / *'.. i-
/

/
/ ;'>

/~
,

'"
/

, ___ - ............ '" / ~


/1
I -- /

/..---~----
....... '"

£ £

Figure 5.9 True stress-strain curves for PMMA. (a) High molecular weight polymer,
(b) low molecular weight. Deformation was measured by lateral contraction of a
waisted specimen. (Reproduced from Hope, Ward and Gibson, 1980, with per-
mission.)

hardening sets in, there is a marked maximum followed by a fall in true


stress. With PVC there is just a small 'blip' at this point. This drop in stress
with increase in strain in a true stress-strain curve is called true strain
softening. After passing through the maximum the stress appears to return
to a form similar to that of the other materials that do not show true strain
softening. Note that strain softening is also observed in compression, and
therefore is not a tensile instability phenomenon. Finally, it is also charac-
teristic that the final rise in true stress takes place at widely differing strains
and stresses for different polymers.
There is very little doubt that the general shape of the tensile curves
illustrated above is truly representative of the true stress-strain response of
thermoplastics. All the curves reported by the several workers with different
experimental techniques show similar strain hardening characteristics. This
applies even to strain softening, though, as will be described later, this is
strongly affected by the sample's thermal history. For example G'Sell et al.
(1992), using their improved technique with PMMA, observed a marked
peak and obtained a curve similar to those in Figure 5.9. Similarly Boyce
and Arruda (1990) found a peak with PC similar to that in Figure 5.8 and
Hasan and Boyce (1993) have observed this behaviour in compression in
polystyrene, PMMA and Pc.
Developments in the measurement of true stresses and strains 229

While the overall nature of the true stress versus strain curve has been
observed for a variety of polymers in various tests as discussed above, it is
important to recognize that the true stress-strain curve for a polymer has
a significant dependence on the state of strain tested. For example, Arruda
and Boyce (1993a) have illustrated this difference for both PC and PMMA
by testing in uniaxial compression and plane strain compression. In both
sets of tests, they obtain globally homogeneous deformation. These results
for PC are shown in Figure 4.9. Note the apparent increase in yield stress
in plane strain over that in uniaxial compression - this is a result of both
the plane strain constraint and also the pressure dependence of yield. As
reported in Chapter 4, this is a general property of glassy polymers. The
results for PC also illustrate the dramatic difference in the strain hardening
behaviour, where the material is observed to harden more rapidly with
strain under plane stain compression conditions than under uniaxial com-
pression conditions. Arruda and Boyce attribute this difference to be a direct
result of the different state of molecular orientation obtained in these
different states of strain and have verified this using birefringence measure-
ments. They have successfully modelled this behaviour, as will be discussed
in a later section. Cross and Haward (1978) also compared tension and
plane strain compression measurements with PVC with similar results
(section 5.6.2), while G'Sell and Gopez (1985) observed this state of strain
dependence in PC when comparing their tension and simple shear stress-
strain curves.

5.2.6 EFFECT OF PREORIENTATION


As distinct from the type of orientation imposed during the straining
process, the true stress versus strain behaviour is also found to be strongly
dependent on any preorientation existing in the polymer. Allison and Ward
(1967), and Long and Ward (1991) illustrated this dependence via uniaxial
drawing of PET fibres where the PET had first been predrawn to various
initial stretches. They found that if this initial stretch was accounted for by
a shift of the stress-strain curve along the strain axis (Figure 5.10), then all
curves superposed. This result indicates the existence of an underlying
molecular network which controls the drawing behaviour where all stretches
are relative to the limiting stretch of the network. The extensions are also
accompanied by an increase in the proportion of trans rotational isomers
(Ward, 1984).
The effect of pre orientation on subsequent mechanical behaviour is a
three-dimensional effect. The three-dimensional nature of preorientation
is shown in the experiments of Arruda, Boyce and Quintus-Bosz (1993).
In this work, PC samples were preoriented by initial uniaxial or plane
230 The post-yield deformation of glassy polymers

O~----~----~ ____ ~ ____ ~ ____- J


00 04 oa 12 16
a I n )..

b In )..

Figure 5.10 True stress- strain curves for spun PET yarns. In (a) the yarns have all
been prestrained by different amounts as shown on the x-axis. and in (b) each curve
has been matched against the lowest orientation precursor yarn by a horizontal shift
to give a value of the network draw ratio. (Reproduced from Long and Ward. 1991.
with permission.)
Developments in the measurement of true stresses and strains 231

z x Cili eJ
,
simulation ----'7'
I' y
.xperim.ntal~

f,
I:,
I'
//
I
f
:
'
Z I
100 I ,
I

50

a ~~~--~~~~~~--~~~--~~~
0.0 0.5 1.0
True Strain
Figure 5.11 The effect of preorientation on the compression of Pc. True stress-
strain results for uniaxial compression of PC comparing the response of initially
isotropic PC (center line) to those of a preoriented PC subjected to compression
along two different axes (,Y' and 'Z' curves). The preorientation was transversely
isotropic, being via an initial 8 = -0.57. Also shown is the ability of the Arruda and
Boyce constitutive model to capture the effects of preorientation on initial yield and
subsequent strain hardening (section 5.7.1)
232 The post-yield deformation of glassy polymers

strain compression. The preoriented samples were then compressed in


different directions. Figure 5.11 shows one example of the stress-strain
curves obtained in the different directions; a reference true stress-strain
curve on an initially isotropic sample is also shown. The material is found
to exhibit a lower yield stress and much delayed hardening when com-
pressed along an axis of pre orientation (here, the network will first 'deorient'
then reorient in a plane normal to the compression direction), whereas a
higher yield and early hardening are observed when compressed normal to
the plane of initial orientation (here the network is continuing to be further
oriented in the same state as the initial orientation, similar to Ward's work
on PET uniaxial drawing). The three-dimensional effect of pre orientation
also has dramatic consequences on the deformation patterns of the cross-
sectional shapes, which are highly anisotropic and dependent on direction
of straining as shown in Arruda, Boyce and Quintus-Bosz (1993). The
Arruda and Boyce results are also consistent with the concept of an
underlying macromolecular network governing post-yield hardening/defor-
mation and point to the need for a three-dimensional representation of this
behaviour. This effect has been successfully modelled by Arruda, Boyce and
Quintus-Bosz (1993); (see also section 5.7.5).

5.3 PHYSICAL AGING AND LARGE DEFORMATIONS

5.3.1 TRUE STRAIN SOFfENING


In this section we continue the account of the physical aging process already
given in Chapter 3, where some of its fundamental features were discussed
and related to the glass-liquid transition and to free volume changes. The
effects on small deformation processes were also described. Here we deal
with the influence of physical aging on the deformation processes which take
place after yielding has commenced.
As already described in section 5.1.5 and Chapter 4, most thermoplastics
exhibit necking at yield in a conventional tensile test, and this is accom-
panied by a fall in the engineering stress from the 'yield stress' to the
'drawing stress'. Much of this fall in stress is only apparent and is geometric
in origin following the Considere principle, but many glassy polymers also
show a fall in true stress at the same time. For this reason, strain softening
in tension can only be clearly demonstrated when true stress-strain curves
are measured, as in Figures 5.8 and 5.9, where this effect is clearly shown for
polycarbonate in Figure 5.8 and for high molecular weight PMMA in
Figure 5.9a. However, these problems do not arise in other types of test,
where strain softening was observed in compression some time ago by
Binder and Muller (1961) for polystyrene, polycarbonate and PVC, and
Physical aging and large deformations 233

r--------r--------------~--~.--~/~~
,... /

cooled from 90 /0 :noC "","" 250 101


01 OS'C/h _ _ _ __
~ 70'·

0: requenched ofler
slow cooling
SOI---L---+"'--~'----------+_--------- .---+

_ extension role, em/min

10" 10'

Figure 5.12 The effect of quenching and annealing on the yieid strength of Pvc.
The samples indicated were quenched from 90 to 50°C and then aged for various
times as shown, after which they were quenched to 21°C for testing. The top line
is for materials cooled slowly from 90 to 21°C at O.5°C h-l, which is shown to be a
very effective aging procedure. These specimens were then reheated to 90°C and
quenched to 21°C to give the points marked o. (Reproduced from Struik, 1978,
with permission.)

Sternstein, Ongchin and Silverman (1968) observed the same effect for
PMMA in shear.

5.3.2 PHYSICAL AGING AND STRAIN SOFfENING


The effect of thermal or physical aging in raising the yield stress of glassy
polymers has been known for some time. It appears to have been first
observed by Horsley (1958) with Pvc. Since that time many other cases
have been reported. For example, increases in yield strength were found by
Golden, Hammant and Hazell (1967), when polycarbonate was heated at
temperatures of 80-140°C, i.e. below the glass transition temperature (~)
of 142°C, and this type of behaviour is now known to be a characteristic
feature of glassy plastics. One of the most instructive accounts of this effect
has been given by Struik (1978), whose results with PVC (Struik, 1978,
p. 27) are given in Figure 5.12. These measurements illustrate several
features of the physical aging process, including the characteristic increase
in yield stress with storage at a constant temperature. Materials quenched
234 The post-yield deformation of glassy polymers

Sample A •
60 Sample B •
r;;-
'E
Z
2 55
Vl
Vl
[':'
+-'
Vl

-0

'>-" 50
0'
c:
;:
''c:""
c;,
c:
w 45

400~---0~5----1~0----~1~5--~2~0~--~25·

Enthalpy change IJ g-I)


Figure 5.13 The relation of yield stress to enthalpy change during aging with PET.
Also note that polymer A has a higher molecular weight than B. (Reproduced from
Aref-Azar et at. 1983, with permission).

to 50°C from 90°C (above Tg ), and then kept at that temperature, show a
continuous increase in yield stress depending on the time allowed. It is also
shown that slow cooling from 90°C is a comparatively effective way of aging
Pvc. The reversibility of the aging process is also demonstrated, by taking
the well aged material from the slow cooling procedure and heating it to
90°C followed by quenching to 21°C, which gives the points (circles) at the
bottom of Figure 5.12, with the lowest yield stress. This ability to reverse
the aging process and rejuvenate the material by heating above ~ has been
confirmed in much subsequent work.
In Chapter 3 an account has been given of the enthalpy changes
associated with physical aging their evaluation by DSC. This makes it
possible to measure the energy deficit caused by aging, as compared with a
material which has been quenched or treated in some other specified way.
Sometimes it is found that when a glassy polymer is cooled very rapidly
from a temperature above ~, it gives a smooth curve with a zero or minimal
peak when evaluated in the DSC. In such cases the quenched polymer
provides a particularly good basis for estimating the energy change due to
aging. This can then be related to the increase in yield stress as in Figure
5.13, taken from the work of Aref-Azar et ai. (1983) on PET. The two curves
show a clear relation between the two quantities. However, even when
plotted in this way, there is an influence of molecular weight which leads to
Physical aging and large deformations 235

slightly higher yield stresses for the lower molecular weight samples. The
tendency is often observed for low molecular weight polymers to be more
strongly affected by physical aging than is the case for higher molecular
weight materials. It has been described by Marshall and Petrie (1975) for
polystyrene and by Adam et al. (1976) for polycarbonate. However PMMA
gives conflicting results (Figure 5.9).
It has already been noted that when a material which has been subjected
to a physical aging process is heated above Tg , the effects due to aging are
erased. The same is true for mechanical deformation at or beyond the yield
point. This was recognized by Struik (1978, p. 76) who reported the partial
elimination of aging effects in a tensile test even before the yield stress had
been attained. This type of behaviour was exploited by Bauwens (1978), who
produced a PVC test piece free of aging effects by a continuous alternate
bending treatment while it was cooled down from a temperature above ~.
A more striking demonstration of this phenomenon, which was reported
by Adam, Cross and Haward (1975), is illustrated in Figure 5.14a. These
results are similar to those given for PS in Figure 4.1, where the effect of
aging on the stress-strain behaviour is 'erased' after sufficient plastic
deformation, i.e. the stress-strain behaviour of aged and unaged samples are
the same after sufficient straining. DSC measurements for PS are shown in
Figure 5.14b (Hasan and Boyce, 1993), where it will be seen that the
enthalpy peak characteristic of the annealed material is gradually eliminated
as the deformation is increased, and at strains greater than 25% the DSC
curve coincides with that of the quenched material (cf. Figure 4.1). Subtrac-
tion of the quenched and annealed DSC curves for undeformed material
then give the enthalpy change due to annealing and the dIfference between
the stress-strain curves (shaded in Figure 5.14a) gives the associated
mechanical work. In the earlier work by Adam, Cross and Haward (1975)
the two quantities were found to be the same, but in the later measurements
of Hasan and Boyce using improved DSC techniques the two quantities
were found to be similar but not identical. These experiments also illustrate
the way in which the increased stresses due to physical aging are neutralized
during the early stages of yielding. Additionally, Hasan et al. (1993) have
observed an increase in free volume with compressive straining beyond yield
using positron annihilation lifetime spectroscopy measurements on annealed
(aged) and also quenched specimens of PMMA. The increase in free volume
on straining was accompanied by strain softening. Prior to deformation, the
annealed specimen was found to have a lower free volume content than the
quenched specimen; after 30% strain the two materials exhibited the same
stress-strain behaviour and also had the same free volume content, which
was higher than that observed in the undeformed specimens.
The DSC thermograms obtained by Hasan and Boyce (1993) also provide
more detailed information than was previously available, so that it is
236 The post-yield deformation of glassy polymers

0·2 O·L
a rnvE STRhl (I" 1/1.1

0.20
a Annealed

0.15

-;;0 I
........
~
'" 0.10 'f"
I'
:I:
<I
07-
,,/,'.
,iI. _
-rr-_
0.05
~~ /.
- 5 7-
- 10 7-
- 157-
'5--
__ ~~,~- ~-25 7-
............... - -.~.:'='
0.00
50 100 150
Temperalure (Oe)

Figure 5.14 A comparison of the work required to erase physical aging with the
equivalent enthalpy change. (a) Compression curves for aged and quenched polycar-
bonate coincide at a strain of -0.5. The shaded area measures the energy required
to erase the ageing effect which could be compared with energy differences
measured by DSC. Samples were annealed for 16 h at 130°C. (Reproduced from
Adam, Cross and Haward, 1975, with permission.) (b) DSC curves for annealed
polystyrene after different degrees of compression. The difference between the curve
at 0% and the similar curve in Figure 4.20 gives the energy deficit due to annealing,
and the difference in the work of compression can be derived from Figure 4.1 as in
(a) here. (Reproduced from Hasan and Boyce, 1993, with permission.)
Physical aging and large deformations 237

0 . 15

---..
00
"- -607.
f:.. 0 10
- 807.
.:x:
<l - 1407-
- 1707.

0 .05

0 .00
50 100 150
Temperalure (Oe)

Figure 5.15 DSC measurements for annealed polystyrene specimens compressed to


different applied strains. These results demonstrate the existence of storage energy
caused by strains above 40% .

possible to observe enthalpy differences in materials which have been


compressed to strains greater than 0.3, shown in Figure 5.15. There are two
main effects to observe on the DSC scan: first, the previously noted gradual
erasure of the ~ endotherm with strains up to 25%; and, second, develop-
ment of a pre- ~ exotherm indicating an energy storage mechanism during
plastic straining (also observed in Figure 5.14b). The pre- ~ exotherm
reaches a steady state configuration after 25% strain, i.e. it is identical from
25% up to 170% strain. Also, comparing Figure 5.14b with Figure 4.20, the
DSC scan of initial quenched versus annealed are identical after 25% strain.
However, F igure 5.15 also exhibits a second exotherm above ~ which
increases at very high strains (> 60%) and indicates a second energy storage
238 The post-yield deformation of glassy polymers

mechanism during large strain plastic deformation. Hasan and Boyce


associate the second storage mechanism with the molecular orientation
process which leads to strain hardening at large strains. The energy!entropy
effects due to orientation provide for the recovery (restoration of the
random configuration of chains) observed upon heating, as shown earlier in
Figure 5.3.

5.3.3 PLASTIC INSTABILITY


As described above, many amorphous polymers, especially those which have
been subjected to some form of physical aging, exhibit true strain softening.
In such cases, the deformation process itself progressively erases the effect of
aging. This leads either to an increase in the strain rate at constant applied
stress, or to a decrease in stress at a constant strain rate - i.e. to the
conditions which promote plastic instability and accentuate the localization
of plastic strain. Examples of this type of behaviour will now be given.

(a) Shear Bands


The role of thermal history and of strain softening in the development of
shear bands was demonstrated by Bowden and Raha (1970) and Bowden
and Jukes (1972); their work was summarized in the first edition of this
book (Bowden, 1973). Here it is presented in Chapter 4, where photographs
of shear bands formed under plane strain compression for PS and PMMA
are given in Figure 4.17. It will be seen that the sample of PS quenched from
100°C shows rather diffuse shear bands, whereas in that with the material
'cooled normally', the shear bands were very sharply defined. In their
analysis of these phenomena both Bowden (1973) and Argon and Bessonov
(1977) proposed that true strain softening was a prime requirement for the
formation of shear bands and therefore for the localization of plastic strains.

(b) Tensile Tests


It has already been noted that, although physical aging will generate true
strain softening in a conventional tensile test, because of geometric necking,
it is not observed so unambiguously as in other types of test. Nevertheless,
with some materials, and particularly with PVC, quite striking changes
occur as shown in Figure 5.16 after Cross, Haward and Mills (1979). In this
case the quenched material, at the lowest extension rate, gave nearly uniform
extensIon and the aged material necked decisively. Less striking, but similar,
differences in necking behaviour were observed in studies by Aref-Azar et al.
(1983) on the thermal aging of PET. The aged materials not only had a
higher yield strength and a greater fall in stress after yield, but also a sharper
Physical aging and large deformations 239

N
~
z
::!:
- 70
b

~ 60
UJ

'"
>-
Vl

e> 50
z ANNEALED

'"
OJ
IJUENCHED
~ 40
e>
zOJ
30

20

10

o0<-----------------.-~----1-0---I-I-c::-EN-GtNEERII'I3

.,
STRAIN

Figure 5.16 Engineering stress-strain curves for quenched and ar<nealed Pvc. The
material was first quenched in ice water from 90°C. The small maximum in the
resulting curve was not enough to cause necking at slow strain rates. When annealed
for 1 h at 65°(, 5 h at 50°C and 16 h at 40°(, the polymer necked strongly, as with
the normally received material which ages slowly at room temperature. (Reproduced
from Cross, Haward and Mills, 1979, with permission).

boundary to the neck or the incipient neck, in a manner similar to Bowden's


shear bands.
When necking occurs, the stress required to extend the test piece falls from
the yield stress to the drawing stress, thus securing a mechanical advantage
for the necking process, as compared with uniform deformation of the whole
test piece. The same factors should operate, on a microscopic scale, within
a craze. Both processes are therefore promoted by physical aging.

5.3.4 IMPACT STRENGTH


A transition from ductile behaviour to crazing with brittle fracture was
observed by Mininni et al. (1973) when PET was subjected to thermal aging.
Unannealed material gave a ductile stress-strain curve in a conventional
test; however, after annealing, brittle fracture, ascribed to crazing, occurred
at yield. With some very long thermal treatments the crazes proliferated to
such an extent that substantial extension of the test piece took place, as with
rubber-toughened polystyrene (Chapter 8).
240 The post-yield deformation of glassy polymers

10

-;;-70
E
-,
oX

"I
l-
60
\.?
~50 --. A
a: B
··0· -
l-
V)
.... 40
U
<t
~30
o
~20
U
I-
~10 .0--0-_
• 0 •• _
I

.o.d

-150 -100 -so 0 so 100 150


TEMPERATURE (OC)

Figure 5.17 The impact strength of quenched and annealed polycarbonate. Both
materials show a sudden transition from brittle to tough fracture as the temperature
is raised. The annealed material is always more brittle and the transition takes place
at a much higher temperature. A untreated, B annealed. (After Allen, Morley and
Williams, 1973.)

The initiation of fracture by crazes is also believed to determine the result


of many notched impact tests. For example, Allen, Morley and Williams
(1973) studied the effect of thermal aging and temperature on the Charpy
impact strength of polycarbonate. Except at temperatures below - 50°C, the
annealed samples always showed a lower impact strength than the untreated
material (Figure 5.17). Both materials showed a relatively sudden transition
from brittle to tough behaviour at a particular temperature which, however,
differed by as much as 100°C between annealed and untreated material and
did not correspond with the low temperature relaxation peak. However in
another Charpy test it was shown that the embrittlement was associated
with a drastic reduction in the volume of deformed material (Haward, 1980).
The effect of annealing in causing a transition from tough to brittle fracture
was confirmed and extended by Zurimendi et al. (1982), and by Cheng,
Keskkula and Paul (1992). Both groups additionally studied the fracture
surface obtained under the different conditions, which made it possible to
show that brittle fractures were initiated by crazes whereas the tough
materials gave a fracture surface characteristic of a ductile fracture process
(Comes and Haward, 1974). Zurimendi et al. (1982) also demonstrated that,
Physical aging and large deformations 241

as in the work of Aref-Azar et al. (1983) with PET, when annealing


treatment was intensified under brittle fracture conditions (higher yield
stress), the point of craze initiation moves closer to the crack tip. These
results suggest that the measured reduction in impact strength is caused by
the localization of strain in a craze and the consequent initiation of brittle
fracture. With material which has not been annealed, bulk deformation and
orientation can occur in front of the craze and this inhibits its further
growth. Fracture is then delayed until a ductile process is initiated at the
surface of the notch. This competition between the occurrence of craze
initiation versus a ductile shear initiation has been attributed 10 the effect of
aging, which raises the yield stress to a level which is greater than the stress
required to propagate the craze, thus, after sufficient aging, crazing and
brittle fracture occur instead of shear yielding and ductile fracture.

5.3.5 VOLUME CHANGES DURING AGING


The changes in mechanical properties which occur during the physical aging
process must be related to structural change within the polymer, which can
be eliminated by heating above ~ or during post-yield deformation.
Attempts to interpret the nature of the changes centre around estimates of
free volume and its distribution, enthalpy changes and on spectroscopic
studies of chain conformations, some of which have been briefly discussed
earlier.
That volume and density changes playa part in determining the observed
variations in mechanical properties can hardly be doubted. An interpreta-
tion on these lines has been strongly supported by Struik (1978) and has
already been presented in Chapter 3 and does not need to be further
developed here. Instead we shall consider the possibility of other factors
contributing to the process. In the first place we may note that the actual
changes in macroscopic volume, even during prolonged aging, are often
rather small and seldom exceed 0.2% (Bartos, Muller and Wendorf, 1990;
Lee and McGarry, 1991). The latter used different cooling programmes with
polystyrene to reach a particular temperature at which density changes were
measured against time. The graphs so obtained occasionally intersected,
which meant that material at the same temperature and volume could
contract at different rates. In other work on polystyrene, Roe and Millman
(1983) found that when, after a period of aging, the enthalpy deficit !1H had
reached a constant value, the creep compliance continued to decrease.
Another example of unexpected behaviour is described by Struik (1978,
p. 112); he found that after PVC had been held at a lower temperature,
annealing at 60°C led first to a decline in yield stress before the normal
increase began. All these observations emphasize the complexity of the
physical aging process and, although they may be explicable in terms of a
242 The post-yield deformation of glassy polymers

distribution of free volume, they are also strongly suggestive of the possibil-
ity that more than one type of structural change may be taking place.

5.3.6 STUDIES IN CHAIN CONFORMATION


These apparent contradictions have stimulated investigations of possible
changes in chain conformation by spectroscopic methods and especially by
means of Fourier transform infrared spectroscopy (FTIR). In this way
Koenig and Antoon (1977) measured difference spectra with aged and
untreated PVc. They found spectral changes which indicated that a certain
ordering occurred during aging which led to an increase in planar zigzags,
with no evidence of crystallization. This was also in line with the evidence
of Reddish (1966) who concluded from dielectric measurements that an
ordering process took place as PVC was cooled down towards its glass
transition temperature.
Somewhat similar conclusions have been reached in the case of
poly(ethylene terephthlate) (PET) where Ito et al. (1978) observed a reduc-
tion in the intensity of the trans deformation band during aging. These
results have been confirmed by Moore, O'Sloane and Shearer (1981), while
in a less specific sense Carfagna et al. (1988) found evidence of conforma-
tional changes during the annealing of poly(ether ether ketone) (PEEK).
There is however still some uncertainty about the extent to which different
chain conformations in general can contribute to variations in internal
energy. Roe and Tonelli (1978) carried out a series of calculations on
rotational states and concluded that they could only make a small contri-
bution to ~Cp (at ~) and hence to ~H. However, more recent work by
Mott, Argon and Suter (1993) on molecular mechanics simulations of the
structure and deformation of polypropylene and also by Bulatov and Argon
(1995) on the Monte Carlo modelling of general amorphous structured
materials shows the presence of very high internal stresses. These internal
stresses are distributed within the overall structure of the glassy state
material. This distribution in internal stress is a governing factor in local
inelastic events and also evolves with deformation, leading to the non-linear
and time-dependent recovery of polymers. The atomic level simulation
results are in direct accord with the DSC measurements of Oleynik (1990)
and Hasan and Boyce (1993).

5.4 THERMAL EFFECTS DURING THE DEFORMA nON OF


GLASSY POLYMERS
The deformation of a polymer glass is associated with several different types
of thermal response, some of which may act in opposite senses depending
on whether the material is subject to tension or compression. Recently this
Thermal effects during the deformation of glassy polymers 243

·4 20
o
15

~ ·2
c
ell
.r:.
()
~O~--~ __J -_ _ _ _ _ _- L_ _ _ _ _ _ ~ ___
E 250 300 350
~
Temperature (K)

Figure 5.18 The reversible heating effect with PMMA in compression. The reversible
Joule-Thomson temperature change is linear with stress and temperature (K). The
stress values in MPa are marked at the end of each line. (Reproduced from Gilmour,
Trainor and Haward, 1978a, with permission.)

whole field has been reviewed by Godovsky (1992) whose work will be
quoted in this discussion. Here the different types of behaviour will be
summarized under several headings, beginning with the thermoelastic effect.

5.4.1 REVERSIBLE THERMOELASTIC EFFECT

The thermoelastic effect is a property of all Hookean materials which obey


the classical laws of elasticity. Such materials exhibit a small heating effect
in compression or a cooling effect in tension often known as the Joule-
Thomson effect after its originators (Thomson, 1853; Joule, 1857), who
derived the now well known equation:
8T/da = TIXdJCpp
where 1X, is the coefficient of linear expansion, J the mechanical equivalent
of heat and P the density. Thus, when a plastic is subject to a tensile stress
(a negative) there is first a small cooling stage before the material heats up
due to viscoelastic processes. This was demonstrated on PVC by Binder and
Muller (1961). Following this work a more extensive and accurate study was
carried out by Gilmour, Trainor and Haward (1978a) who measured small
increases in temperature (~1 K) under compression. Good estimates of
these increments could be obtained by the repeated application and reversal
of the applied stress, and it was found that the changes in temperature were,
as predicted, proportional to the absolute temperature and to the applied
stress (Figure 5.18). After measuring Cp by DSC, it was possible to derive
values of 1X, which were found to be in good agreement with literature values;
indeed the method has the capability of improving the accuracy of measure-
ment of either one of the quantities 1X, or Cpo It can also be used for the direct
244 The post-yield deformation of glassy polymers

calculation of the thermodynamic Gruneisen constant (Gilmour, Trainor


and Haward 1978b).

5.4.2 ENERGY STORAGE DURING YIELD

A systematic discussion of the yield behaviour of glassy polymers has


already been given in Chapter 4. However, in describing the thermal effects
associated with the post-yield process it is also necessary to consider thermal
changes during yield, since any changes in temperature would significantly
influence the observed mechanical response.
Most theories of yielding propose changes in internal energy as an
essential feature of the process. For instance, Robertson (1966, 1968), Argon
(1973a) and Boyce, Montagut and Argon (1992) suggest that under the
influence of the shear component of stress, molecular chain segments rotate
and thus change their surrounding stress state. On the other hand, Bowden
(1973) and Oleynik (1990) have treated a glassy polymer in a manner more
analogous to a crystal, postulating some evolution in defect structure of the
polymer. Both treatments require that energy should be stored during yield,
a process which has been investigated by Adams and Farris (1988),
Nazarenko (1988), Chang and Li (1988), Rudnev et al. (1991) and Oleynik

II

20

~ '"
:::>
<l ....,,<»
0
10
:to

~

~.

10 20 30 40 10 20 30 40
a C ('" ) b £ (:t,)
Figure 5.19 The relation between the applied work W. the heat generated Q and
the change in the internal energy I1U measured under compression. Experiments
made at 30 0 ( in a deformation calorimeter for (a) PS, (b) Pc. (Reproduced from
Godovsky, 1992, with permission.)
Thermal effects during the deformation of glassy polymers 245

(1989, 1990) by deformation calorimetry and/or DSC measurement. For


most of these works, a Calvet-type calorimeter was supplied with a special
arrangement for the uniaxial compression of cylindrical samples. In Figures
5.19a and 5.19b, results obtained by Nazarenko (1988) are reproduced from
the review by Godovsky (1992). The small Joule-Thomson effect is not
visible in this work, which reports results for polystyrene and polycarbonate
in compression. It will be seen that there is a steady increase in internal
energy at compressions of up to 30%, but that with further compression up
to about 40% the mechanical work is largely converted into measurable
heat. This is consistent with the DSC measurements of Hasan and Boyce
(1993) discussed earlier in Figures 4.19, 4.20, 5.11 and 5.15 which showed
that the pre- ~ exotherm saturates at 25% strain.
The actual means by which this energy is stored during yielding is not yet
fully understood. It could be the consequence of distorting or rotating
covalent bonds, but Oleynik prefers to ascribe the process to inhomo-
geneous shear transitions or shear defects rather like dislocations in crystals.
He does not consider that the energy deficit is associated with the confor-
mation of the molecular coil although, in the case of PET, it was shown that
yielding was associated with a marked increase in trans configurations
(Oleynik, 1989). The molecular mechanics simulations of Mott, Argon and
Suter (1993) and the Monte Carlo simulations of Bulatov and Argon (1995)
presented earlier also show the presence of very large internal stresses which
evolve with straining and contribute to the non-linear nature of unloading
in these materials. While the actual energy storage mechanism is still open
to debate, the experimental data clearly show that a significant portion of
post-yield energy is stored and a portion is dissipated.

5.4.3 CONVERSION OF MECHANICAL WORK TO HEAT


According to the theory to be presented at the next part of this chapter, the
post-yield deformation is dominated by viscous and entropic forces. The
latter are treated according to established theories of rubber elasticity, but
the interpretation of this model in terms of the release of measurable heat
during post-yield plastic deformation is not clear and two hypotheses have
been applied.

1. If the two processes (viscous and entropic) are considered to operate


separately, as in a literal interpretation of the model (Figure 5.25a), the
entropy difference appears as measurable heat in the same way as in the
stretching of rubber. In this case, when the largest deformations are
employed, only relatively small proportions of energy may be stored,
either around yield or as internal energy associated with the extension of
the polymer chain (Erman and Mark, 1994). In this case high conversions
246 The post-yield deformation of glassy polymers

of work into heat are predicted, e.g. 90% or greater. Reversal of the
deformation occurs when the temperature is raised above ~ and the
retarding viscous forces become small.
2. The two processes are not separate and that part of the energy which is
treated mathematically according to rubber elasticity theory is stored as
a back stress in the deformed material together with the internal energy
contributions included in (1) above. This proposal, which is described by
Boyce, Montagut and Argon (1992), predicts a much lower production
of heat, e.g. 50-80%, and also provides an explanation for the ability of
a polymer to recover fully to its initial un deformed state upon heating.

5.4.4 ADIABATIC AND ISOTHERMAL DEFORMATION


The significance of adiabatic heating during the deformation of thermoplas-
tics has been recognized for a long time and is important in the drawing of
poly(ethylene terephthalate) filaments. A case of this type was analysed by
Marshall and Thompson (1954). As PET has a relatively high yield stress
and a draw ratio of about 4, it is inherently able to develop significant
quantities of heat. These workers measured a series of force extension curves
for this polymer at different temperatures and then calculated the adiabatic

5~-----------------------------------------------'

-d
_.L
1 3

.,..,.
~
f
- .I-
Q
5 2
'~
~

4 5
ExfenSlon rOhO

Figure 5.20 The isothermal and adiabatic extension of amorphous PET. Isothermal
extensions (a) 20°(, (b) 30°(, (c) 40°(, (d) 50°(, (e) 60°(, (f) 70°(, (g) 80°(,
(h) 100°(, (i) 140°e. A is the adiabatic curve calculated from 20°e. (Reproduced
from Marshall and Thompson, 1954, with permission.)
Thermal effects during the deformation of glassy polymers 247

curve assuming complete conversion of mechanical work into heat, i.e. they
did not allow for any energy storage during the initial yield step. Their
results are presented in Figure 5.20 where the difference in the nominal
stress-strain curve due to adiabatic heating may be clearly seen. They used
these calculations to describe the fast necking process during the drawing of
PET filaments, allowing also for the transfer of heat backwards through the
neck.
Because of the nature of the necking process in a tensile test or in the
drawing of a filament, both deformation and the generation of heat are
localized in a small volume of material, and this reduces the heat loss as
compared with uniform deformation. This means that the necking process
is particularly susceptible to heating effects, as pointed out by Marshall and
Thompson (1954). In a conventional tensile test when necking starts in the
parallel length of the test piece, the type of polymer can affect the relative
activity of the two necks which are initiated. For example, when polycar-
bonate was tested it was found that both necks at the ends of the deformed
part of the test piece were active, as shown by a rise in temperature. Under
similar conditions with PVC, which has a lower Tg and a yield stress more
sensitive to temperature than is the case with polycarbonate, all the
deformation took place at one neck while the other remained cool (Cross,
Hall and Haward, 1975).
However, in order to study this process and to determine the proportion
of energy which appears as measurable heat, it is necessary to measure the
heat output during necking, and this raises a number of problems.

5.4.5 EXPERIMENTAL DETERMINATION OF HEATING AT HIGH STRAINS


Measurement of the heat generated during large deformations of thermo-
plastics gives significant problems and the summary of published results by
Godovsky (1992, p. 227) includes values of - Q/W ranging from 0.5 to 1
where - Q is the heat released and W the work applied. However, only a
few studies have been published for polymer glasses and here we shall
discuss the work of Adams and Farris (1988), on the one hand, and of
Maher, Haward and Hay (1980) and Koenen (1994) on the other. All these
measurements were concerned with measuring the heat given out during
necking. Two quite different methods were used. Adams and Farris meas-
ured the heat generated using a Calvet-type deformation calorimeter with
different rates of extension. The other workers measured the temperature in
the neck from infrared radiation.

(a) The Work of Adams and Farris, and Supportive Work


These workers used a Calvet deformation calorimeter, which measures the
pressure changes in the gas surrounding the sample situated in a sealed
248 The post-yield deformation of glassy polymers

chamber and relates it to the emission of heat. Polycarbonate samples were


drawn uniaxially to their natural draw ratio under essentially isothermal
conditions. This limited the drawing speed to 5 mm min - 1, but below this
level the rate of extension was varied by a factor of 10. Their results gave
values of Q/W which increased with an increasing rate of extension from
0.56 to 0.79. These measurements correspond with the predictions of Boyce,
Montagut and Argon (1992) and so favour the hypothesis (1) listed in
section 5.4.3. Of course, the stored energy which does not appear as heat
includes a proportion of that produced during the yielding process, as
previously described. Evidence that this energy is retained during further
deformation comes from DSC work by Hasan and Boyce (1993) with
polystyrene, as shown in Figure 5.15. The figure also shows that there is a
significant if somewhat smaller amount of energy released above ~, also
observed by Chang and Li (1988). Hasan and Boyce (1993) showed that this
takes effect above 40% strain and increases further with additional strain.
More recently, Arruda, Boyce and Jayachandran (1995) used an infrared
detector to monitor the surface temperature (at a single spot) during
compression testing of both PC and PMMA at constant strain rates of
-0.001, -0.01 and -0.1 s -1 The temperature rise and corresponding
thermal softening of the material was found to increase with increasing
strain rate where a maximum AT of 20°C at a strain rate of -O.1/s after a
true strain of -0.9 was measured. This rate did not result in fully adiabatic
conditions, and a first-order heat transfer analysis indicated that heat loss
due to conduction into the steel plates did occur. These results, including
the dramatic post-yield thermal softening observed in the stress-strain data,
were modelled according to the principles presented in Boyce, Montagut
and Argon (1992) and good agreement was obtained.

(b) Estimates based on Infrared Measurements


The use of infrared radiation to measure temperature directly has the
advantage that it does not limit the rate of extension that may be used,
although at low rates temperatures are limited by the loss of heat to the
environment. With this method Maher, Haward and Hay (1980) calibrated
an infrared camera to measure the temperature of black plastic polycarbon-
ate sheets. This made it possible to record the distribution of temperatures
in the neck as shown in Figure 5.21. They also measured the difference AT
between the maximum temperature in the neck and the ambient tempera-
ture as a function of the rate of extension Vn applied to the strip of film. A
plot of l/AT against l/v n , based on a semi-empirical argument, was found
to be linear, which enabled a value of AT to be extrapolated for adiabatic
conditions and hence to estimate the proportion of the mechanical work
appearing as measurable heat. More recently the infrared technique has
Thermal effects during the deformation of glassy polymers 249

":i~tt:tt(== h)
" t-+H-t-i-- -Il)
121
U7''-I-+-+..--- - ( 4)
b 1'---I--tr-- - (SI
:l;r-l--tr- - -(61
:-tC-..tJ-- -17I

1 --I-t--- - -(81

1 (91

-1->,1
~ I
" I
Dlogrom 01 photograph 01 t~mperotUf~ profile
ot 0 n«kong polycorbonole speCimen
(I' 19l'C,I2J l7l 'C, (l, l53 'C,(41 3H ,(51 3IS ' C,
(6) 29 9' C,I1I 111'C, (8) 2S6'C ,(91 23 O' C

Figure 5.21 Temperature distribution in a polycarbonate test piE'ce during necking.


A moderate rate of extension was used, which allowed significant heat losses to the
environment. (a) Infrared photograph of neck, (b) temperature map.

been significantly improved by Koenen (1994), both by improving the


accuracy of the infrared camera and by introducing a mathematical solution
of the one-dimensional equation for heat conduction and loss.
Koenen's measured and calculated results are presented in Figure 5.22
where the effect of heat loss to the environment and the consequent
distribution of temperature during and after neck formation is presented. As
with Maher et al. (1980), Koenen reports near adiabatic conditions with
drawing speeds above 10 - 3 m s -1 where 100% conversion of applied work
to heat was also observed. At lower rates of extension the figure illustrates
the fall in temperature after necking. Under these conditions, however,
Koenen calculates separately the rate of heat generation in unit volume
during neck formation, designated as a stress, and reports a lower conver-
sion of work into heat as observed by Adams and Farris (1988).

(c) Some Factors Which Influence Observed Heating Effects


The apparent effect of strain rate on the conversion of work to heat is
unexpected, but it should also be noted that two very different experimental
250 The post-yield deformation of glassy polymers

=
Vn=2.06 10.5 mis, qt=18MPa vn=1.17 10-4 mis, qt=20MPa

<t' -0.05
x/m
o
J ;'~C/L -0.05
x/m
o

Figure 5.22 The temperature in a polycarbonate test piece ploUed against the
distance from the point of neck formation (0). Different drawing velocities were
used, and the energy input q, was separately estimated as a stress. The continuous
lines were calculated from a model for conduction and heat losses and the
discontinuous curves were as measured. At the highest extension rates it was found
that all the work was converted to heat. The rise in temperature flT is represented
by S in this figure. (Reproduced from Koenen, 1994, with permission.)

methods have been used. Adams and Farris employed low strain rates and
isothermal conditions, while in I.R. measurements where generally higher
extension rates were used there were temperature rises of 20-30 K. Is it
possible that the storage of energy is time dependent?
The tendency in the infrared experiments for - Q to exceed W by a small
fraction is also unexpected and is within experimental error in the work of
Maher, Haward and Hay (1980) if not with that of Koenen (1994). It is not
however impossible. As noted by Godovsky (1992, p. 225), the conforma-
tional energy may both increase and decrease, depending on the sign of the
value of d(ln(r 2 )0)/dT (Erman and Mark, 1994), where (r2)0 represents the
sum of the squared end-to-end vectors between crosslinks (or entangle-
ments). Some measurements by Stolting and Muller (1970) suggest that it is
possible under certain conditions for this quantity to decrease with tensile
Thermal effects during the deformation of glassy polymers 251

strain. They extended polystyrene at 112°C, well above ~ and compared the
heats of solution of deformed and isotropic materials. Under these condi-
tions the extension will take place effectively at zero yield stress, as with a
rubber, so that the type of energy storage at yield described previously will
not be present. Their results pointed to a lower energy state for polymer
extended under these conditions. Of course, similar arguments apply to
rubbers; Erman and Mark (1994) reported that with isoprene and butadi-
ence rubbers, internal energy absorbs 10-20% of the work applied.
From the above arguments it is obvious that the storage of energy and
the production of heat during the deformation of a thermoplastic is a much
more complex process than had been initially supposed. Much work
remains to be done.

5.4.6 SELF-OSCILLATION IN THE DRAWING OF FILMS AND FIBRES


Some time ago Hookway (1958) observed an unexpected phenomenon
during the drawing of thermoplastic filaments. When the polymer is being
drawn through the necking stage the process at first proceeds smoothly, but
as the length of the drawn material increases a characteristic type of
self-oscillation is observed. This apparently peculiar behaviour has now
been confirmed by a number of workers including Richards and Kramer
(1972), Barenblatt (1970) and Toda (1993). The latter two developed a
theoretical account of the process.
An extensive experimental investigation was also carried out by An-
drianova, Kechakyan and Kargin (1971) who studied the extension of a
strip of PET film. As with Hookway (1958), they observed a normal yield
point followed by necking and steady drawing (Figure 5.23a). However, as
the length of the drawn material increased, regular fluctuations began to
appear and after a time this process became firmly established. This was
shown to be a thermal phenomenon. As the drawn material grew longer, its
ability to store elastic energy increased, so that when the stress would
normally have fallen during the necking process, the rate of strain increased
instead, leading to an adiabatic softening process which continued until
strain hardening became equal to the stored elastic stress in the length of
the already necked material. The neck could then cool down, and no further
extension would take place until the stress had once more risen to the level
required for the cold necking process to start. Thus all the deformation took
place in short time intervals following stress peaks, as shown in Figure 5.23b
and 5.23c. When the sample was cooled in a water bath, all the fluctuations
disappeared. It was also shown, by dusting the film surface with heat
sensitive crystals that the sudden extension of the neck was accompanied by
an increase in temperature up to 90°C and even up to 140°C at high
252 The post-yield deformation of glassy polymers

700

~ 500
2 3
'"
-'"
.,;
"'~ 300
(/)

100

10 20 30 40
Time, sec
(0)

400
N
E
v
"-
~
350
"'"'~
(/)

300

Time, sec
(bl

30·0,

:;; 20'0f-
::::
E
E
U
Q)
Q) 10'Ot-
Q
(/)

50f-
0·725 ~
2 3 4 5
Time, sec
Ie)

Figure 5.23 Thermal self-oscillation in the necking of PET film. (a) Actual stress
development and growth of self-oscillation during extension at a constant rate.
(b) Variation of stress with time. (c) The accompanying change in the true rate of
deformation in the region where the polymer passes into the neck. The broken line
shows the imposed average rate of deformation. (Reproduced from Andrianova,
Kechekyan and Kargin, 1971, with permission.)
Thermal effects during the deformation of glassy polymers 253

deformation rates. Very similar results were obtained with nylon but in this
case visible banding was observed on the drawn filament (Richards and
Kramer, 1972).

5.4.7 THERMAL FRACTURE


During the self-oscillation process described above, PET reaches relatively
high temperatures, well above the glass transition temperature of the
amorphous material. However, even under these conditions the material
strain hardens to an extent sufficient for it to bear the elastic stress
remaining in the length of necked material. In the case of PET this process
is probably reinforced by crystallization. However, this explanation raises
the further question concerning what would happen with a polymer which
does not easily crystallize and which has a relatively low ~. This led Cross
and Haward (1973) to undertake a series of tensile experiments with PVC
strips (I;, 67°C) in which the length was increased in a series of steps. With
test pieces of a conventional size, necking was initiated at a shear band and
then propagated in the normal way, but as the length of the test piece was
increased, the small dip observed immediately after yield (as in Figure 5.23a)
became larger. At the same time the proportion of fractures which occurred
at yield increased, until at a length of 0.30 m all the test pieces broke. When
the broken test pieces were photographed in the infrared camera immedi-
ately after fracture, the results shown in Figure 5.24 were obtained and the
temperature rise in the fracture zone was estimated as 24' C. Micrographs
of the fracture surface showed that melting had occurred in much the same
way as with polystyrene (Figure 5.1), i.e. there was further heating during
the fracture process. When polycarbonate (I;, 147°C) was used instead of
PVc, no fracture occurred with the lengths available in the Instron
ten so meter. However, when the polycarbonate was physically aged by
heating at a temperature below I;, to increase both the yield stress and the
fall in stress after yield, thermal fractures were observed.
This work raises the question as to how far the fracture of glassy materials
might be promoted by temperature rises. Although relatively few measure-
ments have been made in this field, there is strong reason to believe that
such effects may be considerable. For example, Fuller, Fox and Field (1975)
studied the progress of a crack in PMMA and PS by means of high speed
infrared photography. As the crack approached and passed the camera there
was first a small drop in temperature (thermoelastic effect) followed by a
large rise as fast crazing was observed. The temperature rise measured in
this way within a period of 10 !is was estimated as 450 K above the starting
temperature for PMMA and 400 K for PS. Thus temperatures well above
the glass transition temperatures, as normally measured, occurred in each
case.
254 The post-yield deformation of glassy polymers

Figure 5.24 The thermal fracture of Pvc. When the test piece is part of a long strip.
the initial shear band extends adiabatically and fracture is promoted by thermal
runaway. This infrared picture illustrates the temperature rise of 24°( when the
fracture takes place.

More recently Dickenson et al. (1994) have studied the temperatures of a


polystyrene surface during fracture by measuring the velocity distribution of
the volatile species emitted. The estimated values of 600 K and above agree
generally with the previous results. They also made electron micrographs
which showed evidence of melting during fracture similar to that given in
Figure 5.1.

5.5 MODELS FOR LARGE STRAINS

5.5.1 GENERAL HISTORY


Spring and dash pot models have been employed to illustrate the deforma-
tion of high polymers from the time that their mechanical properties were
first studied, for example by Tuckett (1943) and Alfrey (1948). The former
used a three-part model, the main feature of which was a spring and dash-
pot in parallel. However, the spring used in this and other similar models
Models for large strains 255

was generally of a Hookean character, where stress was proportional to


strain. Such models work well enough for small deformations and could
be characterized by a relaxation time. When problems were encountered
in reproducing experimental results, recourse was made to combining
a series of such models with different constants and having a distribution
of relaxation times. These ideas worked well for small deformations at
low stresses, but are not applicable to the large deformations under high
stresses.
However, from the beginning, a number of workers who studied the large
deformations of plastics and rubbers recorded and recognized analogies
between the two. For example, it was found that there was a finite limit to
reversible, or potentially reversible, highly elastic deformations and several
workers tried to relate the limit to molecular characteristics. For example,
Meyer and Van der Wyk (1946) suggested that large deformations depended
on a form of chain straightening which took place in segments shorter than
the whole molecule, while Hoff (1952) considered that the relevant chain
lengths were limited by entanglements. Haward (1949) listed examples of the
maximum extensibility for several polymers including both rubbers and
cellulosics, and considered that the differences were largely influenced by the
structure of the polymer chain. Such ideas were developed further by the
work of Allison and Ward (1967) into the concept of an underlying elastic
network (section 5.2.6).

5.5.2 QUANTITATIVE MODEL FOR LARGE STRAINS


Following the earlier work, Haward and Thackray (1968) specified a
particular, modified spring and dashpot model, which will be referred to as
the H-T model and is illustrated in Figure 5.25. Superficially the model was
much like those which had been proposed by Tuckett (1943), Alfrey (1948)
and other workers over the intervening period. The novelty lay in the nature
of the spring and dashpot employed. For the viscosity, any relation could
be used appropriate to a high polymer. In the ensuing calculations to
demonstrate the application of the model, the Eyring (1936) viscosity
equation was applied and the spring was then assumed to follow the laws
of rubber elasticity as presented by Treloar (1958). To this end the Langevin
equations, derived from non-Gaussian statistics, were selected as they
provided a limit to chain extensibility in accordance with earlier experimen-
tal work. Since the model was intended to cover the behaviour of solid
polymers, the constants of the viscosity equations had to be appropriate to
a high viscosity dashpot and wen, derived from measurements at yield.
Under these conditions, except at the highest deformations, any possible
stress due to the rubber component would lead only to an exceedingly slow
256 The post-yield deformation of glassy polymers

linear
elastic
spring

visco plastic rubber


element elasticity
Langevin
'spring'

(a)

200

0
0.0 1.5
(b) true strain

(c)
~ 10
X

Imax/lO=l..max
I max
X

~ 10
,~
(d) Ilia =1..
Models for large strains 257

rate of retraction and hence to the experimentally observed permanence of


the deformation in the absence of applied stress.
The above argument avoids the question as to whether or not there is
actually a stress present in deformed material due to the rubbery spring.
According to accepted rubber elasticity theory, the various chain elements
of a deformed elastomer coil are free, without further energy requirement,
to assume their highest entropy situations. This may involve millions of
local adjustments which are easily made in the rubbery state, but which
entail a significant energy loss under the high viscosity conditions appropri-
ate to a solid polymer. Further, as reported in Chapter 4, certain changes
are associated with the beginning of the yield process. Before retraction
yielding can occur, there may be a need for a segmental rotation of chains
against intermolecular barriers (Argon, 1973a) or large cooperative rearran-
gement of many chain segments, as shown in recent atomistic simulations
(Mott, Argon and Suter 1993) or for shear transformations (Oleynik, 1990)
as previously described (section 5.4.2). Thus the capability of the polymer
chain to move freely from one conformation to another, which is essential
for rubber elasticity, may be absent in the proposed model, and in that sense
there is no stress. However, in the same paper Argon, while endorsing the
model in general, indicated that the employment of rubber elasticity theory
for the spring was really just a particular way of representing entropy
changes. Indeed, as shown in Chapter 2, this proposition is supported by the
work of Mckechnie et al. (1993) who showed that the strain hardening could
be derived from a model which did not include the crosslinks required by
the theory of rubber elasticity. Thus it is appropriate at this point to
consider the way in which entropy changes are introduced into the theory
of high polymer deformation processes (Treloar, 1975).

5.5.3 ENTROPY AND POLYMER DEFORMATION PROCESSES


For a reversible process, as in the deformation of rubber, we have the
relation (Treloar, 1975)
dU=TdS+dW

Figure 5.25 Diagram for the H~ T model for large deformations in thermoplastics.
(a) The original model as shown here comprised a Hookean spring in series with a
Langevin rubber elasticity spring (TR and a dashpot (Tv in parallel. Later a Gaussian
spring was also used for the rubber component. Any appropriate equation could be
used for the viscosity. (See also Chapter 4.) (b) H~ T model as developed in section
5.7 to include the effects of physical aging. The initial aging peak is erased as
deformation takes place and the stresses are represented as in (a). (c) The deforma-
tion of a polymer coil between entanglements according to the Langevin model
where the coil is completely extended and a limit of extensibility Amax is reached. (d)
The Gaussian equation represents the deformation of a coil which does not reach a
limit.
258 The post-yield deformation of glassy polymers

where U, Sand Ware respectively the internal energy, the entropy and the
work done on the system by external forces. It is also convenient to
introduce the concept of the Helmholtz free energy A, where A = U - TS,
and at constant temperature dA = d W Then, since the work done on a
system by a force f acting over a length dl is f dl, we obtain

f = (dAjdlh = (dUjdlh - T(dSjdlh

In the classical rubber elasticity theory the quantity (dU jdlh is assumed
to be zero (although small changes are known to occur; Erman and Mark,
1994). Statistical theory is then applied to calculate the entropy differences
associated with changes in the initially random configurations of long chain
molecules, and when crosslinks are introduced these lead to the established
equations for rubber elasticity. The quantitative results are, of course,
dependent on the particular assumptions relating to the configurations of
the statistical chain and to the nature of the constraints, e.g. crosslinks which
are introduced, but the entropy changes are related directly to differences in
the spatial distribution of the polymer chain whether or not cross links are
present.
For glassy polymers it has been shown by neutron scattering measure-
ments (Kirste, Kruse and Schelten, 1973; Wignall, Schelten and Ballard,
1973) that, as originally proposed by Flory (1949), the random statistical
chain associated with the melt is preserved when the polymer is cooled
through the glass transition. It is also clear that orientation of the macro-
molecules also takes place during large deformations (e.g. Ward, 1984)
which must involve drastic changes in molecular conformations. Thus any
large deformation must entail entropy differences and the inclusion of
appropriate terms in any quantitative treatment of the process is clearly
necessary.
In semicrystalline polymers the situation may be more complicated,
since strain hardening is also observed in non-polymeric crystalline
materials. However, Flory and Yoon (1978) have shown with polyethylene
that crystallization is so fast that the random structure present in the melt
should be preserved during crystallization and their conclusion has been
confirmed by neutron scattering studies (Yoon and Flory, 1977; Yo on,
1977; Schelten et al. 1976). It is also clear that, as with glassy polymers,
large scale orientation of the polymer molecules takes place during defor-
mation (Ward 1984). Recent studies by Lee et al. (1993) clearly identify
the role of both the crystalline phase and the amorphous phase in deforma-
tion where both experiments and simulations were used in their investiga-
tions.
Models for large strains 259

5.5.4 DEVELOPMENT OF THE ENTROPY OR H ~ T MODEL

The above discussion provides the theoretical foundation for the H - T


model for large deformation processes in thermoplastics. As shown in Figure
5.25, the model includes a Hookean spring in series with the spring and
dash pot in parallel. The Hookean strain is essential to model the first part
of the stress-strain curve, where it supports the general picture but is not
accurate in detail, for which purpose a more complex treatment would be
required. However, errors of a few percent in the first part of the curve are
not of major significance in relation to a large deformation of several
hundred percent. The dashpot may be represented by any of the several
viscosity or yield stress equations which have been proposed, e.g. the
Eyring viscosity equation as originally suggested or one of its modified
versions (Bauwens, Bauwens-Crowet and Homes, 1969), or other theories
proposed by Robertson (1966, 1968), or Argon (1973a) as described in
Chapter 4. An important feature of the model is that, at temperatures well
below ~, the viscosity factor is in principle entirely responsible for the effect
of strain rate, which can be separated from entropy changes. Experimental
support for this approach has been reported by Haward (1942), Ender
(1968) and G'Sell and Jonas (1979, 1981). See also Figure 4.11.
The entropy contributions may be provided by either the Gaussian
(Argon, 1973b; Haward, 1993) or the Langevin (Haward and Thackray,
1968; Boyce, Parks and Argon, 1988a) theories of rubber dasticity, though
they also raise a fundamental difficulty. Both rubber elasticity theories
require the presence of chemical crosslinks which are absent in thermoplas-
tics. To overcome this difficulty, it is common to introduce the concept of
entanglements and the associated number of active chains as a substitute for
crosslinks. This was done in the original paper (Haward and Thackray,
1968) where, as in some more recent examples, the results seemed to be in
accord with measurable chain characteristics. It is probable that there is still
much to be learned about the best way of representing the topological
constraints fundamental to a system of condensed but separate polymer
chains. Indeed, recent work by Arruda, Boyce and Jayachandran (1995)
proposes that the size of the effective statistical unit can depend on
temperature. Many well established variations within the theory of rubber
elasticity also remain unexplored.
It should also be stated that a number of empirical relations, some of
them related to those in use for metals, have been applied to thermoplastics.
Some of these, e.g. those proposed by G'Sell and Jonas (1979, 1981) provide
a good representation of experimental results and are convenient for
mathematical manipulation. They have not been detailed here for the
reasons given above. In the next section the use of the Gaussian and
260 The post-yield deformation of glassy polymers

Langevin theories will be explored within the general framework of the H - T


model.

5.6 APPLICATION OF THE GAUSSIAN THEORY TO THE STUDY


OF LARGE DEFORMATIONS IN THERMOPLASTICS

5.6.1 THE GAUSSIAN EQUATION


In the Gaussian treatment the polymer coil does not approach a fully
stretched condition and the elastic stress in tension is given by Treloar
(1975) as
true stress = (TR(true) = G(,F - 1/,.1,)
Here the modulus G = pRT/Me where p is the density and Me is the average
molecular weight of the chain between effective crosslinks or entanglements
and Me/Me ~ 2 (Ferry, 1980). In the present context it is written as Gp ' the
strain hardening constant. Further, when A ~ 1 at the beginning of the large
strain process, (Ttrue ~ (Tn = (Ty' the yield strength. Then, at constant strain
rate,
(5.1)
and
(5.2)
The Gaussian treatment described by Treloar includes different types of
stress. For plane strain compression these are given by G(A2 - 1/,.1,2), and for
shear Gy, where y is the shear strain. The latter is an important conclusion
of the Gaussian theory, generally described as Hooke's law in shear. For
example, Rivlin (1948) demonstrated that equation 5.1 can be derived from
it without further assumptions. The Gaussian theory therefore predicts that
the effect of strain hardening in tension will be less than for plane strain
compression but greater than for shear. However, because the stresses under
consideration are relatively large, both (Ty and Gp may be expected to depend
on state of tension/compression of the material (sections 5.2.6 and 4.3.3).

5.6.2 MEASUREMENTS IN TENSION


As indicated in Figure 5.25, the extension ratio may be corrected, if
appropriate, for the small Hookean deformation. True stress-stain curves
may then be plotted according to equation 5.1. The relative simplicity of
these equations has many advantages, making them easy to comprehend
and manipulate. Physically they are a combination of a constant viscosity
and a modulus, as in the model. Their utility depends on the extent to which
The study of large deformations in thermoplastics 261

the strain hardening term dominates other influences in determining large


strain properties. However, as already pointed out, only a small proportion
of the stress-strain curves recorded in the literature are true stress-strain
curves suitable for this type of evaluation. The important question is then
how far the Gaussian relations represent the relevant experimental results.
This is best elucidated by plotting measurements of true stresses against
(Je 2_1j),) in tension, or against the appropriate Gaussian function in other
cases.
Measurements of true stresses and strains should preferably be made at a
constant true strain rate, but in accordance with the model, it has been
shown for polyethylene that the change of strain rate during a constant rate
of extension experiment causes only a small change in the slope Gp (Haward,
1993; Figure 4.11). So good straight lines may be obtained with either
method. Many of the published tensile true stress-strain curves, including
many semicrystalline polymers, have now been plotted in the Gaussian form
and in many cases they follow equation 5.1 very well (Haward, 1993; G'Sell
and Jonas, 1981). In some cases linear results are obtained up to very high
values of A with no sign of a limiting network strain. Polymers not giving
good Gaussian lines are high molecular weight PMMA (Hope, Ward and
Gibson, 1980), PS, PET and, most importantly, polycarbonate (G'Sell,
1982), although the latter has been modelled using a Gaussian relation up
to moderate strains (0.4; Lee and Luken, 1986). This is not surprising since
Gaussian and Langevin equations give similar results at low or moderate
strains. PMMA, PC and PS have been found to follow a Langevin relation
up to high strains and have been rigorously studied by Boyce and co-
workers, as described later.
Several examples of Gaussian plots will now be presented. In Figure 5.26
the results of Cross and Haward (1978, 1980) for PVC are shown for tension
and plane strain compression. The tension measurements are made at a
constant rate of extension on material which had been quenched in ice-water
and then extended uniformly at a low strain rate at normal temperatures. A
good straight line was obtained with a slope (G p ) of 13.2, comparing well
with the value of 14 reported by G'Sell and Jonas (1981) in a series of
Gaussian plots (Figure 5.27) based on their measurements at a constant true
strain rate (Figure 5.8). Figure 5.27 does not include a line for the
polycarbonate curve shown in Figure 5.8 as this polymer deviates from the
Gaussian relation in a manner expected of a polymer which follows the
Langevin equation as described in section 5.7. Figures 5.28a and 5.28b
represent the results of Bessonov et al. (1987) with polyimide OPO, an
amorphous polymer which deforms uniformly. It should be noted that the
slope (G p ) falls as the temperature is raised, a trend also observed with the
Langevin model (section 5.7). The general applicability of the Gaussian
262 The post-yield deformation of glassy polymers

Figure 5.26 Gaussian plots for Pvc. Open points for quenched PVC which exhibited
uniform deformation in tension. Filled points for lightly plasticized Pvc. The line for
plane strain compression showed a steeper slope and a higher yield stress (intercept)
than the one for tension. (Reproduced from Cross and Haward, 1978, 1980, with
permission.)

equation has been considered by Haward (1993) who showed that, in


addition to the above examples, a goed measure of agreement with equation
5.1 was obtained with low molecular weight PMMA (Hope, Ward and
Gibson, 1980) and PVC (Vincent, 1974). Certain other polymers which
extended uniformly also gave linear plots, namely cellulose nitrate and
acetate butyrate, and an amorphous ethyl, propyl isocyanate copolymer.
A case of particular interest is that of poly(aryl ether ether ketone). Here
the results of G'Sell et al. (1992) and of Biddlestone et al. (1986) both gave
a good linear plot for the semicrystalline material. Using measurements of
post-necked material, the latter also obtained a good straight line for the
amorphous polymer (Figure 5.29). There was, however, a difference between
the slopes and the intercepts associated with the two different structures.
The study of large deformations in thermoplastics 263

150

o
<>-
=100
b

OL-______L-____ ~ ________L -_____ ~ __ ~

o 2.5 5
(~- ~A)

Figure 5.27 Gaussian plot reported by G'Sell and Jonas (1981). These results include
many semicrystalline polymers, viz: PA, (polyamide) 6 and 66; PVC, poly(vinyl
chloride); PP, polypropylene; HDPE and LDPE, polyethylenes; and PTFE, polytetra-
fluoroethylene. The slope here represented by K is equivalent to Gp in the present
treatment. The value of 2.2 tor LOPE compares with values of 1-2 measured by Mills,
Hay and Haward (1985), where it was shown to depend on molecular weight. The
value of 14 for PVC compares with 13.2 from Cross and Haward (1978). The
measurements for polycarbonate (Figure 5.8) are not included. Polycarbonate has
not been found to fit the Gaussian equation under normal conditions.

The presence of crystallinity was found to increase both (Jy and Gp , although
it did not affect the form of the relationship.

5.6.3 GENERALIZATIONS BASED ON THE GAUSSIAN EQUATION


The results for plane strain compression (Figure 5.26) suggest that, as with
the Langevin system, the Gaussian treatment may be capable of describing
deformations other than those occurring under simple tension. It should
also be capable of including terms to allow for true strain softening, as will
be described later for polycarbonate. However, as already pointed out, the
simplicity of the tensile equation lends itself to manipulation and makes it
possible to illustrate certain features of the tensile test. Examples of this
approach will now be given.
264 The post-yield deformation of glassy polymers

0, MPalm 2

250 ,,-196°

/(,.'25'
rr
200

·25' ') -II

150 x-~

,,+200°

(a)
E, %

300
___ -25C
250
-+--50C
_ _ 100C

---150
200 -Er-200C
CIS
a..
~ .
.6
150

0 100

50

0
(b) 0 0.5 1.5 2 2.5 3
f...2 -1/f...

Figure 5.28 (a) Nominal stress-strain curves for polyimide OPO at different tem-
peratures, e= 1.7 x 1O- 4 s- 1 . (b) Gaussian plots for polyimide OPO at different
temperatures (OC). The polymer deforms uniformly and gives good Gaussian plots
whose slope decreases as the temperature is raised (section 5.7). The formula of the
repeating unit is shown. Curves kindly supplied by Or Laius. (Reproduced from
Bessonov et al., 1987, with permission.)
The study of large deformations in thermoplastics 265

200
CRYSTALLINE
190
180
(.
"'E 170 <>

e.Z 160
(fJ 150 0
(fJ
LU 140
a:
I- 130 AMORPHOUS
(fJ
LU 120
::J
a: <>
I- 110
0
/
100 /'

90
80

.1 .2 .3 .4 .5 .6 .7 .8 .9 10
(t-11A)

Figure 5.29 Gaussian plots for PEEK, measured by the further uniform extension of
material that has passed through the neck. Both crystalline and amorphous materials
give good linear plots, but the crystalline material has a higher value of Gp and yield
stress.

(a) Condition for Necking


Starting from the 'theoretical' nominal stress-strain curve given by
O"n = O"y/)' + Gp (). - 1/).2)
we obtain
(5.3)
For necking to take place there has to be a maximum in the nominal
stress at yield followed by a fall. At this point we may put). '" 1 and the
requirement for necking is then that dO"n/d)' should be zero or negative. For
this the critical condition is O"y/Gp > 3. At the present time there are no
known exceptions to this rule (Haward, 1993). For example, cellulose
acetate and acetate butyrate, each of which have O"y/G p < 2, both extend
uniformly under tension.

(b) Minimum Engineering Stress


The fall in the theoretical engineering stress at yield is followed by a
minimum, after which the stress starts to rise again. This occurs where
266 The post-yield deformation of glassy polymers

Minimum on the True Engineering stress/strain curve


6

5
~I
C '

I' 4
~
0'" 3

o
o 5 10 15 20 25 30
Y/G ,()..2+2/)..)
p

Calculated Draw Ratios for


12

10

1 3 5 7 9 11

Figure 5.30 Generalizations based on the Gaussian equation , (a) A minimum in


nominal stress according to equation 5.4 which coincides with the second tangent
in Figure 5.6. (b) Estimates of the natural draw ratio in terms of the quantity (J/Gp
(shown here as Y/G p ) according to equation 5.5.

do'n /dA = 0 for the second time and A is defined by the equation (Figure
5.30a)
(5.4)
This condition also corresponds to a tangent to the true stress-strain curve
The study of large deformations in thermoplastics 267

and means that all Gaussian stress-strain curves which meet the condition
for necking also meet the Vincent -Considere condition for stable necking
(Vincent, 1960; Figures 5.5 and 5.6). In addition, when any part of the
material passes through the extension ratio corresponding to the minimum
in the theoretical engineering stress-strain curve, a further increase in strain
will lead to an increase in the engineering stress, i.e. such material will
deform uniformly. This situation may be achieved in practice by orienting
the polymer, e.g. by a controlled rolling process where subsequent uniform
extension could be achieved at high levels of orientation (Broutman and
Patil, 1971; Snyder, Hiltner and Baer, 1994). This work has much in
common with the results of Ward and coworkers on preoriented filaments
(section 5.2.6).

(c) Estimation of the Natural Draw Ratio

The theoretical engineering stress-strain curve may also be used to provide


an estimate of the 'natural draw ratio' of material which has passed through
the neck in a conventional tensile test. This quantity is somewhat difficult
to define or measure exactly, especially where a significant amount of
deformation takes place after necking is apparently complete. However,
using the Gaussian equation the assumption can be made that the comple-
tion of the necking process occurs when the 'theoretical' engineering stress
rises to the original yield stress. The associated extension ratio will then be
an approximate measure of natural draw ratio, i.e.

and
(5.5)

where )'D is the natural draw ratio. A plot of this relation is given in Figure
5.30b.
It is believed that the differences between polymers that follow the
Gaussian and Langevin treatments respectively are real. Where the same
results have been studied by the two authors, as with Ward's results with
PMMA (Figure 5.9), the high molecular weight result& do not fit the
Gaussian equation but comply with the Langevin treatment while the low
molecular weight material is Gaussian. Polycarbonate which fits the Lan-
gevin equations has not been fitted to a Gaussian curve.
Because the strain hardening in tension of polymers which follow the
Gaussian relation may be represented by a single constant, it may prove of
particular interest to those concerned with making polymeric materials and
wish to obtain a measure of this important property (Patel et al., 1996).
268 The post-yield deformation of glassy polymers

5.7 THREE-DIMENSIONAL MODELLING OF LARGE STRAINS


In the preceding sections, the basic physics of post-yield deformation of
glassy polymers was presented. In particular, the phenomena of yield, strain
softening and strain hardening were discussed; the effects of rate, tempera-
ture, aging, molecular orientation and instability (necking and shear band-
ing) on these aspects of mechanical behaviour were detailed. A one-
dimensional model considering a viscous element in parallel with an entropy
hardening spring was also presented and shown to capture many of the
basic features of observed behaviour. The actual deformation of polymer
components and structures, whether simple (e.g. a tensile bar) or complex
(a soft drink bottle, a computer housing, an automobile bumper), are
three-dimensional and inhomogeneous in nature. In this section we utilize
much of what has been established or at least hypothesized concerning the
basic physics of amorphous polymer deformation, together with advanced
continuum mechanics formulation of finite strain kinematics, to develop and
implement a three-dimensional representation of large strain amorphous
polymer deformation. The model is then used together with numerical finite
element modelling to illustrate an example case of inhomogeneous deforma-
tion of polymers. This work draws on the research of Boyce and coworkers
over the past several years.
The three-dimensional constitutive model for glassy polymer deformation
essentially consists of three elements: (1) a linear elastic spring in series with
(2) a viscoplastic dashpot, and (3) a non-linear hardening spring, where
elements (2) and (3) act in parallel. The one-dimensional framework of this
three-element model is that of Haward and Thackray (H-T) discussed
earlier. H - T used an Eyring model to represent the viscous element and, as
presented in section 5.6, a Gaussian rubber-elastic model to represent the
hardening spring. (Note that in the original H-T model, a non-Gaussian
representation was used.) Boyce and coworkers have prescribed an equival-
ent three-dimensional representation for each of these elements. Each
element will be described below and will draw on similar concepts to
those of H-T and others as discussed earlier, but in some instances
represent significant advances in the features of deformation that can be
predicted. Below we first briefly give the constitutive representation for
three-dimensional deformation, which includes a thermally activated repre-
sentation of yield and a non-Gaussian network treatment of the strain
hardening behaviour; example homogeneous deformations are then studied
with the model and compared to experiments; the general finite strain
kinematics of large strain deformation are then reviewed; finally, an example
of the inhomogeneous deformation process of necking is studied using the
general three-dimensional representation together with the finite element
method.
Three-dimensional modelling of large strains 269

5.7.1 MATERIAL CONSTITUTIVE DESCRIPTION

The material constitutive representation can be broken down into the three
elements of linear elastic spring, a viscoplastic dashpot and a non-linear
hardening spring in a manner analogous to the H - Tone-dimensional
model. Below, we first provide the three-dimensional relations for these
elements in terms of principal components in order to show the three-
dimensional extension with minimal mathematical tensor complexity. The
three-dimensional element components are shown in the schematic of
Figure 5.25 where tensorial quantities of stress act on each element. In what
follows, standard indicial tensor notation will be followed where an upper-
case symbol with two indices refers to a second-order tensor, an uppercase
scripted symbol with four indices refers to a fourth-order tensor, repeated
indices imply summation, and the indices vary from 1 to 3 (three mutually
orthogonal directions). For a thorough description of tensor notation in
continuum mechanics see, for example, Malvern (1969).

5.7.2 ELASTIC BEHAVIOUR


The Cauchy stress tensor, T;j' acts on the linear elastic spring element as
defined through the conventional fourth-order modulus of elasticity

(5.6)

where T;j is the elastic strain tensor, J is the volume ratio and !e~jkl is the
elastic modulus tensor. The elasticity tensor can, in general, be anisotropic
and, indeed, in polymers the elastic properties will typically evolve from
isotropic to anisotropic during large scale straining due to the molecular
orientation of the macromolecular network. However, the anisotropy of the
elastic behaviour of amorphous polymers has been found to be small
(Wright et ai., 1971) in comparison to the anisotropy of the plastic
behaviour; here we will approximate the elasticity to remain isotropic.
Therefore,

(5.7)

where f.l is the elastic shear modulus, K is the bulk modulus and Dij is the
Kronecker delta (Dij = 1 when i = j, Dij = 0 when i =I j).

5.7.3 INELASTIC BEHAVIOUR


As in the H - T framework, the viscoplastic element acts in parallel with the
hardening element. Thus, part of the total Cauchy stress acts on the
270 The post-yield deformation of glassy polymers

viscoplastic element to drive plastic flow - this portion of the stress is termed
the driving stress tensor and is denoted by 7;;'. The remaining part of the
stress tensor develops due to the orientation of the underlying molecular
network and captures the anisotropic strain hardening behaviour of the
material - this portion is termed the network stress tensor and is denoted
Bij. Bij is taken to evolve with plastic strain assuming a non-Gaussian
statistical representation of the network, as will be described later.

5.7.4 VISCOPLASTIC BEHAVIOUR


The viscoplastic element is modelled assuming plastic flow to be a thermally
activated process. As discussed in earlier sections, several mechanisms of
inelastic flow have been postulated for glassy polymers. Here we consider
the model proposed by Argon where plastic flow is considered to result once
the intermolecular barriers to chain segment rotation have been overcome:

(5.8)

where Yo is the pre-exponential factor proportional to the attempt frequency,


AGo is the zero-stress level activation energy, 't is the effective equivalent
shear stress (the scalar equivalent of (7;j'), to be defined later), s is the
athermal shear strength, J1. is the elastic shear modulus, k is the Boltzmann
constant and 8 is absolute temperature (8 is now used for temperature in
order to avoid confusion with 7;j' the stress tensor). Note that equation 5.8
can be rearranged to give the shear flow stress as a function of strain rate
and temperature:

7: = s [ 1 - AGo (Y)J
k8 In y~ 6 5
/ (5.9)

Strain softening has been discussed in detail in section 5.3 and can be
considered to be a result of local dilatation accompanying shear which
produces a lowering of the barrier to deformation, i.e. a softening of the
material. It is also known to depend on the thermal history of the material
(physical aging; e.g. Hasan et al. 1993). Although strain softening is a result
of localized events, the effects of strain softening can be modelled in an
averaged sense. Here we model softening by taking the athermal shear
strength s to evolve to a preferred state sss:

S=h(1-~)YP
sss
(5.10)

where h is the softening slope. The existence of such a preferred state has
Three-dimensional modelling of large strains 271

been shown in the positron annihilation spectroscopy (PAS) and differential


scanning calorimetry (OSC) data of Hasan et al. (1993) on incrementally
deformed specimens, where both the PAS and OSC data found a steady
state profile to be reached after sufficient strain (about 25°/.)), regardless of
initial thermal history (section 5.3). The pressure dependence of yield can
also easily be incorporated into equation 5.8 by modifying the athermal
shear strength to be pressure dependent due to the well known pressure
dependence of the modulus (Boyce, Parks and Argon, 1988a).
The effective equivalent shear stress, acting on the viscoplastic element
is the scalar equivalent of the driving stress tensor T;;':
, = [-21T'J*'T*']1/2
J'
(5.11)
where T;;' is the tensorial difference between the total stress and the network
stress Bij (where Bij will be mathematically defined shortly). Thus the
three-dimensional description of the rate of plastic stretching, DL, is given by
Dfj = yPN ij
1 *, (5.12)
Nij = (2)1/2, T;j

where Nij defines the tensorial direction of the flow of the material.

5.7.5 STRAIN HARDENING BEHAVIOUR


The hardening element acts to model the second barrier to inelastic
straining, which evolves with strain due to the development of molecular
orientation. Once the intermolecular barriers to deformation have been
overcome, i.e. once the material has 'yielded', extensive plastic flow occurs
and the molecular chains begin to align or orient in a manner coincident
with the applied stretch state. The orientation of the underlying molecular
network will produce a bias in strength making this behaviour highly
anisotropic. Because of the anisotropic nature of orientation-induced har-
dening, characterization of strain hardening requires a three-dimensional
representation of the network and its stretching behaviour with strain.
Many rubber elastic network-type models have been proposed in the past,
such as the Flory and Rehner four-chain tetrahedron model and the Wang
and Guth three-chain cube. These and other models have been reviewed in
the Treloar text on rubber elasticity (1975). The various models have
assumed either or both Gaussian and non-Gaussian treatments of the chain
statistics. As discussed in section 5.4, while a Gaussian treatment captures
some basic features of polymer hardening, it has been found to be unable to
capture strain hardening behaviour at very large strains where a steep
272 The post-yield deformation of glassy polymers

Figure 5.31 Schematic of the eight-chain network of Arruda and Boyce showing the
network (a) in the undeformed configuration, (b) under uniaxial tension and (c)
under biaxial tension.

upturn in the stress-strain curve occurs. This behaviour is characteristic of


many commercially important glassy polymers such as polycarbonate,
poly(ethylene terephthalate), polystyrene and poly(methyl methacrylate),
among others, most probably due to their high concentration of 'physical
entanglements'. Thus, in this section we limit the discussion to non-Gaussian
treatments of chain stretching. Furthermore, Arruda and Boyce (1993a,b)
have found deficiencies in the three-chain cube and four-chain tetrahedron
network models with regard to their ability to predict three-dimensional
aspects of polymer deformation. Arruda and Boyce have proposed an
eight-chain network model (Figure 5.31) which will be utilized here to model
the strain hardening element. In the eight-chain network model, eight
non-Gaussian chains emanate from the centre of a cube, one to each corner
(Figure 5.31). The cube is taken to be oriented in principal stretch space and
the principal values of the network stress, termed Ri , are given by

(5.13)

where CR = nkE> is the rubbery modulus (equivalent to Gp in the non-


Gaussian formulation of section 5.4), n is the chain density where a chain is
defined as that portion of a chain between entanglements, N is the number
of 'rigid links' between physical entanglements, N are the principal
stretches, A~hain is the stretch on each chain in the network and is equal to
the root mean square of the applied stretches. The inverse Langevin function
is given by

2(f3) = cothf3 -~, f3 = 2- 1 (~~;;n)

and captures the limiting stretch behaviour of polymers where, as a chain


Three-dimensional modelling of large strains 273

stretch approaches its limiting stretch N 1 / 2 , the stress rapidly increases. As


will be demonstrated shortly, this eight-chain model has been found to
successfully capture the strain hardening behaviour of both glassy (Arruda
and Boyce 1991, 1993a) and rubbery polymers (Arruda and Boyce, 1993b)
in a variety of strain states. The model acts to simulate the essential features
of a more realistic, complicated random network where, in the eight-chain
model, the chains act to stretch and rotate towards the direction(s) of
largest stretch, as illustrated in Figures 5.31b and 5.31c for the cases of
uniaxial stretch and biaxial stretch. The success of the eight-chain network
model in capturing three-dimensional deformations has been demonstrated
by both Arruda and Boyce (1993a,b) and Wu and Van der Giessen (1993),
where Wu and Van der Giessen also compared the model to a full network
model.
The Arruda and Boyce eight-chain network model as stated in equation
5.13 acts to capture the three-dimensional large strain behaviour of strain
hardening, as will be demonstrated below; however, equation 5.13 does not
capture the temperature dependence of hardening that is observed in many
polymers. As has been discussed, molecular chain orientation evolves with
magnitude and state of strain to produce strain hardening in the polymer.
Molecular orientation also produces optical anisotropy in the polymer that
can be measured by birefringence. Indeed, Arruda and Przybylo (1995) have
successfully modelled the evolution of optical anisotropy using the eight-
chain network representation. The temperature dependenc(: of optical an-
isotropy has been measured and modelled by other researchers such as Raha
and Bowden (1972). Raha and Bowden found a temperature-dependent
optical anisotropy in PMMA specimens which had been plastically de-
formed via plane strain compression. They proposed a model which con-
siders the macromolecular network structure to evolve with deformation as
secondary valence interactions dissociate with strain. A thermally equilib-
rated number of these essentially weak entanglements was found to exist,
which accounts for the observed temperature dependent birefringence. Raha
and Bowden modelled the thermally equilibrated network to possess a
temperature-dependent chain density (where a chain is defined to be that
portion of a molecular chain between entanglements) of

no(0) = A' - B' [1 - exp ( - :0)] (5.14)

where no(0) is the chain density at absolute temperature 0, A' represents a


strong network that does not undergo thermal dissociation; the expression
containing B' represents the thermally evolving chain density, with Ea being
the thermal dissociation energy and R is the universal gas constant.
274 The post-yield deformation of glassy polymers

Kahar, Duckett and Ward (1978) found similar temperature dependence of


optical anisotropy in PMMA subsequent to extrusion at different tempera-
tures.
In addition to a temperature-dependent chain density, Boyce (1986) and
Arruda, Boyce and Jayachandran (1995) note that there must also exist a
temperature-dependent number of statistical 'rigid links' between entangle-
ments, N. If mass is conserved, nand N must evolve in a consistent manner:
n(0)N(0) = constant
Arruda, Boyce and Jayachandran (1995) then used these two concepts of a
temperature-dependent network structure and mass conservation to model
the temperature dependence of strain hardening. Strain hardening is still
modelled using the eight-chain network of equation 5.13, however nand N
now depend on temperature:

n(0) =B- Dexp(-~)


R0
(5.15)
n(0)N(0) = constant
This model predicts the chain density to decrease with an increase in
temperature and, consequently, the number of links between entangle-
ments to increase with an increase in temperature - thus modelling the
material to exhibit a softer strain hardening behaviour as the temperature
is increased. This model was found to capture successfully the temperature
dependence of strain hardening in both PMMA (Arruda, Boyce and
Jayachandran 1995) and PC (Arruda, 1992).

5.7.6 MODEL RESULTS FOR SIMPLE STATES OF DEFORMATION


Figure 5.32 shows both experimental and constitutive model results for the
cases of uniaxial compression and plane strain compression of polycarbon-
ate. The material constants for the model are fitted to the uniaxial compres-
sion data; the plane strain compression simulation is thus a true prediction.
Details for obtaining the material constants are given in Boyce, Parks and
Argon (1988a) and Arruda and Boyce (1993a). The model is found to
capture the uniaxial compression data and then found to predict the plane
strain compression data very well. This capability has also been shown on
other amorphous polymers. The predictive capability demonstrates that the
constitutive model contains the essential physics of the three-dimensional
anisotropic nature of the network deformation behaviour.
Figure 5.10 depicts the model prediction of the uniaxial compression
behaviour of pre oriented PC samples versus experimental results. The
material was preoriented by a uniaxial compressive logarithmic strain of
Three-dimensional modelling of large strains 275

200
I
-.,
<1l
a... !
~
r.n /
~
~
J
r.n 100 ,I
~ {)~,~
/'
~ j
J ........... uniaxial data
-- -- - - plane strain data
--- uniaxial eight chain
- - - plane strain eight chain

o
0.0 0.5 1.0 1.5
TRUE STRAIN
Figure 5.32 Model versus experimental results for the uniaxial and plane strain
compression behaviour of polycarbonate at a strain rate of - 0.01 s -1. The Arruda
and Boyce constitutive model was fitted to the uniaxial compression data and found
to be predictive of the plane strain behaviour.

- 0.57; samples were then cut and recompressed at different directions to the
initial compression. Note that the constitutive model accounts for network
orientation and the same material constants obtained from the uniaxial
compression on initially isotropic PC are used in these simulations on
preoriented Pc. Again, the model is found to predict the highly anisotropic
effects of preorientation, which act to either increase or decrease the yield
stress as well as increase or decrease the hardening slope, depending on the
direction of loading. The data illustrate the anisotropic nature of the
mechanical behaviour of oriented polymers and the simulation results
demonstrate the ability of the three-dimensional constitutive model to
predict this behaviour. Several other cases are given by Arruda, Boyce and
Quintus-Bosz (1993), where the ability of the model to predict the actual
highly anisotropic deformation paths as well as the stress--strain curves of
pre oriented samples is demonstrated.
Figure 5.33 depicts both experimental and model results for the tempera-
ture dependence of strain hardening in PMMA; the results show uniaxial
276 The post-yield deformation of glassy polymers

--
-
«l
Il.
:::e
t il
til
Q)
...,'"' 100
U1
Q)
:=
e-'"'

, .. ,.--_ .. -

0.5 1.0
True Strain
Figure 5.33 Model versus experimental results for the temperature-dependent
behaviour of poly(methyl methacrylate) subjected to uniaxial compression. The
Arruda and Boyce constitutive model was found to capture both the temperature
dependence of initial yield and strain hardening.

compression true stress versus true strain curves at different temperatures.


The simulations considered the temperature-dependent representation of the
network structure given in equations 5.15. Note that data are needed at two
temperatures to obtain the temperature-dependent material constants; the
third curve is a prediction. The strain hardening is found to be lower with
increasing temperature and is well predicted by the model. This same model
was used to simulate the rate dependence of PMMA, as shown in Figure
5.34 where both experimental and model results are given. The expected
increase in initial yield stress with increase in strain rate is observed and
predicted; however, note the greater amount of softening observed at the
higher strain rate. The highest rate ( - 0.1 s - 1) does not provide enough time
for heat transfer to occur and there is a rise in the material temperature
during straining which produces thermal softening in addition to strain
softening. The thermal softening is also predicted by the simulation, as
shown in Figure 5.34. This model considers the network deformation
Three-dimensional modelling of large strains 277

200
data -O.OOl/sec theory -O.OO1/sec

-_.::s
data -O.Ol/sec theory -O.Ol/sec
data -O.l/sec theory -O.l/sec
Ie
p..
~"'.
(Il
(Il
It:'~
: .....~.:.~
II)
.....s..
....~, ;:--
.~.~.:.---.
100
C/1
IIJ
::l
s..
Eo-

0.5 1.0
True Strain

Figure 5.34 Model versus experimental results for the strain rate dependence of
PMMA subjected to uniaxial compression. The initial yield stress was found to
increase with increasing strain rate; however, at high strain rates the high rate
response was found to be softer than that of the low rate response, due to thermal
softening. The Arruda and Boyce constitutive model results were able to capture
these effects in a predictive manner.

contributions to the total energy to be stored and not dissipated; a full


discussion is provided in Arruda, Boyce and layachandran (1995) as well as
Boyce, Montagut and Argon (1992); this approach to modelling storage
versus dissipative contributions to deformation as well as an alternative
approach were discussed earlier in this chapter. It is particularly important
to understand the high dependence of the stress-strain behaviour on rate
and temperature in cases of processing polymers and designing polymer
components for impact loading situations, where thermal softening can alter
the expected deformation response of the material. Walley et al. (1991) and
Walley and Field (1994) have experimentally examined very high rate
behaviour on a variety of polymers using Hopkinson bar tests, showing the
increase in yield with increase in rate as well as the thermal softening after
yield.
278 The post-yield deformation of glassy polymers

5.8 THREE-DIMENSIONAL KINEMATICS OF DEFORMATION


As indicated earlier, many glassy polymer deformation processes are not
homogeneous. Finite element analysis offers a methodology for examining
inhomogeneous deformation. Prior to using finite element analysis, the
constitutive representation of the material must first be generalized to
arbitrary three-dimensional deformations. Below we first provide a three-
dimensional generalization of the constitutive equations above (where
three-dimensional deformations were indeed examined, but only in a prin-
cipal stretch state) and then move on to examine inhomogeneous polymer
deformations using finite element analysis.
A general state of deformation of a material point can be described by the
deformation gradient,
(3x.
F .. = - '
'J ax.J
which maps a material point from the reference configuration into the
current configuration. Fij may be multiplicatively decomposed into elastic
and plastic components Fij = F~kF~j. Here, the relaxed configuration Ffj is
obtained by elastically unloading without rotation, i.e. F~j = F;i (see sche-
matic of Figure 5.35). Note that since F~j is taken to be symmetric, it
represents only stretch. F~ may be further decomposed into stretch and
rotation components:
(5.16)
For later reference, the eigenvalues of V;} are denoted Af.
In cases where either the material or the geometry exhibit a non-linear
behaviour, problems are solved in an incremental manner. Therefore, the
incremental quantities associated with deformation are important and are
typically expressed in terms of rate quantities. The velocity gradient
Lij = FikF~ 1 may be decomposed into its elastic and plastic components:

Lij = L ije + peLPpe-l


ik kl ij

Here,
(5.17)
where D~ is the rate of shape change of the relaxed configuration and lti}
is the corresponding spin. lti} is algebraically defined due to the kinematic
restriction on P~j (e.g. Boyce, Weber and Parks, 1989). The rate of shape
change, D~, must be constitutively defined as outlined below.
With the kinematic representation for arbitrary states of deformation in
hand, the material constitutive model described earlier can now be expressed
Three-dimensional kinematics of deformation 279

Ml~ ~'"
~ Fl,=~h)' ~ ~
"" (Fittl = (FhJ- T

F;',---~.J

§
Figure 5.35 Schematic illustrating the deformation gradient and its multiplicative
decomposition into elastic and plastic components.

in a general three-dimensional form. The Cauchy stress T;j acts on the linear
elastic spring as defined through the fourth-order tensor modulus of
elasticity 2~jkl:

(5.18)

where J = det F~j is the volume change and In F:, is the logarithmic strain
often referred to as the elastic Hencky strain. As stated earlier, in the model,
2~jkl is taken to remain isotropic since the developing anisotropy of
elasticity is small compared to the anisotropy of the plastic behaviour in
amorphous polymers. However, one could easily envision having this tensor
evolve with strain due to molecular orientation.
The rate of plastic straining is described by the plastic stretching tensor
DPij'•
280 The post-yield deformation of glassy polymers

where

1 *,
Nij = (2r)1/2 Ii j (S.19)

Here yP is the scalar expression of the plastic strain rate described in detail
earlier; Nij describes the three-dimensional tensorial direction of plastic
straining which is coaxial with the driving stress tensor Ii1'- Ii1' is the stress
acting on the viscoplastic element which is essentially the difference between
the Cauchy stress Ii j and the network stress Bij (where the network stress
must be converted to the same configuration as the Cauchy stress):

(S.20)

and r is the effective equivalent shear stress:


r = (-21 T*'T*')1/2
IJ JI (S.21)
The general expression for the network stress, B ij , is found by first obtaining
the eigenvalues (An of the plastic deformation gradient (l';}) via an
eigenvalue decomposition.
(S.22)
The eigenvalues of the network stress, Bi , are then found using the
eight-chain model of rubber elasticity:

B. = C N 1/ 2 2-1 (A~hain) Ai 2 - A~~ain (S.23)


I R N 1/ 2 AP.
cham

The network stress is then obtained by rotating the eigenvalues to the


configuration represented by the plastic deformation gradient:
(S.24)
Equations S.16-S.24 together with equation S.8 constitute a complete
three-dimensional description of the finite strain elastic-viscoplastic behav-
iour of glassy polymers. This general description is needed in order to
describe cases involving arbitrary deformation.

5.9 NUMERICAL SIMULAnON OF INHOMOGENEOUS


DEFORMAnON
Post-yield behaviour of polymeric materials, their products and their
processes typically involves inhomogeneous deformation - the entire speci-
men or product is not deforming in a uniform, homogeneous manner. The
Numerical simulation of inhomogeneous deformation 281

classic example is the necking behaviour of many polymers during a tension


test. Such inhomogeneous deformation can now be analysed in detail using
accurate constitutive models of the material stress- strain behaviour such as
the one described in this chapter, together with numerical techniques such
as the finite element method.
No attempt to describe the finite element method will be given here due
to the mathematical detail required to do it justice. Numerous textbooks are
available on the subject such as Cook, Malkus and Plesha (1989), Bathe
(1982) and Hughes (1987) to which the reader is referred fo r detail regarding
formulation of the finite element method. In a finite element analysis, the
geometry of the problem is discretized into numerous 'elements'; each
element contains an approximation or interpolation of a particular field
(typically the displacement field) over the element from which other quan-
tities (for example, strain and stress fields) are calculated: calculations are
executed such that compatibility is maintained and equilibrium is satisfied
in an approximate manner using either a work or energy minimization
principle. Advances in the finite element method, particularly aspects
involving non-linear deformation, have been numerous over the past 20
years. These developments in the finite element method now permit the
analysis of complex deformation problems involving non-linear and evolv-
ing deformations, geometries, material behaviour and boundary conditions.

Figure 5.36 Finite element mesh of a polymer cylinder which will be subjected to
tension (elapsed time 0 s) Note that axisymmetry was used in the model and the
figure shows both the mesh and its mirror image about the axis of symmetry.
282 The post-yield deformation of glassy polymers

(a)

(b)

(c)

(d)

Figure 5.37 (a-h) Model versus experimental results for the necking of polycarbon-
ate. Elapsed times (a) 343 s, (b) 443 s, (c) 721 s, (d) 1145 s, (e) 1490 s, (f) 2100 s,
(g) 2400 s, (h) 3046 s. Calculations of the deformed geometry are depicted for
comparison with the experimentally observed deformation; neck initiation, stabiliz-
ation and propagation are shown .
Numerical simulation of inhomogeneous deformation 283

(e)

(f)

(g)

(h)

Figure 5.37 (Continued)


284 The post-yield deformation of glassy polymers

Below, we illustrate the power of this technique, using the example of the
necking behaviour of polycarbonate (Hasan, 1994).
The tensile necking phenomena was described earlier in section 5.15.
Here, we numerically simulate the necking of a cylindrical specimen of
polycarbonate during a tensile test and compare the simulation results to
experiments. The simulations and experiments are taken from Hasan (1994).
The simulation of thermomechanical effects in necking has been discussed
in Boyce, Montagut and Argon (1992). In the polycarbonate example to be
shown here, the constitutive model of the material stress-strain behaviour
is essentially the model described in the previous section where, however, a
distributed Eyring model of the yield process is used to model initial yield
instead of the Argon model. Further details of this aspect regarding
modelling the initial yield of the material are provided in Hasan (1994) and
Hasan and Boyce (1995). Also, operational details of finite element model-
ling of necking of polymers are discussed in detail in Boyce, Montagut and
Argon (1992). The finite element model of the specimen is depicted in Figure
5.36 (axisymmetric elements were utilized taking advantage of the symmetry
of the problem which permits modelling of only one-half of the specimen
cross-section; Figure 5.36 depicts both the finite element mesh and its
reflection in order to visualize the entire specimen). Both the specimen
and the model contain an imperfection at the same location (slightly above
the specimen centre) in order to predefine the neck initiation location. The
specimen model is then loaded in tension by monotonically increasing the
displacements of the specimen ends.
Figures 5.37a-h depict results of the evolution in the specimen geometry
with deformation, calculated using the finite element method as well as
photographs of the experimentally deformed specimen at the same stages of
deformation. Note that the constitutive model parameters used in the
simulation were obtained by fitting data obtained in uniaxial compression.
Therefore, the tensile test simulation results are true predictions (no data
obtained in the tensile test were used to calculate the tensile test behaviour).
Excellent agreement is obtained between the numerical prediction of the
tensile specimen geometry and that actually observed in the experiment -
note the initiation of the neck, the stabilization of the neck and the steady
state propagation of the neck. Also note the steady state profile of the
shoulder region of the neck as the neck is propagating.
The evolution in the neck with deformation can also be described by
examining the contours of equivalent plastic strain rate. The equivalent
plastic strain rate yP is the quantity which indicates where active plastic
deformation is taking place - i.e. the regions of material that are currently
deforming. Figures 5.38a-h depict contours of yP for the stages of deforma-
tion corresponding to those shown in Figures 5.37a-h. Initially (a) all
Numerical simulation of inhomogeneous deformation 285

deformation is occuring at the location where the neck initiated, indicated


by yP being highest in this region; indeed material away from the zone is not
plastically deforming yet. As the material in the neck begins to strain
harden, the zone of plastic deformation begins to spread out from the
necking region (b). Once the necked region has sufficiently hardened, it
becomes easier to deform material outside the neck and the neck then ceases
to deform plastically, thus stabilizing. This is shown in (c) where the zones
of active plastic deformation are now travelling up and down away from the
location of initial neck formation. In the deformed profiles of Figure 5.37,
the shoulder region of the neck took on a steady state profile; in the
contours of plastic strain rate, a steady state profile in strain rate is also
observed in the shoulder region. This is also shown in Figure 5.39, which
plots strain rate as a function of axial position along the neck as one
transitions from un-necked to actively necking to fully necked regions. This
neck stabilization and propagation phenomena continues as shown in (d-h)
where, as the material in the actively deforming region hardens, the regions
of active deformation propagate until the entire specimen has necked and
strain hardened. At this point, the specimen is again of uniform diameter.
The curve calculated in Figure 5.39 is similar to that measured on PC by
Nazarenko et al. (1994). Its form is characteristic of the behaviour of many
thermoplastics under constant load. For linear polyethylene similar curves
have been measured by Coates and Ward (1980), and can also be calculated
from the one-dimensional Gaussian equation (Haward, 1995).
The global specimen load -displacement curve for the specimen is shown
in Figure 5.40, where both experiment and finite element prediction are
provided. The finite element calculation is shown to provide an excellent
prediction of the observed load versus displacement curve, showing the
initial increase in load followed by a dramatic decrease upon neck initiation.
This is followed by a load plateau as the neck propagates and finaIIy a
gradual increase in load once the entire specimen has necked and the
material continues to strain harden. The drop in load upon neck initiation
is a result of the geometric instability, whereupon the initiation of a local
decrease in specimen diameter leads to the initial decrease in load-carrying
capability of the specimen; the magnitude of the load drop is dependent
upon several factors such as physical aging of the material (the more aged
the polymer, the greater the load drop; see section 5.3 on strain softening),
the natural draw ratio (the larger the natural draw ratio, the greater the load
drop since necking will not stabilize until sufficient strain hardening has
occurred in the necking region) and thermal softening effects (at higher rates,
heating of the polymer will occur due to plastic dissipation; the heating
produces a local thermal softening which further increases the load drop and
can also lead to the oscillatory effect on load displacement discussed in
286 The post-yield deformation of glassy polymers

D'l' VNoAa 1IWt' VAUI'

...,...
-l .U lJ . 1 . tll·1 1
• 1. U I·N . 1. tt ... ...
. J . ..... ..
. t U I,.O. . t .•'I·'"
_l .UI-.' .'.n."
. 1"
. ." .•. n....
· 1, ' I) - 1.UI- "
_:1 .2." ., u.~ . :

. 1. ......
• 2 •• ....
' ..
. :1 • • , ... . , . 1 . nl,. 1
• ) .14.1-1' ·1 u.-•
- 1 . U I,.., -:a . 1 , .... ~
• J . 'f~" _:l .n ••
. ' . I U >aO' _::I . IU'_ '

(a) Elapsed Time 343s (b) Elapsed Time 443s

-..
.,
......
IU~ U
.,."" YAUII
.s ...... u
*1 'U- t4
·I .n .....
., UI- ., I..tl_"

.• u . .. . ·I. u ......

..."".......
. , ••

.,
I' . ., 11
.',"1-
_1 _OtI _.1
."

·l , ltt-OJ *l . :I1.. U

·1 '.1' -, ",1,. 1)
-I 118- 11
.1 t . . . . ,
-112' ..... ,
-1.'U"" .J IJ
·' ."1-" ., I I..... '
.1 U .... ,
· 1 ...... ) ·1IU- 11I

(e) Elapsed Time 721s (d) Elapsed Time 1145s

Figure 5.38 (a-h) Model results showing contours of the plastic strain rate at
deformations shown in Figure 5.37 [with the exclusion of part (f)). The plastic strain
rate indicates where the material is actively deforming at that instant, another
indication of the neck initiation, stabilization and propagation process.
Numerical simulation of inhomogeneous deformation 287

-" ......
.... nl-u
.".", .......
. ] . Uz-u
01.10.,.06 ·1 .".·04
-' . '11- 0 4 " . 411- 04
. 5 . 41 ... 04- ,' . 1,.·04
''' .n''''4 · 6 , "&-
.' •• J..... 4 " ."1'-
. 1.0,,70' . 1. Ut-OJ
"l ." ... t ) '1 . 111-0)
·l. UI· O) · 1.l,.,·O)
-1 . '2",0) ol , UI- O'
- 1.101-0' .1.".0'
.1 " .... 0) - •. t1l-0 J
.l . tU-OJ _'J . UI_OJ
., UI--U .l . nl-O)

(e) Elapsed Time 14905 (I) Elapsed Time 18005

.n. .....
-1 .'''''14
-1 . 711-0"
_. .....,.
. , . , OJ.-U

• '.Ue-04 ·1 . • '.....
• 1.311-0. 'J ,'''-' •
• ., . HI . . .. '4 . "41-t ..
' ' . It • • G..
.f,'"_,,,
-1 . M ... , .t.:"',-."
'\.11,·.,
., 401· .' ' 1 U .... ,
., ",...,
.......,
.J nt· O,
·1 .1,..·.' ·1 , ...... '

ol . tll-"
·1 Ill-OJ
.2 . U ..·.'
.:r .n ... ., . , ""-0)
,' . 14.1-0)

~l

(Q) Elapsed Time 21005 (h) Elapsed Time 24005

Figure 5.38 (Continued)


288 The post-yield deformation of glassy polymers

0.003 .---~---'-------r----,---.---,

CJ)
+'
oj 0.002
~
I=l
.~
r....
+'
(f)

.S
+'
rn
oj 0.001
0::

0.000 L-L-----'--_ _..L-_--'_ _.....L.:>...-_.L......._---1


o 2 3
Axial Distance (LID)
Figure 5.39 Line plot of magnitude of plastic strain rate versus axial position in the
necking polymer.

section 5.4.6). Once the initial neck is fully developed, neck propagation
occurs at a relatively constant load. At the point where the neck has
traversed the entire specimen length, the load increases due to further strain
hardening in the entire specimen until final fracture.
These simulations and comparisons with experiments show the ability to
model complex deformation fields using an advanced, physically based
constitutive model together with the finite element method. Such simula-
tions can predict both the local details of the strain, stress and temperature
fields in a body as well as global loads and displacements. Other problems
have also been examined using the constitutive model of Boyce and
coworkers together with the finite element method; problems include the
simple shear experiments of G'Sell (Boyce, Arruda and Jayachandran, 1994),
necking of a polymer sheet (Hasan), plane strain necking (Wu and Giessen),
hydrostatic extrusion of polymer cylinders (Boyce, Parks and Argon,
1988b), thermomechanically coupled deformations (Boyce, Montagut and
Argon, 1992; Arruda, Boyce and Jayachandran, 1995) as well as indentation
of single and multilayered polymeric coatings (Jayachandran, Boyce and
Argon, 1993, 1995a,b). Such detailed simulations provide a better under-
standing of specimen, product and process behaviour and can be utilized in
the design and development of multiphase polymeric materials, polymeric
products and polymer processing operations.
References 289

4
,,,
~.~_&'w~II\;;;;;;;;.:-::ii;:;;;:-

1 "------"'------"---'-----'------'------'----'-----'
o 10 20 30 40
DISPLACEMENT (mm)

Figure 5.40 Global load versus displacement curve as computed from the finite
element analysis (solid line) compared to that obtained in the exoeriment (dashed
line). Excellent agreement is observed. Note that the finite element model used the
constitutive model parameters obtained in compression, thus demonstrating the
predictive nature of the constitutive model.

REFERENCES
Adam, G.A., Cross, A. and Haward, R.N. (1975) J. Mater. Sci., 10, 1582.
Adam, G.A., Hay, IN., Parsons, I.W. and Haward, R.N. (1976) Polymer, 17, 57.
Adams, G.W. and Farris, R.J. (1988) J. Polym. Sci.: Phys, 26, 433.
Aharoni, S.M. (1983) Macromolecules, 16, 1722.
Aharoni, S.M. (1986) Macromolecules, 19, 426.
Alfrey, T. (1948) The Mechanical Behaviour of Polymers, Interscience, New York.
Allen, G., Morley, D.C.w. and Williams, T. (1973) J. Mater. Sci. 8, 1449.
Allison, S.W. and Ward, I.M. (1967) 1. Appl. Phys., 18, 1151.
Andrianova, G.P., Kechekyan, A.S. and Kargin, V.A. (1971) J. Polym. Sci. A2, 9,
1919.
Aref-Azar, A., Biddlestone, F., Hay, IN. and Haward, R.N. (1983) Polymer, 24, 1245.
Argon, A.S. (1973a) Phil. Mag., 28, 39.
Argon, A.S. (1973b) J. Macromol. Sci. B: Phys., 38(4),573.
Argon, A.S. and Bessonov, M.l. (1977) Polym. Eng. Sci., 17,174.
290 The post-yield deformation of glassy polymers

Arruda, E.M. (1992) PhD Thesis, Massachusetts Institute of Technology, Depart-


ment of Mechanical Engineering, Cambridge, MA.
Arruda, E.M. and Boyce, M.e. (1991) in Anisotropy and Localization of Plastic
Deformation, Proceedings of Plasticity 91 (ed. A.S. Khan), Grenoble, France.
Arruda, E.M. and Boyce, M.e. (1993a) Int. J. Plasticity, 9, 697.
Arruda, E.M. and Boyce, M.e. (1993b) J. Mech. Phys. Solids, 41, 389.
Arruda, E.M., Boyce, M.e. and Jayachandran, R. (1995) Mech. of Mater., 19, 193.
Arruda, E.M. and Przybylo, P. (1995) in Polym. Eng. Sci., 35, 395.
Arruda, E.M., Boyce, M.e. and Quintus-Bosz, H. (1993) Int. J. Plasticity, 9, 783.
Barenblatt, G.I. (1970) Mech. Solids,S, 110.
Bartos, 1., Muller, J. and Wendorf, 1.H. (1990) Polymer, 31, 1679.
Bathe, K.1. (1982) Finite Element Procedures in Engineering Analysis, Prentice-Hall,
N1.
Bauwens, J.-e. (1978) J. Mater. Sci., 13, 1443.
Bauwens, 1.-e., Bauwens-Crowet, e. and Homes, G.A. (1969) J. Polym. Sci. A2, 7,
1745.
Bender, M.F. and Williams, M.L. (1963) J. Appl. Phys., 34, 3329.
Bessonov, M.I., Koton, M.M., Kudryavstev, V.V. and Laius, L. (1987) in Thermally
Stable Polymers. Polyimides (ed. W.W. Wright), Consultant Bureau, New York
and London, pp. 238, 242.
Biddlestone, F., Kemish, D.1., Hay, 1.N. et al. (1986) Polym. Testing, 6, 163.
Biglione, G., Baer, E. and Radcliffe, S.V. (1969) Proc. 2nd Conference on Fracture,
Brighton, England, April 1969, Paper 44/l.
Binder, G. and Muller, F.H. (1961) Kolloid Z., 178, 129.
Bowden, P.B. (1973) in The Physics of Glassy Polymers (ed. R.N. Haward) Applied
Science Pub!., London, Ch. 5.
Bowden, P.B. and Jukes, 1.A. (1972) J. Mater. Sci., 7, 52.
Bowden, P.B. and Raha, S. (1970) Phil. Mag., 22, 463.
Boyce, M.e. (1986), PhD Thesis, Massachusetts Institute of Technology, Depart-
ment of Mechanical Engineering, Cambridge, MA.
Boyce, M.e. and Arruda, E.M. (1990) Polym. Eng. and Sci., 30, 1288.
Boyce, M.e. and Arruda, E.M. (1993) J. Mech. Phys. Solids, 41, 389.
Boyce, M.e., Arruda, E.M. and Jayachandran, R. (1994) Polym. Eng. Sci., 34, 716.
Boyce, M.C., Montagut, E.L. and Argon, A.S. (1992) Polym. Eng. Sci. 32, 1073.
Boyce, M.e., Parks, D.M. and Argon, A.S. (1988a) Mech. Mater., 7, 15.
Boyce, M.e., Parks, D.M. and Argon, A.S. (1988b) Mech. Mater., 7,35.
Boyce, M.e., Weber, G.G. and Parks, D.M. (1989) J. Mech. Phys. Solids, 37, 647.
Brady, T.E. and Yeh, G.S.Y. (1973) J. Mater. Sci., 8, 1083.
Broutman, L.1. and Patil, R.S. (1971) Polym. Eng. and Sci., 11, 165.
Brown, N. and Ward, I.M. (1968) J. Polym. Sci. A2, 6, 607.
Bulatov, V. and Argon, A.S. (1995) Mater. Sci. Eng., 2, 203.
Carfagna, e., Amendola, E., D'Amore, A. and Nicolais, L. (1988) Polym. Eng. Sci.,
28,1203.
Chang, B.T. and Li, 1.e.M. (1988) Polym. Eng. Sci., 28, 1198.
Cheng, T.W., Keskkula, H. and Paul, D.R. (1992) J. Appl. Polym. Sci., 45, 53l.
Coates, P.D. and Ward, I.M. (1980) J. Mater. Sci., 15,2897.
Considere, M. (1888) Die Anwendung von Eisen und Stahl bei Constructionen, Verlag
von Carl Gerold's Sohn, Vienna.
Cook, R.D., Malkus, D.S. and Plesha, M.E. (1989) Concepts and Applications of
Finite Element Analysis, John Wiley, New York.
Comes, P.L. and Haward, R.N. (1974) Polymer, 15, 149.
Cross, A. and Haward, R.N. (1973) J. Polym. Sci.: Phys., 11, 2423.
References 291

Cross, A. and Haward, R.N. (1978) Polymer, 19, 677.


Cross, A. and Haward, R.N. (1980) Polymer, 21,1226.
Cross, A., Hall, M. and Haward, R.N. (1975) Nature, 253, 340.
Cross, A., Haward, R.N. and Mills, N.J. (1979) Polymer, 20, 288.
Dickinson, IT., Jensen, L.e., Langford, S.e. and Dion, R.P. (1994) J. Polym. Sci.:
Phys., 32, 779.
Donald, A.M. and Kramer, E.J. (1982) J. Polym. Sci.: Phys., 20, 899.
Ender, D.H. (1968) J. Appl. Phys., 39, 487.
Ender, D.H. and Andrews, R.D. (1965) J. App. Phys., 36, 3057.
Erman, B. and Mark, J.E. (1994) The Science and Technology of Rubber (eds B.
Erman, lE. Mark and F.R. Eirich), Academic Press, New York and London,
p.207.
Eyring, H.J. (\936) J. Chem. Phys., 4, 283.
Ferry, lD. (1980) Viscoelastic Properties of Polymers, 3rd edn, John Wiley, New
York.
Flory, P.J. (1949) J. Chem. Phys., 17, 303.
Flory, P.J. and Yoon, D.Y. (1978) Nature, 272,226.
Fuller, K.N.G., Fox, P.G. and Field, lE. (1975) Proc. R., Soc., 341, 537.
Gent, A.N. and Madan, S. (1989) J. Polym. Sci. B: Phys., 27, 1529.
Gilmour, I.W., Trainor, A. and Haward, R.N. (1978a) J. Polym. Sci.: Phys., 16, 1277.
Gilmour, I.W. Trainor, A. and Haward, R.N. (l978b) J. Polym. Sci.: Phys., 16, 1291.
Godovsky, y.K. (1992) Thermophysical Properties of Polymers, Springer-Verlag.
Berlin, Heidelberg, New York.
Golden, lH., Hammant, B.L. and Hazell, E.A. (1967) J. Appl. Polvm. Sci., 11, 1571.
Griffith, A.A. (1920) Phil. Trans. R. Soc. London A, 221, 163.
G'Sell, e. (1982) Plastic Deformation of Amorphous and Semi-Crystalline Polymers,
Centre Physique des Houches, Les Houches, France, p. 375.
G'Sell, e. and Gopez, A.J. (1985) J. Mater. Sci., 20, 3462.
GSell, e. and Jonas, J.J. (1979) J. Mater. Sci., 14, 583.
GSell, e. and Jonas, J.1. (1981) J. Mater. Sci., 16,1956.
G'Sell, e., Hiver, lM., Dahouin, A. and Souahi, A. (1992) J. Mater. Sci., 27, 5031.
Gurevich, G. and Kobeko, P. (1940) Rubber Chem. Tech., 13, 904.
Hasan, O.A. (1994) PhD Thesis, Massachusetts Institute of Technology, Department
of Mechanical Engineering, Cambridge, MA.
Hasan, O.A. and Boyce, M.e. (1993) Polymer, 34, 5085.
Hasan, O.A. and Boyce, M.e. (1995) Polym. Eng. Sci., 35,331.
Hasan, O.A., Boyce, M.e., Li, K.S. and Berko, S. (1993) J. Polym. Sci.: Phys., 31, 185.
Haward, R.N. (1942) Trans. Faraday Soc., 38, 391.
Haward, R.N. (1949) The Strength of Plastics and Glass, Cleaver-Hume Press,
London, p. 103.
Haward, R.N. (ed.) (1973) The Physics of Glassy Polymers, Applied Science Pub-
lishers, London.
Haward, R.N. (1980) Coli. Polym. Sci., 258, 643.
Haward, R.N. (1993) Macromolecules, 26,5860.
Haward, R.N. (1995) J. Polym. Sci.: Phys., 33, 1481.
Haward, R.N. and Brough, 1. (1969) Polymer, 10, 724.
Haward, R.N. and Thackray, G. (1968) Proc. R. Soc. London A, 302, 453.
Haward, R.N., Daniels, H.E. and Treloar, L.R.G. (1978) J. Polym. Sci.: Phys., 16,
1169.
Hoff, E.A.w. (1952) J. Appl. Hem., 2, 441.
Holliday, L., Mann, 1., Pogany, G.A. et al. (1964) Nature, 202,381.
Hookway, D.e. (1958) J. Text. Institute, 49, 292.
292 The post-yield deformation of glassy polymers

Hope, P.S., Ward, I.M. and Gibson, A.G. (1980) J. Mater. Sci., 15, 2207.
Horsley, RA. (1958) Plastics Progress (ed. P. Morgan), Iliffe and Sons, London and
New York, p. 77.
Hughes, T.J.R. (1987) The Finite Element Method, Prentice-Hall, Nl
Ito, E., Yamamoto, K., Koboyashi, Y and Hatakeyama, T. (1978) Polymer, 19, 39.
layachandran, R., Boyce, M.e. and Argon, AS. (1993) Adhesion Sci. Technol., 7,813.
layachandran, R, Boyce, M.e. and Argon, AS. (1995a) J. Computer Aided Mater.
Design, 2, 23.
layachandran, R, Boyce, M.e. and Argon, AS. (1995b) J. Computer Aided Mater.
Design, 2, 151.
louIe, l (1857) Phil. Mag., 14, 226.
Kahar, N., Duckett, R.A. Ward, I.M. (1978) Polymer, 19, 136.
Kinloch, A.J. and Young, R.J. (1983) Fracture Behaviour of Polymers, Elsevier
Applied Science, London and New York, p. 132.
Kirste, RG., Kruse, W.A. and Schelten, l (1973) Makromol. Chem., 162, 297.
Koenen, lA. (1994) Thermochim. Acta, 247, 55.
Koenig, lL. and Antoon, M.L. (1977) J. Polym. Sci.: Phys., 15, 1379.
Kramer, E.l. (1978) J. Mater. Sci., 14, 1384.
Kramer, E.J. and Berger, L.L. (1990) Adv. Polym. Sci., 91-92,1.
Lazurkin, Yu.S. and Fogel'son, R.A. (1951) Zhur. Tech. Phys. USSR, 3, 21, 267.
Lee, B.J., Argon, A.S., Parks, D.M. et al. (1993) Polymer, 34, 3555.
Lee, D. and Luken, P.e. (1986) Polym. Eng. Sci., 26, 812.
Lee, H.H.D. and McGarry, F.l. (1991) J. Macromol. Sci. B: Phys. 30, 185.
Long, S.D. and Ward, I.M. (1991) J. Appl. Polym. Sci., 42, 1991.
Maher, lW., Haward, R.N. and Hay, IN. (1980) J. Polym. Sci.: Phys., 18, 2169.
Malvern, L.E. (1969) Introduction to the Mechanics of Continuous Medium, Prentice-
Hall, Nl
Marshall, A.S. and Petrie, S.E.B. (1975) J. Appl. Phys., 46, 4223.
Marshall, l and Thompson, AB. (1954) Proc. R. Soc. London A 221, 541.
McCormick, H.W., Brower, F.M. and Kim, L. (1959) J. Polym. Sci., 39, 87.
McKechnie, 1.I., Haward, R.N., Brown, D. and Clarke, 1.H.R. (1993) Macro-
molecules, 26, 198.
Meinel, G. and Peteriin, A (1971a) J. Polym. Sci.: Phys., 9, 67.
Meinel, G. and Peteriin, A. (1971b) Eur. Polym. J., 7, 657.
Meyer, K.H. and Van der Wyk, lA. (1946) J. Polym. Res., 1,49.
Mills, N.l. (1975) Molecular Behaviour and the Development of Polymeric Materials
(eds A Ledwith and A.M. North), Chapman & Hall, London, p. 443.
Mills, P.J., Hay, IN. and Haward, RN. (1985) J. Mater. Sci., 20, 501.
Mininni, M., Moore, RS., Flick, 1.R. and Petrie, S.E.B. (1973) Macromol. Sci. B:
Phys. 8, 343.
Moore, R.S., O'Sloane, lK. and Shearer, le. (1981) Polym. Eng. and Sci., 21, 203.
Mott, P.H., Argon, AS. and Suter, U.W. (1993) Phil. Mag. A, 67, 931.
Murphy, B.M., Haward, RN. and White E.F.T. (1971) J. Polym. Sci. A2, 9, 801.
Nadai, A. (1950) Theory of Flow and Fracture of Solids, 2nd edn, McGraw-Hill, New
York.
Nazarenko, S., Bensason, S., Hiltner, A and Baer, E. (1994) Polymer, 35, 3883.
Nazarenko, S.I. (1988) Candidate Dissertation, Institute of Chern. Phys., USSR
Academy of Sciences, Moscow (see Godovsky, Y).
Oleynik, E.F. (1989) Prog. Coli. Polym. Sci., 80, 140.
Oleynik, E.F. (1990) High Performance Polymers (eds E. Baer and S. Moet),
Munchen, Ch. 4, pp. 80-101.
Patel, RM., Sehanobish, K., lain, P. et al. (1996) J. Appl. Polym. Sci., 60, 749.
References 293

Porter, R.I. and Johnson, 1.F. (1966) Chem. Rev., 66, l.


Raha, S. and Bowden, P.B. (1972) Polymer, 13, 175.
Reddish, W. (1966) J. Polym. Sci., 14, 123.
Regel, B.P. (1951) J. Tech. Phys. USSR, 21, 287.
Rehage, G. and Goldbach, G. (1967) Angew. Makromol. Chem., 1, 125.
Richards, R.C. and Kramer, E.I. (1972) J. Macromol. Sci.: Phys. B 6, 242.
Rivlin, T. (1948) Phil. Trans. Roy. Soc. London A 240,454.
Robertson, R.E. (1963) J. Appl. Polym. Sci., 7, 443.
Robertson, R.E. (1966) J. Chem. Phys., 44, 3950.
Robertson, R.E. (1968) Appl. Polym. Sci. Polym. Symp, 201.
Robertson, R.E. (1976) Adv. Chem. Series, 154, ACS, New York, p. 89.
Roe, R.-1. and Millman, G.M. (1983) Polym. Eng. Sci., 23, 318.
Roe, R.-1. and Tonelli, A.E. (1978) Macromolecules, 11, 114.
Rudnev, S.N., Salamatina, O.B., Voenniy, V.V. and Oleynik, E.F. (1991) ColI. Polym.
Sci., 269, 460.
Schelten, 1., Ballard, D.G.H., Wignall, G.D. et al. (1976) Polymer, 17, 75l.
Snyder, 1., Hiltner, A. and Baer, E. (1994) J. Appl. Polym. Sci., 52, 217.
Sternstein, S.S., Ongchin, L. and Silverman, A. (1968) Appl. Polym. Symp., 7, 175.
Stolting, J. and Muller, F.H. (1970) Kolloid Z. 238, 459; 240, 792.
Struik, L.C.E. (1978) The Physical Ageing of Amorphous Polymers and other
Materials, Elsevier, New York and Amsterdam.
Thomson, W. (1853) Trans. R. Soc. Edinburgh, 20, 26l.
Toda, A. (1993) Polymer, 34, 2306.
Trainor, A., Haward, R.N. and Hay, 1.N. (1977) J. Polym. Sci.: Phys., 15, 1077.
Treloar, L.R.G. (1958) The Physics of Rubber Elasticity, 2nd edn, Oxford University
Press, Oxford.
Treloar, L.R.G. (1975) The Physics of Rubber Elasticity, 3rd edn, Clarendon Press,
Oxford, UK.
Tuckett, R.F. (1943) Chem. Ind., 62, 430.
Vincent, P.I. (1960) Polymer, 1, 7.
Vincent, P.I. (1974) Deformation and Fracture of High Polymers (ed. H.H. Kausch),
Plenum Press, New York, p. 287.
Walley, S.M. and Field, J.E. (1994), DYMAT J., 1,211.
Walley, S.M., Field, 1.E., Pope, P.H. and Safford, N.A. (1991) J. Phys. III, France,
1, 1889.
Ward, I.M. (1984) The Mechanical Properties of Solid Polymers, 2nd edn, John
Wiley, Chichester.
Wignall, G.D., Schelten, 1. and Ballard, G.D.H. (1973) 3rd Intern. Conf. X-ray and
Neutron Scattering.
Williams, 1.G. (1967) Trans. J. Plast. Inst., June, 205.
Wright, M., Faraday, C.S.N., White, E.F.T. and Treloar, L.R.G. (1971) J. Phys. D, 4,
2002.
Wu, P.O. and Van der Giessen, E. (1993) J. Mech. Phys. Solids, 41, 427.
Wu, P.O. and Van der Giessen, E. (1995) Int. J. Plasticity, 11, 211.
Yo on, D.Y. (1977) 4th Intern. Conf. Small Angle Scattering of X-rays and Neutrons,
Gothenburg, Tennessee, USA.
Yoon, D.Y. and Flory, P.I. (1977) Polymer, 18, 509.
Zurimendi, 1., Biddlestone, F., Hay, 1.N. and Haward, R.N. (1982) J. Mater. Sci., 17,
199.
Crazing 6
A.M. Donald

6.1 INTRODUCTION
In the 20 years or more that have passed since the first edition of this book
was published, our understanding of the phenomenon of crazing has made
substantial progress. We have moved from a phenomenological picture of
when it occurs - in particular with regard to the macroscopic stress state -
to a good understanding at a molecular and mechanistic level of what
governs the crazing process. The advantage is that some degree of predictive
power comes with this understanding, and strategies of improving the
toughness of glassy polymers (even without considering the addition of
second-phase particles as in rubber-toughened systems) can be proposed. In
the first edition of this book, Andrews stated The importance of crazing
in the study of fracture in glassy polymers cannot be overstated.' This
statement is still true. If we are to understand what gives one glassy polymer
a high fracture toughness, but another a low one, we must understand the
formation of crazes which precede the advancing crack front. One of the
recent key advances in understanding the molecular mechanisms which
underlie crazing has been the ability to appreciate why different glassy
polymers have different responses to external parameters such as tempera-
ture and strain rate.
Crazing was first recognized to be distinct from either shear processes or
cracking more than 30 years ago. The fact that crazes could grow right
across a specimen and yet the specimen retain mechanical integrity showed
that a craze must be load bearing. One of the first experimental methods
used to explore the internal structure of a craze was optical interferometry,
and over the years this has continued to yield information regarding the
crazing process (for useful reviews of this approach see Doll, 1983; Doll and

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
296 Crazing

Konczol, 1990; Schirrer, 1990). However interferometry, being an optical


measurement, does not have the spatial resolution to provide details of the
internal fine structure of a craze; this required the development of trans-
mission electron microscopy (TEM) methods to yield a clear picture of the
internal structure.
However, even at the level of optical microscopy much can be learned
about crazes. They grow normal to the principal (tensile) stress. They may
grow to up to centimetres in length and fractions of a millimeter in thickness
if conditions are such as to prevent early failure and crack propagation,
although they are frequently much shorter. They have a refractive index
different from the surrounding undeformed material, which means that they
scatter light: a stressed material that contains a high density of crazes is
sometimes described as 'stress-whitened' because the scattering causes a
normally clear material to become opaque.
Much of the early work to characterize crazes concentrated on the stress
conditions under which they grow. There is an inherent problem in such
studies, because so often crazes are initiated at flaws either within the
specimen or at the surface, where of course the stress conditions are not
accurately known. In addition, if optical methods for detecting the existence
of crazes are used, substantial craze growth beyond the initiation stage is
actually being monitored. Thus in general, the stress for craze propagation
is measured. Sternstein, Ongchin and Silverman (1968) studied the stress
conditions for craze formation in PMMA. They were able to show that
crazes grow perpendicular to the major principal stress, and that this stress
must exceed a critical value, although the precise expression they derived is
not now generally accepted.
In understanding crazing it is clear that there are several stages that have
to be appreciated: initiation - as touched on above; growth, both laterally
and in length; and ultimately breakdown, which will control the overall
strength and toughness of a material. In addition one must understand
under what conditions a polymer will deform by crazing rather than the
alternative mechanism of shear deformation. It is well known that many of
the tougher glassy polymers - such as polycarbonate (PC) and poly(styrene
acrylonitrile) (SAN), which forms the basis for the rubber-toughened poly-
mer ABS - are more prone to show shear than crazing. The stress
conditions promoting crazing were understood comparatively early, because
this could be deduced from optical microscopy of bulk specimens, but
understanding of the nature of the competition between the two processes
at a molecular level has taken much longer to achieve.
The aim of this chapter will be to concentrate on the more recent
developments at the microscopic rather than macroscopic level. In so doing
much of the emphasis will be placed on the results of experiments carried
Craze morphology 297

out on the atypical situation of thin films. It must be appreciated that the
stress conditions in these films will differ from those in a bulk thermoplastic.
In general, thin films will be in a state of plane stress, whilst thick samples
have a triaxial stress state. This difference will undoubtedly be very import-
ant when quantifying the conditions under which craze growth may occur,
but is of much less consequence when attempting to understand the
microscopic fundamentals. However, the reader should be aware of the
distinction between the two situations when reading through this chapter.
For a useful overview of the more macroscopic aspects leading to craze
breakdown and fracture, the reader is referred to the book by Kinloch and
Young (1983).

6.2 CRAZE MORPHOLOGY


The first direct visualization of the internal structure of a craze came from
the work of Kambour (Kambour, 1964, 1969; Kambour and Russell, 1971),
although the idea that there were fibrils spanning the craze interfaces was
proposed rather earlier (Bessonov and Kuvshinskii, 1960). Based on these
papers, plus many more recent studies, it has been shown that the craze
consists of a dense array of fibrils (presumed cylindrical for simplicity) which
span the craze, separated by voids. The existence of voiding instantly
separates craze deformation from shear deformation; the latter occurs
essentially at constant volume, the former has a substantial dilational
component. The interface between the crazed, fibrillar material and the bulk,
undeformed polymer is sharp. For crazes grown under many circumstances,
a narrow region down the centre which looks, in a TEM micrograph, lighter
than the surroundings can be identified. This is known as the midrib.
The fibrils in a craze on average lie perpendicular to the craze-bulk
interface and, since the craze grows normal to the major principal tensile
stress, this means the fibrils are aligned with the stress. Figure 6.1 shows an
electron micrograph of a craze in polystyrene, including a midrib, and a
schematic view of the microstructure (without the detail of a midrib).
Because the fibrillar structure is dense, the craze fibrils in a micrograph such
as Figure 6.1 overlap and it is difficult to quantify the dimensions associated
with the microstructure. For this it is convenient to turn to small angle
scattering.
Historically, small angle X-ray scattering was used first (Legrand, Kam-
bour and Haaf, 1972; Steger and Nielsen, 1978; Paredes and Fischer, 1979;
Brown and Kramer, 1981), but more recently, low angle electron diffraction
(LAED) has also been used (Brown 1983; Yang and Kramer, 1985, 1986).
The latter approach has two key advantages: it can be recorded from the
same area as an image is taken, and with the use of a selected area aperture,
298 Crazing

O.2J,lm

nbrll growth direction

Figure 6.1 (a) Transmission electron micrograph of a craze in polystyrene.


(b) Schematic representation of a simple picture of the internal structure of a craze.

diffraction only from within a single craze can be obtained, removing the
complicating effect of the craze-bulk interface. The findings from the two
methods are in good agreement and for crazes in polystyrene (PS) grown in
air, which have been most extensively studied, the fibril dimensions are
found to be ~ 6 nm (D in Figure 6.1 b) for the fibril diameter and ~ 20 nm
for the fibril spacing (Do) at room temperature. (The effect of changing the
crazing conditions, e.g. raising the temperature or crazing in the presence of
solvent will be discussed later.)
Craze morphology 299

There is a further parameter required to describe the craze microstructure


quantitatively, and that is the craze volume fraction (vr ), or its inverse the
craze extension ratio (denoted 2). Vr is always less than unity. The craze
volume fraction can conveniently be obtained from electron micrographs
(Brown, 1979; Lauterwasser and Kramer, 1979). A craze with a high value
of Vr has a dense array of fibrils. Since the true stress at in the fibrils will be
related to the externally applied stress a 00 by

a
t
= -aVr =
00 1
/l.a
00
(6.1)

if Vr is large, then the stress concentration acting on the fibrils is less high
than for a polymer with a lower value of Vr for a given value of a 00. Thus it
is clear that the value of Vr is of great significance in determining the
propensity for craze breakdown. It is also crucial for understanding the
mechanism of craze growth, and hence why different polymers respond the
way they do under different conditions of testing. The role of Vr (or
equivalently 2) is also important in understanding craze micromechanics, to
be discussed in section 6.4. If a craze has a thickness T(x} as a function of
position x along its length, then the displacement profile w(x} is given by

w(x} = iT(x}[1 - vr(x}] (6.2)

Since both T and Vr can be measured experimentally directly from TEM


micrographs, the form of w(x) can be determined.
In Figure 6.1b, where the internal structure of the craze is crudely
represented, it is implied that the craze microstructure consists only of fibrils
running perpendicular to the craze-bulk interface. This is now known to be
an oversimplification. Hints of the existence of cross-tie fibrils running
between the main fibrils were seen early on in TEM micrographs (Beahan,
Bevis and Hull, 1975). More recently, more quantitative evidence has come
from low angle electron diffraction (Yang and Kramer, 1985; Berger, 1989).
In a LA ED pattern, if only the main array of fibrils were present the
diffraction pattern would consist of a thin single streak perpendicular to the
fibril direction (recall that the correct orientation to the craze is readily
determined in the electron microscope). In practice for PS crazes grown in
air at room temperature, the streak is split into two which are aligned at
± 50 or so to the fibril direction. In addition there are two weak diffraction
spots along the meridian (i.e. parallel to the fibril direction) indicating the
existence of a weakly periodic structure in this direction. This information
leads to a modification of the original structure shown in Figure 6.1b to a
structure of the type shown in Figure 6.2, still at a schematic level. The fibrils
are now no longer precisely aligned with the external stress, but misaligned
300 Crazing

crosstie. - --iji;"'"-
libfil -

Figure 6.2 Refined schematic picture of the craze microstructure, allowing for the
existence of cross-tie fibrils.

by "" ± 5° to this direction (giving rise to the splitting), and there is some
approximate regularity in the spacing of the cross-tie fibrils, which causes
the appearance of a weak reflection on the meridian of the LAED pattern.
A discussion of the origin of this structure will be delayed until the
mechanisms for crazing at a molecular level have been presented.
The description above applies along the bulk of the length of the craze.
The craze slowly decreases in width to a tip with a radius of curvature
comparable with the fibril diameter. Figure 6.3 shows the form of a craze
tip in PS viewed at normal incidence. The fibrillar structure persists right up

Figure 6.3 TEM image of a craze tip in PS viewed at normal incidence.


Initiation and growth 301

to the craze tip. The overall aspect ratio of the craze (is it short and fat or
long and thin?) can be related to the way the craze grows, and will be
discussed further within the context of craze micromechanic:s in section 6.4.
The processes that go on at the tip determine the mechanism of craze tip
advance, and will be discussed in the section on craze growth.

6.3 INITIATION AND GROWTH

6.3.1 INITIATION
Initiation is probably the least well understood part of crazing. As alluded
to above, this is at least in part due to the difficulty in characterizing the
stress state sufficiently locally at the initiation site, which is usually some
sort of inhomogeneity. In addition, since craze initiation cannot be followed
in real time in the TEM (the electron beam would be likely to cause
crosslinking, and hence change the polymer's properties), it is not possible
to see directly the very early stages of craze formation. Recent experiments
using positron annihilation lifetime spectroscopy to examine nanovoid
content in glassy polymers under stress (Xie et ai., 1995) may point to a new
technique to examine this problem.
As indicated above, the stress conditions governing craze initiation may
be hard to obtain with precision locally, but have been well studied at the
macroscopic level. One of the first observations was that there tended to be
a delay in time between the application of a stress and the first appearance
of crazing (Argon and Hannoosh, 1977), suggesting the existence of some
barrier to craze nucleation. Argon and Hannoosh also established the
importance of imperfections in controlling the sites of craze initiation. When
especial care was taken to eliminate dust and surface flaws from PS samples,
it was found that the stress for craze initiation was raised substantially, and
in some cases shear yielding and necking occurred, most unusually for
tensile testing of PS.
This work also highlighted the importance of the detailed state of stress,
in agreement with earlier workers. Bowden and Oxborough (1973) for-
mulated a criterion for craze initiation in terms of a critical tensile strain
which depended on the hydrostatic component of the stress tensor. It can
be written as

X
0"1 - V0"2 - V0"3 = Y +----- (6.3)
0"1 + 0"2 + 0"3

where X and Yare time-temperature dependent parameters. In this


302 Crazing

expression the three o/s represent the three principal stresses (in decreasing
order of magnitude), and v is the Poisson ratio. That the hydrostatic
component of the stress tensor is an important parameter is not surprising
when it is remembered that the void-fibril structure of the craze means that
there is substantial dilatation associated with crazing. Whether crazes can
form at all if the hydrostatic component is zero or even negative is very hard
to ascertain, due to the uncertainties in local stress fields around the flaws
where crazes are likely to nucleate.
Turning now to events which need to occur at a much smaller length
scale, these have been summarized by Kramer (1983). He viewed craze
nucleation as involving three types of processes: (1) local plastic deformation
by shear in the vicinity of a defect; this leads to the build-up of significant
lateral stresses; (2) nucleation of voids to release the triaxial constraints; and
(3) void growth and strain hardening of the intervening polymer ligaments
as molecular orientation proceeds. In this way the incipient craze structure
is stabilized and the craze can thereafter propagate if appropriate stress
conditions hold. If the stress increase occurring in stage 1 promotes further
shear deformation, a craze nucleus will clearly not form. As yet, at the
initiation stage there is no simple molecular picture to determine whether or
not voiding will start to occur.
Given this three-stage process, the question of which stage is the critical
one needs also to be considered. It has generally been considered to be the
void nucleation stage. Argon and Hannoosh (1977), making this assump-
tion, tested their ideas of a critical porosity being required on PS, using a
theoretical model to relate the applied stress to the achieved porosity. If
nucleation is the critical stage, then it seems likely that there is a critical size
associated with this critical nucleus. There is some evidence to support this,
albeit inconclusive, based on work done on high impact polystyrene (HIPS;
Donald and Kramer, 1982a). It has long been known that small rubber
particles are rather inefficient at toughening, and this effect has been related
to the critical nucleus size. It was suggested that in order for a craze to
initiate, it will be necessary for the stress level to remain sufficiently high
over at least this critical distance. The extent of the stress concentration at
a rubber particle scales with particle size, and so it follows that small
particles are less likely to be able to satisfy this criterion, and therefore be
poor at nucleating crazes - as observed. However, recent developments
indicating the importance of cavitation in HIPS as well as ABS (Bubeck et
al., 1991; Lazzeri and Bucknall, 1993; Bucknall, Karpodinis and Zhang,
1994; Magalhaes and Borggreve, 1995) could explain this result by a
different mechanism, namely that the particle cavitation stress is higher for
smaller particles.
Initiation and growth 303

6.3.2 CRAZE GROWTH

(a) Craze Tip Advance


Once a craze is nucleated it must grow both in width and length. Originally
there were two different mechanisms proposed for craze tip advance -
advance by the nucleation and expansion of isolated voids (Argon, 1973)
and growth by an interface convolution mechanism known as the meniscus
instability (Fields and Ashby, 1976; Argon and Salama, 1977) - but the
evidence is now quite firmly down on the side of the latter. One of the
problems with the first mechanism - which was prompted by TEM evidence
on very thin films which appeared to show the existence of isolated voids
ahead of the crack tip (Kambour, 1973) - was that the geometry of the
mature craze with its network of interconnected voids was hard to reconcile
with the proposed initial state of well separated isolated voids, given that
the fibrillar structure was adopted immediately behind the craze tip with no
intermediate structure detectable. The proposed model was also found to
require impracticably high stresses in order to achieve the experimentally
observed craze growth rates. These difficulties are not present in the
meniscus instability. This instability is well known in other classes of
materials, and is based on the ideas of Taylor (1950) who derived it from
experiments on the interpenetration of two fluids of differing densities. In
this context, the voided structure of the craze corresponds to the low density
fluid, propagating into the denser, undeformed polymer.
The physical basis underlying this instability lies in the difference in
hydrostatic pressure across a curved surface: any perturbation which intro-
duces curvature can propagate if the resulting increase in negative pressure,
bah' exceeds that needed to produce this curvature. This condition can be
written as

r
bah >-
R (6.4)

where r is the surface energy and R is the radius of curvature. In principle,


any perturbation satisfying equation 6.4 can grow, but in practice there is a
dominant wavelength which develops, being the one that grows fastest. A
full analysis can be found in the original papers (Fields and Ashby, 1976;
Argon and Salama, 1977), with its application to the specific case of a craze
in Donald and Kramer (1981b), where the first direct TEM evidence
confirming the mechanism was presented. This evidence came from taking
stereopairs of the craze tip viewed at an angle. These confirmed that the
304 Crazing

a) b)

c) d)

Figure 6.4 Schematic view of the interface convolution mechanism (meniscus


instability) at a craze tip, which gives rise to the void - fibril structure of a craze:
(a) craze tip viewed from above, as in conventional TEM images; (b) - (d) successive
stages in the propagation of a craze tip, leaving behind isolated fibrils.

craze tip was not planar but broken up into a series of fingers running
through the thickness of the film. Figure 6.4 shows a schematic view of how
the instability can lead to the generation of the void- fibril structure, and
Figure 6.5 shows a TEM image of a tilted craze tip, in which the fingers are
clearly visible. It can be shown (Kramer, 1983) that this mechanism can
correctly predict the order of magnitude of both craze tip advance and craze
propagation stress. The parameter r which appears in equation 6.4 will be
seen in the following to playa crucial role. The ability to relate this quantity
to molecular processes has been one of the notable advances in our
understanding of crazing.
Initiation and 9rowth 305

Figure 6.5 TEM image of a craze tip viewed at an angle, showing the spreading out
of the tip into a series of fingers. These superimpose when viewed at normal
incidence, and so are not generally seen.

(b) Craze Widening


As the craze tip advances, the craze also grows in width. Again, originally
two different mechanisms were put forward, but the evidence for the correct
one (at least for crazes grown in air at room temperature) is now conclusive.
The two alternatives involved either existing fibrils elongating by fibril
creep, or new material constantly being drawn into the craze fibrils, with the
existing material remaining at a constant draw ratio as widening continued.
TEM methods permit a determination of the local value of ;. from analysis
of the optical density of micrographs. This method shows conclusively that
the draw ratio Ais constant along the vast majority of the craze length, only
increasing immediately behind the craze tip (Lauterwasser and Kramer,
1979; Donald, Kramer and Bubeck, 1982). This can only be consistent with
the surface drawing mechanism. Figure 6.6 shows an example of this
behaviour for polystyrene. As the craze tip propagates, this highly drawn
region remains at the centre of the widening craze and can be identified with
the midrib. The constancy of A means that for a given polymer, drawn under
a given temperature and strain rate e, there is some characteristic A that can
be identified. This finding, discussed further in section 6.S.2(a), led to the
306 Crazing

10 I I I


8

'. .. .... :'r


••
6
...:

• ••• •• •••

4 • -
• •
• ••
\
••
:1
craze tip
0
0 SO 100 ISO 200
position along craze (lim)
Figure 6.6 Plot of the extension ratio A. as a function of position along an isolated
craze in PS. A. is essentially constant everywhere except immediately behind the craze
tip.

realization of the importance of intrinsic molecular parameters, character-


ized by the entanglement network, in controlling craze microstructure and
growth.
Given that the craze widens by a surface drawing mechanism, one must
now consider how this mechanism operates. If one considers the form of the
craze-bulk interface shown in Figure 6.2, it is clear that it is reminiscent of
the convoluted structure of the craze tip represented in Figure 6.4. This is
shown more clearly in Figure 6.7, where the region at the base of the fibrils
is shown in more detail. As in the case of the craze tip, one can visualize the
process of void propagation into the undeformed polymer - corresponding
to craze thickening - as occurring via the meniscus instability. The same
analysis as for the craze tip will therefore apply, with the radius of curvature
R in equation 6.4 now corresponding to half the interfibrillar spacing Do,
and the surface energy r again playing a crucial role. Details of this analysis
are given by Kramer (1983). For the geometry of Figure 6.7, the pressure
gradient is now

v (ITO)m - (ITo).
(6.5)
ITo'" Do/2
and (ITo)., the hydrostatic tension at the tip of the void, is given by
4r
(ITo). '" D (6.6)
o
Craze micromechanics 307

ftbril

Figure 6.7 Representation of the structure at the craze- bulk interface, focusing in
on the base of the fibrils . This structure is reminiscent of the craze tip structure
shown in Figure 6.3 .

(O'O)m, the hydrostatic tension in the layer above the fibril, is assumed (Fields
and Ashby, 1976) to be proportional to the average tensile stress S at the
craze interface.
As before, there will be a dominant wavelength which grows fastest, which
will determine the experimentally determined value of Do. and hence the
geometry of the craze. It will be seen that the craze propagation stress is an
important parameter, and is likely to be dependent on the particular
polymer being used. Determination of this stress takes us into the realms of
craze micromechanics.

6.4 CRAZE MICROMECHANICS


One of the simplest and most commonly used approaches to model craze
micromechanics has been borrowed from the metallurgical literature. This
model, known as the Dugdale model, can be used to describe a single craze
growing at a crack tip (Dugdale, 1960; Goodier and Field, 1963). It was
originally developed to describe plane-stress plastic zones running ahead of
a crack tip in a metal, but actually more realistically describes the case of a
craze. In this model it is assumed that the craze surface stress is a constant,
S, along its entire length. Within this model, this stress then defines the
shape of the craze profile, and hence it also defines the length of the craze
that grows at a crack of a given size ao. A useful summary of this approach
to craze micromechanics is given by Kramer (1979a), where alternative
models for craze micromechanics are also discussed.
The precise form of the displacement profile w(x) is given by

as
w(x) = nE* R(P, Pc) (6.7a)
308 Crazing

where

(6.7b)

(6.7c)

(6.7d)

and
E* = E/(1 - v2 ) for plane strain and
=E for plane stress (6.7e)

Following through this analysis yields

(6.8)

where (J 00 is the applied stress.


This model has been applied frequently to the analysis of interferometric
experiments (Brown and Ward, 1973; Weidmann and Doll, 1976) and,
within the resolution of this optical method, very good agreement was found
between the experimentally determined profile and that predicted by the
model. However, minor discrepancies are found between experiment and
theory if the higher resolution of the TEM is used to characterize the craze
profile. In particular, it is found that the displacement profile in the vicinity
of the craze tip is not well fitted by the Dugdale approach (Wang and
Kramer, 1982).
This difference in the form of the displacement profile in the vicinity of
the tip also has consequences for the craze surface stress, assumed constant
along the entire length within the Dugdale model. There are various routes
to move from an actual displacement profile to the stress profile, reviewed
by Kramer (1983). Whichever method is chosen shows that there is a stress
concentration at the craze tip. How sharp this concentration is depends on
the sample, with monodisperse material, for instance, showing a much more
pronounced concentration than polydisperse (Donald and Kramer, 1983).
However, away from the tip region the surface stress is approximately
constant, and in general lies close to both the value predicted by the
Dugdale model (on the basis of the craze length) and the externally applied
tensile stress. Figure 6.8 shows the stress profile for mono disperse PS.
Craze micromechanics 309

40
35 •
30 •
""'
to
Il.
~
25
'-' 20
+ 0'00
••••••••••••
• .:j. · ·
til
craze tip
...e 15
til

til
10
5
0
0 50 100 150 200
position along craze (11m)

Figure 6.8 Stress profile along a craze in monodisperse PS (molecular weight


200000).

The rise in stress just behind the tip has been related to the increase in A.
there: it is to be expected that the extension ratio will to some extent depend
on the drawing stress, since the material in the fibrils should have a finite
strain hardening rate. A comparison of the surface stress profile with the
extension ratio profile (Lauterwasser and Kramer, 1979) shows exactly this
effect, with the local increase in stress coinciding with the local increase in
A.. The midrib in a mature craze corresponds to this highly drawn material,
which has since been surrounded by less drawn material as the craze widens.
One further manifestation of the limitations of the Dugdale model, with
its assumption of a constant stress along the length of the craze, can be seen
when the surface stress profiles in different polymers are compared (Donald,
Kramer and Bubeck, 1982). Whereas for PS, crazes tend to be long and thin,
with a surface stress close to constant along their entire length and also close
to (J 00' this is not the case for tougher polymers such as poly(phenylene
oxide) (PPO) and poly(styrene acrylonitrile) (PSAN) (which will be shown
below to be those which have high values for the entanglement density). For
these polymers, the crazes tend to have much higher aspect ratios, and for
crazes grown from crack tips their stress drops rapidly away from the crack
tip, with a stress concentration also at the craze tip. This effect can be related
to the much higher stresses required to maintain craze growth in these
polymers - as will be discussed below - with the consequence that the
crazes cannot grow very far before shear becomes a process competitive with
crazing. Thus crazes in PPO and PSAN are often seen to have shear regions
ahead of the craze tip, and this process limits the extent of craze growth.
310 Crazing

To summarize this section on craze micromechanics, one can say that the
simplest model, the Dugdale model, is generally an adequate representation
of the properties of a craze at a crack tip. Its assumption of a constant stress
along the entire craze length is not far from the truth, with deviations only
being significant within a few micrometres of the craze tip, at least for PS.
For isolated crazes, that is those not growing from a crack, there are other
approaches to characterize the profiles of the various parameters, but again
the assumption of a constant stress and A along the craze are good
approximations, except very locally at the craze tip.

6.S MOLECULAR MECHANISMS

6.5.1 INTRODUCTION
So far this chapter has dealt with crazing essentially from a phenomenologi-
cal point of view. This was the best that could be achieved at the time of the
first edition of this book, but in the last 15 years or so there have been
significant developments at a molecular level enabling the subject to move
on rapidly. Much of this progress has stemmed from the work of Kramer
and coworkers. The first step in this progression was an appreciation of the
significance of the finding that the craze extension ratio A was a constant for
a given material (given constant testing conditions) and that it could be
related to other known properties of the material derived from totally
different types of experiments, including those carried out on the melt. It has
long been known that when a polymer necks macroscopically (both glassy
and semicrystalline polymers do this; Allison and Ward, 1967), new material
is drawn in to keep the extension ratio in the neck a constant. The work of
Ward (Coates and Ward, 1978, 1980) is notable in this area. That the craze
fibrils should behave in the same way should perhaps have come as no
surprise, but this understanding meant that it became possible to under-
stand how the molecules themselves were reacting to the externally applied
stress, rather than simply describing the phenomenology of crazing.

6.5.2 THE ENTANGLEMENT NETWORK

(a) Why is A a Material Constant?


There were various early pieces of evidence to suggest that the entanglement
network was of relevance to an understanding of deformation. One key
Molecular mechanisms 311

finding was described by Haward in Chapter 6 of the first edition of this


book, referring to the ability of deformed polymers to revert to their original
undeformed state when heated briefly above their glass transition tempera-
ture ~ (Gurevich and Kobeko, 1940; Haward, 1942; Hoff, 1952; Murphy,
Haward and White, 1971). This observation suggests some memory effect
akin to the retraction of a stretched piece of rubber, which points to the
existence of some sort of network comparable to the rubber's network of
crosslink points (Haward and Thackray, 1968). In addition. it was observed
comparatively early on that above a certain critical molecular weight the
crazing stress for PS at room temperature was molecular weight indepen-
dent, and this critical molecular weight was comparable with the Me known
from melt rheology measurements (Fellers and Kee, 1974). Kramer (1979b)
explored the failure behaviour of PS in more detail as a function of
molecular weight, and established that stable crazes were unable to form for
low molecular weights. However, very narrow crazes were postulated whose
width was defined by the molecular weight - i.e. the craze width was taken
to be equal to the length of the stretched-out chain. By contrast, stable craze
fibrils would need the chains to be entangled.
The idea of an entanglement network has come, as indicated above, from
studies of polymer melt rheology (for comprehensive reviews see Ferry,
1980, for a phenomenological account or Doi and Edwards, 1986, for a more
theoretical approach). The entanglement molecular weight Me is usually
taken from the position of the plateau in the shear modulus G~ (again by
analogy with rubber elasticity), from the relationship
pRT
Me= G~ (6.9)

where p is the density, R the gas constant and T the temperature. Associated
with this value of Me - which will be polymer specific, but independent of
parameters such as overall molecular weight or temperature - are three
other key parameters. These are the entanglement density (the density of
strands in the entanglement network) Ve , the contour length between
entanglement points Ie' and the root mean square end-to-end distance d of
a chain of molecular weight Me. The appropriate relationships between
these three parameters and Me are
pNA
v = - (6.10)
e Me
(NA is the Avogadro constant);
I = lOMe (6.11)
e Mo
312 Crazing

where 10 is the average projected length of a monomer unit along the chain
whose molecular weight is M 0' and
d = k(M .)1/2 (6.12)
M. can readily be determined for the melt (but note that the so-called
critical molecular weight Me derived from melt viscosity measurements is
usually higher than M. by a factor of ",,2 (Ferry, 1980)}; k can be obtained
either directly from neutron scattering studies on the glass or, where these
results are not available, from light scattering studies on a dilute solution in
a () solvent. For a single chain it is easy to appreciate that there will be a
maximum extension ratio Amax given by

Amax = ~ (6.13)

If the entanglement network really does determine the crazing response of


the polymer, then one might expect a correlation between this quantity Amax
and the extension ratio of a craze. Figure 6.9 shows a comparison of these
two measures of extension ratio for a selection of polymers tested at room
temperature (Donald and Kramer, 1982b). These early experiments were
fairly crude, in that the strain rate ewas not controlled: as we will see later,
e(as well as temperature) is an important external variable, and it is perhaps
surprising that the correlation worked so well.
Since these initial experiments there have been a series of further tests of
this simple idea on samples where Ve could be controlled in a more
systematic way than merely by choosing a wide range of polymers. Histori-
cally, the next experiments were carried out on a series of blends of PS with
PPO (Donald and Kramer, 1982c). These two polymers are miscible over
their entire composition range, but they have very different values of M.:
19 100 for PS and 4300 for PPO. Melt rheological data for comparable
blends had previously been obtained by Prest and Porter, (1972), which
allowed an estimate of M. for all the blends to be made, and hence predicted
values for Amax as a function of PPO content which could be compared with
Acraz •• As was the case shown in Figure 6.9, a good correlation between these
two quantities was found. Most recently, the case of PS deliberately
crosslinked to change the network strand density was studied (Henkee and
Kramer, 1985). Again this allowed systematic variations in the predicted
Amax to be made, and again the correlation with A craz • was found to be good.
If an examination is made of which polymers (as in Figure 6.9) show a
high extension ratio (or equivalently low vel, it is found to be those that are
usually described as 'brittle', such as PS. Polymers regarded as 'tough',
including polycarbonate (PC) and PPO, have craze extension ratios which
are much lower. These are also the polymers that at room temperature only
Molecular mechanisms 313

8
7 o

6
S
~
...<i
II
4 o
c<
3
00
2 (JI!)O

1
0
0 1 2 3 4 S 6 7 8
A-max
Figure 6.9 A comparison of Amax and Acraz• for a variety of polymers. The straight
line indicates the line Acraz• = Amax. (Data from Donald and Kramer, 1982b).

craze with difficulty. These polymers will craze more readily in thick samples
than in thin. For the thin films, on which the TEM studies and the
determinations of Acraze have been made, crazing is often suppressed because
of the lack of triaxiality in the stress field (triaxiality being important, as
discussed in section 6.3). In this case for the thin films, a new form of
deformation is frequently seen. This manifests itself in the form of regions of
homogeneous deformation, as shown in Figure 6.10. These zones, known as
deformation zones (DZs), grow normal to the applied tensile stress in just
the same way as crazes do. However, they are distinguished from crazes in
their lack of voiding.
For polymers with high values of Ve (or equivalently low Me), these zones
tend to dominate over crazes in thin films. For the family of blends of
ps- PPO referred to above, as the proportion of PPO increases, crazes are
seen less and less often. Crazing can be promoted by annealing the films
below their ~ prior to deformation; annealing in this way is known to
increase the yield stress (Jy. The deformation zones are regions of shear
through the thickness of the film, equivalent to the inclined shear bands seen
in thick samples. Weare therefore seeing the competition between shear and
crazing, which will be considered in more detail in section 6.5.2(c).
The extension ratio of these deformation zones can be measured from
TEM micrographs in exactly the same way as for crazes. Although only a
limited range of samples exhibit DZ formation, a strong correlation between
their extension ratio and Amax is again seen, as shown in Figure 6.11. If the
extension ratios in a craze and a DZ in the same polymer are compared
314 Crazing

Figure 6.10 A deformation zone in Pc. (Reproduced from Donald and Kramer,
1981 a, with permission.)

(for instance by looking at a DZ in an un annealed sample and a craze in


an annealed one; annealing will affect neither Me nor Amax), it is found that
ADZ is always less than Acraze . This is not surprising given that crazes contain
voids, whereas deformation zones do not. However, the finding draws
attention to the fact that, whereas it is conceivable that DZs form without
destroying any entanglement points, the fact that crazes contain voids points
to the fact that the entanglement network must be being modified by the
crazing process. This 'geometrically required entanglement loss' was first
appreciated by Kramer (1983). He enumerated how many entanglement
points had to be lost to allow the craze microstructure to form. The number
of intact entanglement points left must depend on the precise structure of
the craze, and crucially on the fibril spacing Do. His calculation showed that
if v~ is the entanglement density in the fibril, it is given to a good

3.------r----r-----r-----,

2
o 0
co
o
1

o ~--~--~--~--~
o 1 2 3 4

"-max
Figure 6.11 Plot of ADZ versus Amax . The straight line indicates the line ADZ = O . 6~ax '
(Data from Donald and Kramer, 1982b.)
Molecular mechanisms 315

approximation by

(6.14)

(b) Mechanisms of Entanglement Loss


Given that entanglement points must be lost to permit the void-fibril
structure of the craze to form, one must ask how this loss occurs. Figure
6.12 shows in more detail the nature of entangled chains at the fibril base.
Two alternative processes can be envisaged in which the entanglement point
at C is lost: scission, in which the chain is actually broken, and disentangle-
ment, when the entanglement point is lost by the relative motion of the two
chains involved. Both processes are now known to take place under
appropriate conditions. Which process actually takes place, or whether the
stress required for both is too high so that yield, and hence shear, takes place
in preference, determines the response of the glassy polymer to an external
stress. It therefore also determines how many chains remain intact after
crazing, and hence how resistant the craze is to subsequent breakdown.
Thus we need to find a theoretical approach to understand which of the two
crazing mechanisms operates.
Let us start with the case of scission. Based on the evidence of Fellers and
Kee (1974) that the crazing stress in PS was molecular weight independent
(at room temperature), scission was identified as the operative mechanism.

craze fibri
Figure 6.12 A schematic representation of the surface drawing mechanism during
craze growth. The chains marked A and B are entangled at C. This entanglement
point must be broken for the chains to continue to move into the fibrils. (After
Plummer and Donald, 1990a.)
316 Crazing

Disentanglement would have led to a molecular weight dependence of the


crazing stress, since it would be anticipated (correctly as we will show) that
shorter chains would be more mobile and disentangle more readily than
long chains. Thus, historically, it was initially assumed that crazes formed
via scission.
We now need to return to reconsider the way in which a craze widens,
discussed in section 6.3.2(b), taking into account the occurrence of scission.
When new void surface is created during this widening, the surface energy
r associated with it, and which appears for instance in equation 6.6, is not
simply the van der Waals surface energy y, but contains an additional term
arising from the energy cost to break the required number of chains. The
number of entangled strands crossing a unit area is approximately ve d/2,
and this leads to an extra term of the form Udve/4, where U is the energy
to break a single backbone bond. Thus the total surface energy associated
with the crazing process becomes (Kramer, 1983)

Udve
r =y+-- (6.15)
4

This surface energy term now determines the stress required for the
widening process to continue. Equation 6.6 showed that the hydrostatic
stress just at the tip of the void depended on r. Furthermore, it can be
shown (Kramer, 1983) that the overall stress required for craze widening
(based on the meniscus instability) takes the form
(6.16)
where (Jy is the yield stress (assumed equal to the flow stress) and m is an
empirical power law constant relating the stress to the strain rate E. Hence
S is crucially dependent on r.
Thus if scission is occurring, the stress for craze growth depends through
r on the entanglement density. PS is not a highly entangled polymer (ve is
'" 3 X 1025 m - 3), and this means that the additional term in equation 6.15
due to scission is about equal to the van der Waals term, each contribution
representing about 40 mJ m - 2. However, for more highly entangled poly-
mers such as PC, which has value of Ve approximately a factor of ten higher,
the additional contribution to r is 0.2 J m - 2, and consequently a much
higher stress would be required to cause a craze to form by scission. This
finding provides insight into why PC so often deforms not by crazing, but
readily forms shear bands (or DZs in thin films).
However, PC certainly can craze, even in the absence of solvents -
it is notoriously susceptible to solvent crazing - but it does so more readily
at high temperatures, as was noted (Wellinghof and Baer, 1978; Donald and
Molecular mechanisms 317

Kramer, 1982d) long before understanding of crazing mechanisms came


about. If scission crazing is ruled out on the grounds that the stress required
lies above the shear yield stress, can we understand high temperature crazing
in highly entangled polymers by considering the alternative crazing mech-
anism of disentanglement? Since disentanglement implies chain mobility, it
is natural to expect it to be a more viable mechanism as the temperature is
raised towards the glass transition temperature.
To understand the effect of disentanglement on r, and hence the crazing
stress, we must look in more detail at what happens at the fibril base, and
consider what length of chain must be mobilized for the entanglement point
C in Figure 6.12 to be lost. In addition, some thought must be given to how
the chains can move at all in the glassy state, since naively the glassy state
is thought to be a state in which long range chain motion is frozen out. The
key to understanding this is to realize we are not, as originally thought,
considering the classical picture of melt rheology in whIch reptation is
clearly identified as the source of motion (Doi and Edwards, 1978).
Brownian motion, the source of mobility in reptation, will not be present.
However, the presence of an external stress in itself provides a mechanism
to mobilize the chains, with the consequent motion being dubbed 'forced
reptation' (McLeish, Plummer and Donald, 1989).
When a chain is pulled out of its constraining tube (using the terminology
of Edwards (1967», it experiences a frictional force fd given by
v(oM
fd~~ (6.17)
o
where (0 is the monomeric friction coefficient and v is the rate of chain
pull-out, assumed to be equal to the craze interface advance rate (which in
turn will be proportional to the macroscopic strain rate e). This frictional
force can be compared with the force fb necessary to break a chain

J;~-
u (6.18)
b 2a
where a is the bond length (Kramer, 1983). If fd < fb' then disentanglement
will take place in preference to scission.
However, in practice it is not necessary to mobilize an entire chain, but
only that section of the chain lying entirely on one or other side of the
dividing plane (Figure 6.13), whichever is the shorter. If this shorter section
has a molecular weight of xM (where x :( 0.5), then the relevant frictional
force becomes

(6.19)
318 Crazing

plan or
epa ration

craze fibri
Figure 6.13 Schematic representation of a chain traversing the plane of separation
between two fibrils. As the void tip advances, the chain, assumed to be constrained
to move along the virtual tube, either disengages by forced reptation from the
shorter of the two sections, or undergoes scission .

In this equation, x can take any value between 0 and 0.5, and will do so for
the various different chains near the interface. This means that some of these
chains may satisfy the criterion for disentanglement, while others do not.
The critical value of x, xc' above which chains will break and below which
disentanglement will occur is given by equating equation 6.19 with the force
for scission fb (equation 6.18). This yields an equation for xc:

(6.20)

With both SCISSIOn and disentanglement operative, the surface energy


associated with fibrillation, and therefore the stress for craze propagation,
will be given by an appropriately weighted sum of the energy terms
associated with the two mechanisms.
In order to compute this resulting r it is necessary to make some
assumption about the distribution of x for the chains in the vicinity of the
interface. In the absence of any evidence to the contrary, this has been
assumed (Kramer and Berger, 1990) to be fiat, so that all values of x from
o to 0.5 are equally likely. With this assumption one obtains

dVeU
r = y + ( 1 - xJ -4- for Xc ~ 0.5 (6.21a)
Molecular mechanisms 319

and

1 dveU
r=y+--- for Xc ~ 0.5 (6.21b)
4xc 4

In practice, a chain may cross the dividing plane between fibrils more than
once on average (clearly depending on molecular weight). This effect can be
allowed for (Plummer and Donald, 1990a) and leads to numerical modifi-
cations to equations 6.21. However, the basic principles remain unchanged,
although the modifications do lead to a better fit of the theory with actual
experiment.
It will be noted above that the term for Xc' and hence through r the craze
propagation stress, depends on the monomeric friction coefficient (0' which
will be a function of temperature. It also depends on the craze widening
velocity - which will be determined by the externally applied strain rate. We
thus see how this model allows us to take into account external parameters
which may shift the balance not only between scission and disentanglement
crazing, but also between crazing and shear. This will be discussed in more
detail in the next section. The experimentally observable change in response
to stress as a function of temperature has been one of the key tests that have
verified the correctness of the above approach to understanding the mech-
anistics of crazing.

(c) Competition Between Shear and Crazing


Already, hints have been made about what may control the balance between
shear and crazing. We have seen that it is the highly entangled polymers
which require high stresses for crazing via scission; these polymers, including
PC and PPO, also tend to have high glass transition temperatures. As a
consequence their monomeric friction coefficient is likely to be high at room
temperature, which will lie a long way below 1'g. This means that disen-
tanglement will also be hard to achieve, and so the stress for crazing via
either mechanism will be high. It is for this reason that we associate shear
with the highly entangled polymers.
Armed with this reasoning, it also becomes clear why the appearance of
shear deformation zones tends to be suppressed in favour of crazing if the
polymer is annealed below its 1'g. The annealing has no effect on the various
crazing parameters, but it is well known for polymers such as PC that such
a heat treatment increases the yield stress (Legrand, 1969; Pitman, Ward
and Duckett, 1978). Hence shear no longer occurs at the lower stress and
crazing starts to appear.
320 Crazing

However, annealing is not the only way in which the balance can be
shifted between shear and crazing. Other factors can be seen to come into
play when polymers of intermediate entanglement density are studied: these
are the ones where, at room temperature, the balance between crazing
and shear is most delicate. For polymers with very high values of ve (such
as PC), shear is almost completely dominant; for polymers with low values
(e.g. PS), crazing is essentially always seen. As can be seen from equation
6.20, the critical value of Xc depends inversely on v, the craze widening rate.
This parameter directly correlates with the externally applied strain rate e.
Hence, as the strain rate goes up, the proportion of chains that undergo
disentanglement goes down. However, if one imagines suddently applying a
tensile stress, one can envisage a three-stage process occurring (Donald and
Kramer, 1982d; Donald, Kramer and Bubeck, 1982). First, at the highest
stress levels, chain scission leads to the appearance of crazing. Then, as the

Figure 6.14 Composite structure showing the sequence of events (scission crazing
followed by shear) as one moves out from the centre of a region of deformation in
high molecular weight PS strained at 70°e. (Reprinted from Donald, 1985, with
permission .)
Effect of external parameters 321

average stress level drops below the level where rapid chain breakage may
occur, shear processes start to dominate. Finally, at long times (low l;),
disentanglement becomes possible and fibrillation may reappear. Such a
complicated sequence of events means that the deformation that occurs may
itself be complex and not simply either crazing or shear. An example of such
a composite structure is shown in Figure 6.14.

6.6 EFFECT OF EXTERNAL PARAMETERS

6.6.1 TEMPERATURE AND STRAIN RATE


So far this chapter has concentrated on the effect of deformation at room
temperature. For all the polymers considered, this lies well below their
glass temperature. However, it has been hinted at that as the temperature is
raised, in particular, we may expect to see a decrease in (0' and hence a rise
in the rate of disentanglement. We can use the arguments presented in the
last section to rationalize why it is that PS becomes more ductile as the
temperature is raised - exhibiting more shear and less crazing (Kramer and
Berger, 1990) - in contrast to PC, which is more prone to show crazing at
high temperatures. These conflicting behaviours initially caused puzzlement,
but can now be understood based on an appreciation of the factors that
affect r, and how this in turn determines the crazing stress. In many ways
the effects of temperature and strain rate are inversely related, particularly
within the context of disentanglement, so the two parameters will be
considered together in this section.
As in the previous section, much of our understanding has been derived
from work on thin films, but the arguments can be shown to carryover to
bulk samples. First, let us consider what happens in crosslinked polymers,
for which disentanglement is not an option, so that only the competition
between shear and scission crazing needs to be considered. For crosslinked
PS with various strand densities, it has been shown that Acraze is independent
of temperature (Berger and Kramer, 1988). This is to be expected as the
network of crosslinked chains is unaffected by the temperature. More
interesting, however, is the balance between shear and scission crazing as
the temperature and strand density are altered. Figure 6.15 shows a plot of
how the dominant mode of deformation changes with these parameters for
a strain rate of 10 - 6 S - 1. To understand this figure, at least at a semiquan-
titative level, it is instructive to consider how the stresses for shear and yield
depends on the two changing external parameters.
The shear yield stress (J y should not depend on Ve ' since shear involves
only local segmental motion. This is not of course the case for the crazing
stress, which will be strongly affected due to the dependence of r on Ve
322 Crazing

100
+
80
...
CII
:;,
0 III + +
....nI 60 0 III +
...CII
Q.
crazing
E 40 crazing and shear
CII
I- shear
20 00 III III III +

0
0 S 10 lS 20
Strand density v x 10Z5 m· 3
Figure 6.15 Summary of the dominant mode of plastic deformation observed in
crosslinked PS as a function of strand density and temperature. Open squares:
crazing only; crossed squares: crazing plus shear; crosses: shear. (Data from Berger
and Kramer, 1988.)

(equation 6.15) and hence S (equation 6.16). Figure 6.16 shows schemati-
cally how Sand (Jy depend on temperature. In deriving this plot, it is
assumed that (Jy is linearly dependent on temperature. Since, through
equation 6.16, the crazing stress has a square root dependence on T, the two
curves for Sand (Jy cross, and it is for this reason that there is a shift from
shear at low temperatures to crazing at high. The temperature at which this
shear-to-craze transition occurs will depend on the externally applied strain
rate, since it is well known that (Jy is strain rate dependent, shifting both
curves. Clearly also, the transition will be dependent on the crosslink
density, since the energy penalty associated with scission directly depends
on this parameter.
For uncross linked polymers, which historically were the first to be studied
(Donald, 1985; Berger et al., 1987; Plummer and Donald, 1989), the
situation is more complicated since the balance between the two types of
crazing, as well as with shear, constantly changes as the temperature is
raised. The first concrete evidence to confirm the occurrence of disentangle-
ment came from experiments on PS of different molecular weights (Donald,
1985). For low molecular weights, crazing persisted right up to the highest
temperatures studied close to ~. However, for high molecular weights
(MW) (of the order of 10 6 ) this was not the case with shear intervening
above about 70°C. This can be understood qualitatively by realizing that for
all molecular weights, the shear yield stress has the same (presumed linear)
temperature dependence. In contrast, the crazing stress will drop more
slowly for the low MW PS, since the balance can shift from scission to
Effect of external parameters 323

UI
UI

......
CII

V)

/'

I
//
o 20 40 60 80 100
T -T (DC)
g

Figure 6.16 Schematic representation of the temperature dependence of the shear


yield stress (Jy (taken as linear) and the crazing stress 5. The crossing of the two lines
indicates why there is a shear-craze transition. (After Berger and Kramer, 1988.)

disentanglement crazing with its lesser energy penalty and contribution to


r, than for high MW PS for which the chains are too long for a significant
contribution from disentanglement to occur right up to Tg • Hence the curves
for yield stress and crazing can cross for high MW material but not for low,
and crazing persists over the whole temperature range for the low MW PS.
This can be put on a more quantitative footing by obtaining values of So
from macroscopic mechanical tests explicitly rather than relying on indirect
microstructural evidence. Recalling the surface drawing mechanism for craze
widening described in section 6.3.2(b), equation 6.6 showed the local stress
could be related to the ratio of r and the fibril spacing Do. This analysis can
be taken further to show that the craze widening stress S" is related to the
fastest growing wavelength D6, which will dominate in the microstructure,
by the relation

(6.22)

where f3 is a constant of the order of unity. Thus if we can determine


D6 from micrographs and Sc from mechanical tests, then the behaviour
of r can be explored. When this is done, the fall-off in the value of r
with temperature from the room temperature value corresponding to the
scission-only case is clear (Berger et al., 1987). Thus quantitatively the
picture of a scission-disentanglement transition is confirmed.
As described in section 6.5.2c, highly entangled polymers tend to exhibit
324 Crazing

shear deformation at room temperature because the stress associated with


scission crazing is too high. However, as the temperature is raised and
disentanglement becomes a possibility, crazing can occur more readily. A
study of the deformation mode as a function of temperature for the two
polymers, PC and polyethersulphone (PES), showed clearly how this
transition to crazing was affected by MW (an increase in which will defer
the onset of disentanglement and so permit shear to persist to higher
temperatures) and strain rate. As e increases again there will be a delay in
the temperature at which disentanglement becomes possible. Figures 6.17
and 6.18 show the effect of these two parameters on the strain for deforma-
tion onset (which will of course be correlated with the stress via a
temperature-dependent Young's modulus). At the highest temperatures, the
strain becomes independent of molecular weight, since here the only term
contributing to r is the van der Waals energy y. Disentanglement can occur
so rapidly that there is no scission contribution. However, at lower tempera-
tures the expected delay in the shear-to-craze transition for the higher
molecular weight and higher e samples is clear. It is also apparent that as
crazing starts to occur there is a sharp drop in stress for all samples.
Later on we will consider craze stability. Crucial factors in determining
this are the craze fibril extension ratio A, and the proportion of chains that
have undergone scission. Plots of A for a range of polymers show in all cases
a steep rise as the temperature approaches ~ (Berger and Kramer, 1987;
Plummer and Donald, 1989). This rise in A (shown in Figure 6.19 for PS of
different molecular weights) can be shown not to coincide with the onset of

rxshear
45 1-.. x~.x
• t'" x x

••

.~ 3.5
.......
fI.l 3
• crazing
• x x

• • x
2.5 • •
• x
x • x
2 ~~~~~~~~~~~~x~~~

300 350 400 450 500


Temperature (K)
Figure 6.17 Strain for deformation onset as a function of temperature for two
molecular weights of poly(ether sulphone); e: MW ~ 47000; x: MW ~ 69000.
(Redrawn from Plummer and Donald, 1989.)
Effect of external parameters 325

5
shear
4.5 III • x ••

.-... 4 xx
~

.;= 3.5 ~ • 'lccrazing


J:rIl 3

••
2.5 •• ••
•••••• xXxxx
2 ~~~~~.~~~~••~&U.~~
••~~

300 350 400 450 500


Temperature (K)

Figure 6.18 Strain for deformation onset as a function of temperature at


two strain rates for PES with a Mw of ",47000; x, B= 1 x 10- 2 5-'; e, B =
6 X 10- 5 5-', (Redrawn from Plummer and Donald, 1989,)

disentanglement, but actually occurs some lOoe or so afterwards. At the


onset of disentanglement, the value of A remains at the same value of about
4 as at room temperature. That this is so is due to the fact, discussed in
section 6.5.2(a), that there is a 'geometrically necessary entanglement loss'
to allow the craze structure to develop. This entanglement loss is the same
whether it occurs by disentanglement or, for PS, the room temperature

7 •
I
I
6 I
I
5
,I

4 ~,=-,~,~"::--~ ::"~-+: ~ ~ , , ' ' ' K' • '


..
- .X

3
20 30 40 50 60 70 80 90 100
Temperature
Figure 6.19 Plot of A versus temperature for PS of different molecular weights; e,
MW = 100000; +, MW = 390000; x, MW = 3 x 10 6 , (Data from Berger and
Kramer, 1987,)
326 Crazing

mechanism of scission. Thus the entanglements that are lost as disentangle-


ment starts to occur are only those which at lower temperatures are lost by
scission. It is only at higher temperatures, when disentanglement of chains
that are loaded at less than the maximum can also occur, that A starts to
rise. Since the ease of disentanglement is greatest for the shortest chains, this
rise in A occurs at the lowest temperature for these MWs, as can be seen in
Figure 6.19. The rise will also be determined by the externally applied strain
rate and, as in all these measurements, the precise temperature at which the
onset occurs will be e dependent.
Although it is intuitively obvious that disentanglement will become easier
as the temperature is raised (leading to a decreasing value of Co in equation
6.19), knowing quantitatively how this parameter changes is harder to
establish. Co is derived from melt rheology, and it is not clear how to
evaluate its temperature dependence in the glass. In the absence of any other
information, the best estimate is to assume the same temperature depend-
ence in the glassy state as in the melt (Berger and Kramer, 1988; Plummer
and Donald, 1990a). If this assumption is made, it becomes possible to
predict how r will depend on temperature, and a comparison can be made
with experimental data (obtained from knowledge of D6 and Sc as described
by equation 6.22).
This has been done for both monodisperse PS and polydisperse PES
(Plummer and Donald, 1990a). These results, shown in Figure 6.20 for PS,
show that reasonable agreement exists, particularly if the possibility of
multiple crossing of a chain across the dividing plane between fibrils
(alluded to in section 6.5.2(b» is taken into account. This semiquantitative
agreement is gratifying, but it must be recognized that the picture presented
here is probably still too simple. In particular, at high temperatures at least
two effects may come into play that have not been considered in this
description. First, the mechanism of widening may change as the tempera-
ture approaches ~. This could manifest itself either in some contribution
due to fibril creep, or a change in the strain softened material at the fibril
bases, leading to an increase in the thickness of the so-called 'active zone'
(Kramer and Berger, 1990) there. Second, as the entanglement loss and A
both increase at high temperature, the fibrils are likely to become increas-
ingly fragile due to the effect on the fibril true stress. This effect may cause
additional breakdowns affecting either the measured A or the interfibrillar
spacing D6. Nevertheless it seems clear that the basic picture is adequately
represented. A similar analysis for PES (Plummer and Donald, 1990a) does
suggest agreement between theory and experiment is less good for a
polydisperse material than for the monodisperse PS. This discrepancy could
lie in the effect of a significant fraction of short chains essentially plasticizing
Effect of external parameters 327

.10

e
~
.08 ..
......---~---.;---...-----
c • B
0
·iii
c .06
Ol
Eo-
Ol
Col
.;:....
Vl=- .04

.02
0 20 40 60 80 100
Temperature (0C)

Figure 6.20 A comparison of theoretically predicted values of r with experimentally


determined values for monodisperse PS, MW = 1 150000, at two strain rates:
4 x 1O- 6 s- 1 (solid curve and open squares); 1O- 2 s- 1 (dashed line and filled
squares). (Reprinted with permission from Plummer and Donald, 1990a.)

the material, such plasticization having been suggested to have a very


marked effect on the surface drawing mechanism (Brown, 1989).

6.6.2 ORIENTATION
Comparatively little work has been done systematically on the effects of
orientation on the crazing response, despite the fact that many glassy
polymers are utilized in an oriented state due to their processing history (see
also the discussion in section 5.2.6 on orientation). Many years ago
(Beardmore and Rabinowitz, 1975), it was first noted that crazes form at
higher stresses when the tensile axis is parallel to the molecular orientation
direction than when perpendicular. This effect translates itself into signifi-
cant anisotropy in the bulk mechanical properties such as GIc (Broutman
and McGarry, 1965; Curtis, 1970). Extremely large differences - as much as
an order of magnitude, but depending on the degree of molecular orienta-
tion - have been recorded in the two values of G1c determined parallel and
perpendicular to the orientation. Similar behaviour is observed generally in
large strain studies as described in Chapter 5.
Farrar and Kramer were the first to examine the internal microstructure
of crazes grown in oriented material (Farrar and Kramer, 1981). In these
experiments on thin films of PS, it was difficult to produce films of
328 Crazing

controlled preorientation with uniform thickness. Thus only a single


orientation was examined. However, this conclusively showed that the
fibrillar structure in 'perpendicular' crazes (i.e. those grown perpendicular to
the prior orientation direction) was much sparser than those grown in
parallel crazes.
More recent experiments, in which a controlled and variable degree of
molecular orientation was introduced by using diamond shaped copper
grids on which the PS samples were stretched above ~, have explored the
variation in craze microstructure in a much more systematic way (Maestrini
and Kramer, 1991). First, it is apparent that the craze fibril extension ratio
A. for perpendicular crazes increases steadily with the degree of preorienta-
tion, whereas it decreases for parallel crazes. This finding is clearly consist-
ent with the earlier work (Farrar and Kramer, 1981; Figure 6.21). Second,
using the method of low angle electron diffraction referred to in section 6.2,
both the fibril dimensions and the nature of cross-tie fibrils were examined.
It was found that Do, the interfibrillar spacing, was essentially unchanged
by orientation, but that the fibril diameter in parallel crazes increased
markedly as the molecular orientation increased. For perpendicular crazes
there was only a small decrease in D. In addition, there were comparatively
few cross-tie fibrils for parallel crazes, and they were essentially perpendicu-
lar to the fibril direction, which itself lay fairly accurately parallel with the
tensile axis. This was in contrast to the case for perpendicular fibrils where
the fibrils were misaligned by a few degrees with respect to the tensile axis,
and the cross-tie density was much higher.

5 •
4 • • • •
b
!
« 3 Il Il
Il
Il Il
2

0+----,---,----.---.----+
1 1.2 1.4 A. 1.6 1.8 2
or
Figure 6.21 Craze extension ratio Acr as a function of the preorientation extension
ratio Aor . fl, samples strained parallel to the preorientation; e, samples strained
perpendicular to the preorientation. (Redrawn from Maestrini and Kramer, 1991.)
Effect of external parameters 329

The current understanding concerning the origin of the cross-tie


fibrils has been discussed by Kramer (Kramer and Berger, 1990). They
arise when several polymer strands are strongly stretched at the craze
interface such that they are too strong to be broken. In this situation
it is energetically favourable for the craze to widen while bypassing
this taut clump. In this way a cross-tie fibril is left behind, bridging
between the two main fibrils. Once the craze has widened past the
cross-tie fibril the stresses on these strands will relax, pulling the two
fibrils together and causing the misalignment of the main fibrils referred
to in section 6.2. The higher frequency of cross-tie fibrils in the perpendicular
direction can be understood within this framework: there are more
chains, highly stretched but perpendicular to the fibril axis, which
are being bypassed in the formation of the craze than is the case
for parallel crazes.
Returning now to the higher stresses required to propagate crazes in the
parallel case (Haward and White, 1969; Beardmore and Rabinowitz, 1975),
it becomes apparent that the model presented in sections 5.2 and 6.1 must
be modified. Equation 6.22 would suggest that if Do is unchanged, as the
orientation changes as the experimental evidence shows, since there is no
obvious reason to expect r to change (if anything, one might expect more
scission to be required in the perpendicular case, making Scl. > SCII) there
should be no difference in the crazing stress as the orientation changes.
Maestrini and Kramer have argued that the problem arises in the evaluation
of the stress at the base of the fibrils. The simple model assumes that the
polymer here flows in the same way regardless of orientation, but in fact
allowance should be made for the differing strain hardening characteristics
of chains oriented at different angles to the externally applied stress. If
this is taken into account, an appropriate modification of equation 6.22
ensues, which is then consistent with experiment. However, it has to be
stressed that the nature of the flow in the so-called active layer adjacent to
the fibrils is still not fully understood at a molecular level. A special case of
the effect of preorientation is seen in the manifestation of so-called 'crazes
II' (Dettenmaier and Kausch, 1980, 1981). These were first observed and
most studied in polycarbonate, but have been seen in other polymers. They
form at high temperatures and strains, well past yield. Their microstructure
is fibrillar, but very much coarser than the structures typical of the crazes
discussed so far in this chapter, with fibril diameters upwards of 100 nm and
Vf ~ 0.8. The precise microstructure depends on the extent of pre orientation.
It is believed that these crazes form as an intrinsic phenomenon under
conditions when the entanglement network is nearing full extension. At high
temperature, slippage of the entanglement network may occur to give an
instability. This view is supported by the fact that as the molecular weight
330 Crazing

is increased there is a decreasing tendency for crazes II to form. An overview


of the phenomenon of crazes II is given by Dettenmaier (1983).

6.7 CRAZING IN THE PRESENCE OF SMALL MOLECULES


There are various different types of crazing that fall into this category. True
environmental (or solvent) crazing describes the situation when there is
some external environment that affects the crazing response. This could be
either a liquid or a gas, and both have been studied, although predominantly
at a macroscopic or phenomenological level. In addition, more recently
study has also been directed at the case where the polymer may contain
some small molecule. This may be either deliberately added as some sort of
plasticizer, or present by virtue of the molecular weight distribution of the
components, and it can affect the crazing response. A useful early review of
the first class of environmental effects is given by Kramer (1979a).
The propensity for glassy polymers to be affected deleteriously by low
levels of low molar mass fluids is well known. The early work of Kambour
(Bernier and Kambour, 1968; Kambour, Romagosa and Gruner, 1972;
Kambour, 1973) showed the relationship between the efficacy of a fluid as
a crazing agent and its solubility parameter. When the fluid's solubility
parameter matches closely that of the polymer, the fluid is a strong crazing
agent. Conversely, if the fluid has a very different solubility parameter it will
have little effect on the tendency of the polymer to craze. The reason for this
effect seems to be that the fluid in this case can act as a strong plasticizing
agent (Kambour, Gruner and Romagosa, 1973).
However, plasticization can only be important if the fluid can diffuse fast
enough through the polymer, and in glassy polymers this diffusion will not
occur by normal Fickian diffusion. Instead, so-called anomalous or case II
diffusion is known to occur in which the penetration distance depends
linearly on time, rather than with the square root dependence characteristic
of Fickian diffusion. The consequences of this have been unravelled in a
model due to Brown (1989), in which the Thomas-Windle theory for case
II diffusion (Thomas and Windle, 1980, 1981, 1982), as extended by Kramer
and coworkers (Hui, et al., 1987a,b) is married with the meniscus instability
for craze advance. In this model two regimes are identified: one in which the
craze growth rate is determined by the velocity of the diffusion front, and
one in which the meniscus instability operates sufficiently fast that it can
keep up with the diffusion front. One key unknown in this model is the effect
of an ext(frnally applied stress on the diffusion itself, although there is some
preliminary experimental evidence to suggest that such an effect exists
(More and Donald, 1992).
Crosslinking 331

Whereas, as indicated throughout this chapter, TEM has proved very


powerful for exploring microstructures of crazes grown in air, it is of much
less utility for solvent crazes. This is because TEM necessitates the study of
samples under high vacuum conditions, which leads to changes in micro-
structure as the fluid evaporates. Recent studies have therefore also used
scattering studies, predominantly SAXS (Paredes and Fischer, 1979; Brown
and Kramer, 1981) but there has been some limited small angle neutron
scattering work (SANS; Brown, 1987). The conclusion, on PS at least, is that
the fibril diameter is not greatly changed by the presence of a solvent. This
perhaps surprising effect was attributed to two compensating effects related
to equation 6.22. Since both the crazing stress S and the surface energy r
are known to fall due to plasticization, their ratio - which determines D -
may not be significantly changed (Brown, 1987).
As mentioned above, there is also the possibility that small molecules
present in the polymer itself may affect the crazing response. One example
of this is seen in samples of monodisperse PS whose molecular weight
distribution has been deliberately altered by adding low levels of low MW
PS. This has been shown to affect the craze micromechanics (Donald and
Kramer, 1983), and more particularly the case of craze breakdown (Yang,
et ai., 1986b), as will be discussed in section 6.9. Whether this means that
commercial glassy polymers could have their properties improved by
judicious tailoring of the molecular weight distribution, however, is not
known.

6.8 CROSSLINKING
In general it is believed that crosslinking leads to increased brittleness,
and very often this is true (Broutman and McGarry, 1965; Hinkley and
Campbell, 1983). However, it need not be the case, and understanding why
this should be may make better utilization of the technique possible. In
particular it is also possible to show that all the ideas presented so far in this
chapter can be extended to cover the case of crosslinked thermoplastics and
even, up to a certain point, thermosets, as has already been hinted at in the
data in Figure 6.15. For a polymer such as PS, we have seen that crazing is
driven by scission processes at room temperature, but disentanglement starts
to become increasingly important as the temperature is raised. PS is a
polymer with a low entanglement density, and it has been shown in section
6.5.2 that increasing ve can lead to an increased propensity for shear, which is
inherently a less damaging mechanism of deformation. Thus it is logical to
think that deliberately increasing ve by introducing crosslinks might actually
be advantageous, despite the early historical evidence suggesting the contrary.
332 Crazing

,•
./
4
A ADZ ./
./

.
..!'
• Acraze
./
./
3
:;

./

• •
./

.<
~ ./
./

2 ./ t;. t;. A
./ A
./ At;. A
./ At;.
1/
1 2 3 4
A
mIX

Figure 6.22 Plot of the extension ratio of crazes (e) and shear DZs (.1) versus Amax
for PS crosslinked to various extents. The dashed line corresponds to Amax = Acraz•·
(After Henkee and Kramer, 1985.)

This hypothesis has been shown to be correct by studying the deforma-


tion of PS films with controlled amounts of crosslinking introduced (Henkee
and Kramer, 1985). In this study the crazing response of PS with different
values of v - the sum of entanglement and crosslink density - was
examined. It was found that the crazing response at low values of v changed
to a regime where both crazing and shear (in the form of DZs for these thin
films) occurred, and finally only shear at the highest values of v. In Figure
6.22 these regions where crazes and DZs coexist is clear, and can be
identified with the region in Figure 6.15 (at room temperature) marked by
half-filled symbols covering the range of strand densities from 9-10 x 10 25
to 13 x 10 25 m - 3. Crosslinking also has the effect of increasing the strain at
which deformation is initiated.
In addition, it can be shown (Henkee and Kramer, 1985) that when the
ratio (ln A/In Amax) is plotted as a function of v, all polymers which exhibit
DZs fall on the same straight line. This point can be taken even further with
the realization that this finding extends even to epoxies sufficiently lightly
crosslinked that they still exhibit DZ formation (Glad and Kramer, 1991).
Epoxies (like other thermosets) are considered brittle: they do not usually
show extensive deformation prior to fracture. However, by reducing the
crosslink density an increase in shear can be seen.
Thus there is a complete spectrum of response possible by changing the
crosslink density. Increasing the propensity for shear either (as in PS) by
increasing v so that crazing is suppressed, or (as in the epoxies) decreasing
v from its normally very high value so that shear deformation becomes
possible for the first time, leads to improved mechanical properties. Even for
Crosslinking 333

PS, ultimately if v is increased too far the strain becomes more and more
localized until cracking intervenes with essentially no plastic deformation
preceding: thus there is an optimum strand density to achieve an improve-
ment in PS properties (Henkee and Kramer, 1985). Little systematic work
has been done on macroscopic mechanical properties to explore this route
to improved toughness. In particular, the studies of Henkee and Kramer
referred to here, deal with thin films only. In these, full triaxial stresses are
not usually developed and hence it is particularly easy for shear DZs to form
in preference to crazing. This phenomenon will not be carried over readily
to bulk samples, in which crazes may still form deep within a sample, unless
means of relieving the triaxial stresses are present. One possible route to this
is by the addition of rubber particles which may cavitate, but there appears
to be no work done to test this idea out.
The situation is rather different for polymers that normally craze by
disentanglement, such as PES. For such polymers, crosslinking may be
expected to shift the balance between shear and crazing, because the
disentanglement route becomes impossible as the chains form a network.
Only scission crazing remains a possibility. PES is a polymer which
normally exhibits a tough, shear-type response at room temperature, but it
is known that its mechanical properties are less good at long times and high

-..
~

~ 7
til
C
0 6.
c
0

......
~
e 5
oS:
"CI
4
oS:
.
.;
00
c
3

0 100 200
Temperature (OC)

Figure 6.23 The strain for deformation onset as a function of temperature in thin
films of PES/1 wt% sulphur (crosslinking agent) heated to 350 0 ( to control the gel
fraction <p. Strain rate = 1O- 2 s- 1 . x, brittle fracture; ., shear deformation; 0,
crazing; Y!!J1, mixed mode. (Reprinted from Plummer and Donald, 1991 with per-
mission.)
334 Crazing

temperatures than desired (Davies and Moore, 1987). The explanation for
this lies in the propensity for disentanglement (Plummer and Donald,
1990b). Since crosslinking will prevent this it may solve the problem of long
term failures. Studies of the deformation mode in crosslinked PES have
shown that this picture of craze suppression by crosslinking is correct.
Figure 6.23 shows how the mode of deformation changes as the gel fraction
<p (i.e. the fraction containing crosslinked chains) increases (Plummer and
Donald, 1991). Confirming this picture, as the gel fraction increases, the
range of temperatures for which crazing is seen decreases until for the
highest value of <p crazing is never seen alone. As was the case for PS, a
second effect of crosslinking is to increase the strains at which deformation
is first observed. However, on a practical note, it should be realized
that crosslinking in this way, although attractive in terms of improvement
in mechanical properties, may severely limit the process ability of the
polymer.

6.9 CRAZE FAILURE

6.9.1 INTRODUCTION
If a glassy polymer exhibits multiple crazing, but none of these crazes fail,
then the sample will still possess mechanical integrity. It is the breakdown
of crazes to form cracks that ultimately leads to failures. What is it about
the structure of crazes that causes them ultimately to break down? It was
shown a number of years ago that there was a critical craze opening
displacement beyond which a craze broke down at a craze tip. This view
was mainly derived from interferometric measurements and has been
discussed in some detail by Doll (1983). The value of this critical width
depended on both external parameters, such as crack velocity, and sample-
dependent parameters such as molecular weight. The initial conclusion was
that this finding was a consequence of fibril creep causing the final loss of
all entanglement and hence rupture of the fibrils. With the discovery of the
surface drawing mechanism for craze growth (Lauterwasser and Kramer,
1979), this explanation seemed less plausible. However, recent findings
suggest this view of fibril breakdown may be correct, although perhaps not
quite in the way originally envisaged (Yang et al., 1986a; Kramer and
Berger, 1990), and the picture is still incomplete. In particular it is now clear
that the craze does not fail at its midrib - the oldest and most highly drawn
part of the craze - as might be expected from the idea of fibril creep. On
the contrary, failure appears always to initiate at the edge of the craze where
new material is still being drawn in (Figure 6.24), with the midrib remaining
undamaged in the early stages.
Craze failure 335

Figure 6.24 TEM micrograph of the site of the initiation of a craze failure in PS. The
midrib can be seen to be intact, with the damage starting at the edge of the craze.

6.9.2 MOLECULAR ASPECTS OF CRAZE BREAKDOWN


When a craze fibril forms, as we have seen, there is a necessary entanglement
loss. It is possible to compute how many strands remain and what the
degradation of the molecular weight is as a consequence, for those polymers
which undergo crazing by scission. (There seems to be no data covering the
case of fibril breakdown when the initial craze is formed by disentangle-
ment.) The first part of the calculation involves enumerating the total
number of strands no in the initial material which goes into making the
fibril, the so-called phantom fibril (Kramer, 1983):

(6.23)
336 Crazing

The number of remaining effectively entangled strands ne after fibrillation is


then given by Kramer and Berger (1990):

(6.24)

where q is the fraction of entangled strands which do not undergo scission


during the fibrillation (the term in brackets represents a correction to
allow for chain ends). The quantity q has been calculated, and shown to be
~0.6 for PS at room temperature (Kuo, Phoenix and Kramer, 1985).
Based on this approach, it becomes possible to evaluate the median strain
at which craze breakdown should occur. It transpires that the prediction is
wildly out by comparison with experiment (Yang et al., 1986a), indicating
that other processes are occurring to cause fibril breakdown over and above
the necessary scission. To understand what these may be, it is helpful to
reconsider the location of the breakdown site. As shown in Figure 6.24,
initiation of the pear-shaped void is at the craze-bulk interface (this is
consistent with the so-called mackerel pattern observed earlier on fracture
surfaces as the site of fracture moves between one craze surface and the
other; Murray and Hull, 1970). In other words it is at the location of the
active zone where new material is being drawn in. It must be that here
additional breakdown of the entanglement network is occurring, either by
scission or disentanglement. In his review of this topic Kramer (Kramer and
Berger, 1990) comes down firmly in favour of disentanglement, rejecting the
idea of additional scission because there is no reason for this to be favoured
at this location rather than randomly along the fibril. His contention is that
in the active zone, before any strain hardening within the fibrils has occurred
to push up the value of (0' disentanglement (particularly if short chain
fragments are present) may be relatively easy locally.
Recent experiments, totally away from the area of crazing, lend support
to the idea that there may be additional mobility at this interface, although
not for the precise reasons mooted by Kramer. It has been shown, by
studying the expansion of ultrathin films by spectroscopic ellipsometry, that
the ~ of PS films thinner than about 100 nm is thickness dependent
(Keddie, Jones and Cory, 1994). The analysis of this finding suggests that in
the vicinity of a free surface, polymer chains possess additional mobility, and
this effect persists for some tens of nanometres (for PS the scaling length
describing the effect was found to be ~ 10 nm). Since the base of the fibril,
where the drawing is occurring, is of course a free surface, this result may
indicate why the mobility of the chains in the active zone is so high. Thus
the origin of the additional disentanglement at the craze-bulk interface may
be linked to this enhanced mobility. Additional experiments to verify this
hypothesis would be very illuminating.
Conclusions 337

6.9.3 PHENOMENOLOGICAL ASPECTS OF CRAZE BREAKDOWN

Many experiments have been carried out exploring the factors controlling
crack propagation into crazes using interferometry, as alluded to above
(Doll, 1983; Doll and Konczol, 1990; Schirrer, 1990) but, for the study of
molecular aspects, thin film work has again proved invaluable. A statistical
approach has been adopted, studying many individual grid squares, each
covered by a thin film which will behave independently. By slowly increasing
the strain, and determining at each strain the proportion of squares in which
(1) a craze has initiated, (2) a craze has begun to break down and (3) failure
has occurred with a crack spreading across at least half a grid square, it is
possible to explore how failure is controlled by parameters such as molecu-
lar weight and strain rate (Yang et al., 1986a,b). However, in order to obtain
meaningful statistics, it is desirable to work with ultraclean films so that
minute dust particles do not precipitate premature failure.
A convenient way to represent the data gathered in this way is to use the
strain at which 50% of the grid squares have crazed (cJ, have shown first
signs of failure (cb) and have fractured (cr), as a measure of the response.
Figure 6.25 shows data obtained in this way for two strain rates. It is evident
that the strains are pushed to higher values, particularly Cb and cr, as the
strain rate drops. The median plastic strain for fibril breakdown (cb - cJ is
a useful way of characterizing the breakdown of crazes in different samples.
In particular it has helped to show the dramatic effect short chains may have
on stability in a study of blends of monodisperse PS (Yang, et al., 1986b).
When blends are prepared of ultra-high MW PS (MW of 1.8 x 106 ) with
polymer with MW of 2000 (i.e. well below the entanglement molecular
weight), both Cb and Cf fall dramatically as the proportion X of the high
molecular weight component falls below about 0.3, with (l;b - cJ dropping
to zero as shown in Figure 6.26. The value of X = 0.3 does not correspond
to the proportion of high molecular weight chains at which the blend
average molecular weight between entanglements falls below Me for PS -
that is a much lower value of X - and that finding emphasizes the need for
a full molecular description of the type outlined in section 6.9.2.

6.10 CONCLUSIONS
This chapter has been devoted to a study of crazing, concentrating on the
molecular mechanisms. Crazing is of course the precursor of fracture, with
a crack propagating through a craze as the fibrils break down as described
in the last section. Thus if toughness is to be improved, what is needed is
either some strategy which shifts the balance away from crazing, or a means
of making the crazes that do form more resistant to breakdown. A third
338 Crazing

a)

• Ef
."..----------
20 •
Median • •
Strain
(%) E~


10 •

EC
0~--~~~==I6~-~~=-~6~~~6=~~=-~6~~~6=---
1~ 1~
Molecular Weight, M

b)

20
Median
Strain D D

::
(%)

?
D
10

qJ
i

• •
..- 0_ _ 0_6
10 5 106
Ec

Molecular Weight, M

Figure 6.25 Median strains for crazing 1'c(~)' fibril breakdown 1'b(e) and catastrophic
failure 1'1(0), (.) as a function of molecular weight at two strain rates: (a)
8 = 5 X 10- 7 S -1; (b) 8 = 3 X 10- 6 S -1. (Reprinted from Yang et at., 1986a, with
permission).

route, that of rubber toughening to encourage the formation of many crazes


(or extensive shear depending on the polymer matrix) prior to anyone
breaking down, is slightly different because the crazes themselves are
unchanged. This topic of rubber toughening will be covered in detail in
Chapter 8. These other strategies are really only examples of ideas underly-
ing the molecular picture presented in this chapter; however, as often as not
they have been arrived at empirically and an appreciation of why they work
may well have come after their successful exploitation.
It is clear that over the last couple of decades, tremendous progress has
been made in understanding the molecular processes that underlie crazing.
We have moved from a phenomenological appreciation to a much deeper
References 339

0.2,
t
• t
0.15

WU
I 0.1
w'"

0.05

0
• •
0 0.2 0.4 0.6 0.8
x
Figure 6.26 Plot of Eb - Ee for blends of monodisperse PS with molecular weights
of 2000 and 1800000 as a function of X the proportion of high MW polymer. The
strain rate was 3 x 10- 6 S-1. (Redrawn from Yang et at., 1986b.)

understanding which gives us some ability to make predictions. The import-


ance of the parameters of the entanglement network in determining the
crazing response of glassy polymers is firmly established. However, there are
some gaps in our understanding, in particular in moving on from concepts
extrapolated from behaviour in the melt to those strictly valid in the glass.
It is in these areas that work is most needed to complete the description of
the crazing process.

REFERENCES
Allison, S.W. and Ward, I.M. (1967) Br. J. Appl. Phys., 18, 115l.
Argon, A.S. (1973) J. Macromol. Sci. B. Phys., 8, 573.
Argon, A.S. and Hannoosh, J.G. (1977) Phil. Mag., 36, 1195.
Argon, A.S. and Salama, M.M. (1977) Phil. Mag., 36, 1217.
Beahan, P. Bevis, M. and Hull, D. (1975) Proc. R. Soc. London A, 343, 525.
Beardmore, P. and Rabinowitz, S. (1975) J. Mater. Sci. 10, 1763.
Berger, L.L. (1989) Macromolecules, 22,3162.
Berger, L.L. and Kramer, E.I. (1987) Macromolecules, 20, 1980.
Berger, L.L. and Kramer, E.I. (1988) J. Mater. Sci., 23, 3536.
Berger, L.L., Buckley, D.l. Kramer, E.I. et al. (1987) J. Polym. Sci., 25, 1679.
Bernier, G.A. and Kambour, R.P. (1968) Macromolecules, 1, 393.
Bessonov, B.1. and Kuvshinskii, E.V. (1960) SOV. Phys. Sol. State, 1, 132l.
Bowden, P.B. and Oxborough, R.I. (1973) Phil. Mag., 28, 547.
Broutman L.I. and McGarry, F.I. (1965) J. Appl. Polym. Sci., 9, 609.
Brown, H.R. (1979). J. Mater. Sci., 14, 237.
Brown, H.R. (1983) J. Polym. Sci., Phys. Edn, 21, 483.
Brown, H.R. (1987) Mater. Sci., Rep. 2, 315.
340 Crazing

Brown, H.R. (1989) J. Polym. Sci., Phys. Edn, 27, 1273.


Brown, H.R. and Kramer, EJ. (1981) J. Macromol. Sci. B: Phys., 19,487.
Brown, H.R. and Ward, I.M. (1973) Polymer, 14, 469.
Bubeck, R.A., Buckley, D.J., Kramer, EJ. and Brown, H.R. (1991) J. Mater. Sci. 26,
6249.
Bucknall, C.B., Karpodinis, A. and Zhang, X.c. (1994) J. Mater. Sci., 29, 3377.
Coates, P.D. and Ward, I.M. (1978) J. Mater. Sci. 13, 1957.
Coates, P.D. and Ward, I.M. (1980) J. Mater. Sci. 15,2897.
Curtis, l (1970) J. Phys. D: Appl. Phys., 3, 1413.
Davies, M. and Moore, D.R. (1987) ICI Internal Report.
Dettenmaier, M. (1983) Advances in Polymer Science, 52/53, 53 Springer-Verlag,
Berlin.
Dettenmaier, M. and Kausch, H.H. (1980) Polym. Bull., 3, 571.
Dettenmaier, M. and Kausch, H.H. (1981) Colloid. Polym. Sci., 259, 937.
Doi, M. and Edwards, S.F. (1978) J. Chern. Sci. Faraday Trans. II, 74, 1789, 1802,
1818.
Doi, M. and Edwards, S.F. (1986) The Theory of Polymer Dynamics, Clarendon
Press, Oxford.
Doll, W. (1983) Adv. Polym. Sci., 52-53. Springer, Berlin.
Doll, W. and Konczal, L. (1990) in Crazing in Polymers 2. Springer-Verlag, Berlin,
pp.215-62.
Donald, A.M. (1985) J. Mater. Sci., 20, 2630.
Donald, A.M. and Kramer, E.J. (1981a) J. Mater. Sci., 16, 2967.
Donald, A.M. and Kramer, E.l (1981b) Phil. Mag. A, 43,857.
Donald, A.M. and Kramer, EJ. (1982a) J. App/. Polym. Sci., 27, 3729.
Donald, A.M. and Kramer, E.l (1982b) J. Polym. Sci., 20, 899.
Donald, A.M. and Kramer, EJ. (1982c) Polymer, 23, 461.
Donald, A.M. and Kramer, EJ. (1982d) J. Mater. Sci., 17, 1871.
Donald, A.M. and Kramer, EJ. (1983) Polymer, 24, 1063.
Donald, A.M., Kramer, EJ. and Bubeck, R.A. (1982) J. Polym. Sci., Phys. Edn, 20,
1129.
Dugdale, D.F. (1960) J. Mech. Solids, 8, 100.
Edwards, S. (1967) Proc. Phys. Soc., 92, 9.
Farrar, N. and Kramer, E.l (1981) J. Mater. Sci., 22, 691.
Fellers, J. and Kee, P.F. (1974) J. Appl. Polym. Sci., 18, 2355.
Ferry, J.D. (1980) Viscoelastic Properties of Polymers, John Wiley, New York.
Fields, R.J. and Ashby, M.F. (1976) Phil. Mag., 33, 33.
Glad, M. and Kramer, E.l (1991) J. Mater. Sci., 26, 2273.
Goodier, J.N. and Field, F.A. (1963) Proc. International C01iference on Fracture of
Solids, Interscience.
Gurevich, G. and Kobeko, P. (1940) Rubber Chern. Technol., 13, 904.
Haward, R.N. (1942) Trans. Faraday Soc., 38, 391.
Haward, R.N. and Thackray, G. (1968) Proc. Roy. Soc. London A, 202, 453.
Haward, R.N. and White, E.F.T. (1969) Polym. Lett., 7, 157.
Henkee, C.S. and Kramer, EJ. (1985) J. Polym. Sci., Phys. Edn, 22, 721.
Hinkley, lA. and Campbell, FJ. (1983) J. Mater. Sci. Lett., 2, 267.
Hoff, E.A.W. (1952) J. Appl. Chern., 2, 441.
Hui, c.-Y., Wu, K.-c., Lasky, R.C. and Kramer, E.J. (1987a) J. Appl. Phys., 61, 5129.
Hui, c.-Y., Wu, K.-c., Lasky, R.C. and Kramer, EJ. (1987b) J. Appl. Phys., 61,5137.
Kambour, R.P. (1964) Polymer, 5, 143.
Kambour, R.P. (1969) J. Polym. Sci. A2, 7, 1393.
Kambour, R.P. (1973) J. Polym. Sci. D: Macro. Rev., 7, 1.
Kambour, R.P. and Russell, R.R. (1971) Polymer, 12, 237.
Kambour, R.P., Gruner, C.L. and Romagosa, E.E. (1973) J. Polym. Sci.: Phys., II,
References 341

1879.
Kambour, R.P., Romagosa, E.E. and Gruner, c.L. (1972) Macromolecules, 5,335.
Keddie, lL., Jones, R.A.L. and Cory, R.A. (1994) Europhys. Lett., 27, 59.
Kinloch, A.I. and Young, R.I. (1983) Fracture Behaviour of Polymers, Applied
Science, London.
Kramer, E.I. (1979a) Developments in Polymer Fracture-I, Applied Science, Barking,
55-120.
Kramer, E.I. (1979b) J. Mater. Sci., 14, 1381.
Kramer, E.I. (1983) Adv. Polym. Sci., 52-53, Springer, Berlin.
Kramer, E.J. and Berger, L.L. (1990) Adv. Polym. Sci., 91,2, Springer, Berlin.
Kuo, c.c., Phoenix, S.L. and Kramer, E.I. (1985) J. Mater. Sci. Lett., 4, 459.
Lauterwasser, B.D. and Kramer, E.I. (1979) Phil. Mag., 39A, 469.
Lazzeri, A. and Bucknall, C.B. (1993) J. Mater. Sci., 28, 6799.
Legrand, D.G. (1969) J. Appl. Polym. Sci., 13, 2129.
Legrand, D.G., Kambour, R.P. and Haaf, W.R. (1972) J. Polym. Sci., A2, 10, 1565.
Maestrini, C. and Kramer, E.I. (1991) Polymer, 32, 609.
Magalhaes, A.A.M. and Borggreve, R.I.M. (1995) Macromolecules, 28, 5841.
McLeish, T.c.B., Plummer, C.I.G. and Donald, A.M. (1989) Polymer, 30, 1651.
More, A.P. and Donald, A.M. (1992) Polymer, 33, 3759.
Murphy, B.M., Haward, R.N. and White, E.F.T. (1971) J. Polym. Sci. A2, 9, 801.
Murray, 1 and Hull, D. (1970) J. Polym. Sci. A2, 8, 583.
Paredes, E. and Fischer, E.W. (1979) Makromol. Chern., 180,2707.
Pitman, G.L., Ward, I.M. and Duckett, R.A. (1978) J. Mater. Sci., 13,2092.
Plummer, C.I.G. and Donald, A.M. (1989) J. Polym. Sci., Phys. Edn, 27, 325.
Plummer, C.I.G. and Donald, A.M. (1990a) Macromolecules, 23, 3929.
Plummer, C.lG. and Donald, A.M. (1990b) J. Appl. Polym. Sci., 41, 1197.
Plummer, C.I.G. and Donald, A.M. (1991) J. Mater. Sci., 26, 1165.
Prest, W.M. and Porter, R.S. (1972) J. Polym. Sci. A2, 10, 1639.
Schirrer, R. (1990) in Crazing in Polymers 2, Springer-Verlag, Berlin, pp. 263-300.
Steger, T.R. and Nielsen, L.E. (1978) J. Polym. Sci., Phys. Edn, 16, 613.
Sternstein, S., Ongchin, L. and Silverman A. (1968) Appl. Polym. Symp., 7, 175.
Taylor, G. (1950) Proc. Roy. Soc. London A, 201, 192.
Thomas, N. and Windle, AH. (1980) Polymer, 21, 613.
Thomas, N. and Windle, AH. (1981) Polymer, 22, 627,
Thomas, N. and Windle, AH. (1982) Polymer, 23, 529.
Wang, W.-c.v. and Kramer, E. (1982) J. Mater. Sci., 17, 2013.
Weidmann, G.N. and D6ll, W. (1976) Colloid Polym. Sci., 254, 205.
Wellinghof, S. and Baer, E. (1978) J. Appl. Polym. Sci., 22, 2025.
Xie, L., Gidley, D.W., Hristov, H.A. and Yee, AF. (1995) J. Polym. Sci., Phys. Edn,
33, 77-84.
Yang, A-C. and Kramer, EJ. (1985) J. Polym. Sci., Phys. Edn, 23, 1353.
Yang, A-C. and Kramer, E.I. (1986) J. Mater. Sci., 21, 3601.
Yang, A-C., Kramer, E.J., Kuo, c.c. and Phoenix, S.L. (1986a) Macromolecules, 19,
2010.
Yang, A-C., Kramer, E.J., Kuo, c.c. and Phoenix, S.L. (1986b) Macromolecules, 19,
2020.
Fracture mechanics 7
1. G. Williams

7.1 INTRODUCTION
This chapter will give methods for describing the fracture behaviour of
glassy polymers. The schemes are all macro in nature, such that certain
conditions have to be fulfilled before they give sensible values. The con-
ditions are usually expressed in terms of the relative values of parameters
and give rise to size criteria. The precise nature of the molecular mechanisms
involved in the processes are not important to the methods, providing they
can be characterized quantitatively. Insights can be gained from the param-
eters about the nature of such processes, but they are not necessary to the
methodology.
Fracture can be defined as the creation of surfaces within a solid, and as
such is distinct from deformation, which preserves continuity. Such a
definition is self-evident in a brittle fracture, but is less so in ductile fracture
where blunting accompanies crack growth. The violation of continuity is
critical in the analysis, however, since it is this which gives rise to energy
release and provides the driving force for fracture. The origins of the
schemes described here go back to Griffith [1] and Obrimoff [2] who
computed energy changes in brittle fractures and equated these with the
surface energy necessary to create those surfaces. Because the fractures were
brittle, the separation of the elastic processes generating the energy and the
separate mechanisms of surface creation did not present any problems.
Indeed Griffith demonstrated the relationship of his surface energy to
surface tension in the inorganic glass he studied. In most materials, and
certainly for most glassy polymers, fracture is accompanied by plastic
deformation and the effective surface energies are much higher than the true
surface energy. It is this factor, of course, which renders them physically
useful because of the enhanced toughness. In this case, however, the

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
344 Fracture mechanics

deformation of the body is involved in the release of the energy and its
dissipation. Some assumptions are necessary in order to proceed and Irwin
[3] concluded that one could treat this local deformation as if it was
separable and proceed as in the Griffith case. For this to be sensible,
however, the deformation would have to be localized and hence a length
scale invoked. This is the essence of elastic fracture mechanics and forms the
bulk of the methods used for glassy polymers. When the extent of the
deformation becomes significant with respect to other sizes such as thickness
and crack length, then the schemes become more complex and will be
addressed later in this chapter. First, however, it is sensible to consider the
elastic fracture case in some detail.

7.2 ELASTIC FRACTURE MECHANICS


This scheme of analysis is based on the assumption that the energy
dissipation processes at the crack tip are sufficiently local that the specimen
can be treated as an elastic body without the local dissipation. Thus elastic
analysis can be used to compute the energy release rate for the body at
fracture and hence the toughness, expressed as energy per unit area, may be
deduced. Figure 7.1 shows this local process zone of size rp where the energy
is dissipated. Within this zone plastic energy will be dissipated, but other
micro mechanics such as void growth and micro cracking may also occur.
The effects of all of these are simply summed within the measured fracture
toughness, Gc . This is achieved by performing an elastic analysis on the
body to compute the energy release rate G at fracture, thus determining Gc .
Such a parameter would have little value if it turned out to be different in
different specimens and crack forms and so one would hope that it is a basic
material property and, as such, transferable from one geometry to another.

Figure 7.1 Plastic zone at the crack tip.


Elastic fracture mechanics 345

Even if the elastic analysis works and we can find Gc ' there is no guarantee
of this.
The energy dissipation is a sum of different processes and one cannot
make a priori statements about its behaviour. A helpful notion here is to
postulate that these processes will be governed by the stress state within the
local zone. In general one would suppose that the higher the degree of
triaxiality generated, the more difficult it is for the material to dissipate
energy and the lower the toughness. The stresses within the zone are shown
in Figure 7.1, and around the crack tip a 1 = a 2 = a and the triaxiality arises
from a 3. For a small zone embedded in a large thickness B, the transverse
strain 1:3 = 0 and hence

a 3 = v(a 1 + a 2) = 2va
where v is the Poisson ratio. Constraint is conveniently expressed via the
Von Mises yield condition:
2a; = (a 1 - ( 2 )2 + (a 2 - ( 3 )2 + (a 3 - ay

and we may write


a 1
s=-=--
ay 1 - 2v

For v = 1 we have a hydrostatic tension field and s -+ ex and for v = ~,


s = 3. If B is not much greater than rp then a 3 -+ 0 and s -+ 1. In Figure 7.1
rp is shown increasing near the surfaces since s -+ 1, i.e. a plane stress
condition, at the free surface. A schematic diagram of the relationship
between sand Gc is shown in Figure 7.2 with Gc decreasing from a

G.
_ plane IItre8a

----------.~-~-~---------- ---
plane llraln


Figure 7.2 Toughness as a function of constraint.
346 Fracture mechanics

G

-------------~-----------
(i)

o r-- a. ---4.11-----· Aa a

Figure 7.3 Schematic resistance curve.

maximum plane stress value when s = 1 to a minimum plane strain value at


high constraint. Thus the material property is this relationship and not any
single value. If we wish to determine a conservative measure of toughness
we must look at plane strain conditions, but if we are interested only in the
behaviour of thin films then plane stress is appropriate. The complete
characterization of the full curve is uncommon, but recent work to define
constraints more precisely have been published, e.g. [4]. In most practical
schemes the extremes are measured and for glassy polymers the plane strain
condition is relatively easy to achieve since r p is small, i.e. '" 50 J.1m, so that
practical specimen sizes easily achieve plane strain.
A further factor of major practical importance is the definition of the
fracture event. In all fracture mechanics analyses it is assumed that an
existing crack is loaded until it grows. The condition when this growth
occurs is termed 'initiation', whilst subsequent growth is termed 'propaga-
tion'. If one plots a graph of Gc versus crack length, as shown in Figure 7.3,
then an ideal simple relationship is as shown in lines (i). The original crack
length remains at a o until G reaches a critical value, Gc ' after which all
subsequent propagation proceeds at this value. This is a reasonable expec-
tation for plane strain conditions since the stress state should remain
unchanged and any crack tip deformation is small. In some tests, however,
there is evidence to suggest that Gc increases during propagation, giving a
rising curve as shown in line (ii). In addition, a close examination of the
crack tip will reveal movement at quite low values of G, and often crack
blunting, which is not fracture, can be observed, as sketched in line (iii). It
is sometimes argued that one cannot practically separate blunting and crack
growth and that the behaviour can only be characterized by a resistance
Standard for linear elastic fracture tests 347

curve of the form

say, where A and N are constants and .1a is the crack growth. This
relationship combines together blunting and fracture, and initiation is
indeterminate. There is some truth in this claim in highly ductile fractures,
where it is usually employed, and it will be discussed later in section 7.4. For
elastic fracture, however, it is believed that the distinction can be drawn and
initiation is a real phenomena, although its precise definition is sometimes
difficult. The rising resistance, or R, curve is also a reality that is encountered
in some tests. However, it usually signifies a loss of constraint and an invalid
test for elastic fracture mechanics.
A further influential factor in testing is the stability of the loading system
used. This arises from the mechanics and will not be described in detail here,
but is important in that it determines the accessibility of parts of the
resistance curves. Load-controlled tests give energy release rates which rise
with crack growth since external work is fed to the system. In such systems
fracture initiation is often accompanied by instability, i.e. it goes 'bang', and
the problem of deciding on which point of the curve the measurement
should be made is avoided. In displacement control, as used in most very
stiff modern testing machines, the load drops with crack growth and the
crack grows in a stable fashion, and so the curve can be found. In some cases
stable growth can precede an instability, so great care must be exercised in
interpreting results.
Initiation is also greatly influenced by the initial crack. It is assumed that
a natural crack is formed and then reinitiated in the test. The creation of the
initial crack can often be quite difficult and lead to blunt cracks, residual
stresses or damage around the crack tip. All of these will tend to suppress
initiation, so that when it does occur it is at a falsely high value and will
give instability since, as the crack grows the resistance will fall to the true
value, but the driving force remains high.
All the factors outlined above influence the determination of toughness
and any practical scheme must, of necessity, be a compromise. This is
reflected in the current standards that are available and they will now be
discussed.

7.3 STANDARD FOR LINEAR ELASTIC FRACTURE TESTS


The tests method used for polymers are based on that developed for metals,
ASTM E399 [5], and is available as ASTM 05045 [6] or an ESIS
document [7]. Geometries which are predominantly bending are preferred
because they have higher constraint than tension. Thus three-point bending
348 Fracture mechanics

(a)

(b)

Figure 7.4 Test geometries: (a) three-point bend test and (b) compact tension.

and compact tension, as shown in Figure 7.4, are used. The latter is often
convenient since it uses less material, but the former is a simpler geometry
and is most often used. Details of the test such as specimen proportions and
roller support sizes are prescribed and appropriate calibration factors given.
Usually deep notches with ajW'" 0.5 are used, but other values are
permitted.
The initial notch is machined and then the final sharpening is performed.
For low toughness glassy polymers, and particularly thermoset resins, this
is done by tapping a fresh razor blade placed into the root of the notch. This
forms a natural crack ahead of the blade and is the preferred method of
notching. In general it will work well for glassy polymers, but some degree
of skill is involved and it is more difficult with opaque materials. For higher
toughness materials or when the yield stresses are lower, as in modified
materials, it is sometimes not possible to generate a crack by this method
and sliding the blade must be used. Pushing the blade into the notch should
be avoided since it generates residual compressive stresses which inhibit
initiation.
Specimens are loaded at a convenient rate, usually about 10 mm min - 1
which avoids inertia effects, but is not too time consuming, though the study
of rate dependence may require a wider range (section 7.6). The load
deflection curves to fracture are of the form shown in Figure 7.5. There is
Standard for linear elastic fracture tests 349

o Dltlplaoement
Figure 7.5 Load-displacement diagram.

often an initial displacement, due to slack in the loading system, but then
the curve is linear in the initial stages. There is then usually some non-
linearity leading to a maximum load which may be accompanied by an
instability. The materials are assumed to be linearly elastic, which they are
not, so the non-linearity can arise from various sources; there can be
viscoelastic effects, the evolution of the crack tip plastic zone or slow, stable
crack growth, which would all give non-linearity. It is, of course, possible to
make visual observations in transparent glassy polymers. but the iden-
tification of crack movement signifying initiation is notoriously difficult and
subjective. Similarly, deciding on when non-linearity has occurred is arbit-
rary, even if one was confident of its source. Because of these uncertainties
a compromise is adopted in which the initial compliance is defined, as
shown in Figure 7.5, and then a line for a 5% increase in that compliance
is drawn. If this intersects the loading line, then this load, designated
P50/o' is taken as the initiation point. If the maximum Pmax occurs within the
two lines, then that is taken as initiation. This definition is arbitrary, but can
be defined with reasonable accuracy. It would correspond to a crack growth
of about 2.5% in the geometries used if all the non-linearity came from crack
growth, and so in an 'R' curve as shown in Figure 7.3 the initiation point
would be geometry dependent. For valid elastic tests, however, the curves
are usually rather fiat, so the effect is not significant.
The two loads are also used to limit the degree of non-linearity that can
be tolerated in the linear analysis. For a valid test,

Pmax < 1.1


P5%
350 Fracture mechanics

to ensure reasonable accuracy. The appropriate load, designated PQ' is then


used to compute a stress intensity factor K Q , via

where J(ajW) is the calibration factor given in the standards [6,7]. This
provisional value is then tested for the size criteria to ensure plane strain
conditions, which is

a, W - a, B > 2.5 (::Y


where (Jy is a yield stress determined at an equivalent loading rate.
For polymers this is taken as the maximum stress in a tensile test, and for
glassy polymers a brittle failure will often precede yielding. In these cases a
compressive value is taken and then reduced by an empirical factor of 0.7.
If the value of KQ satisfies the criteria it is declared valid and becomes K c '
the critical stress intensity value.
The derivation of Gc requires a knowledge of the elastic modulus, and for
metals this is well defined so that the conversion from Kc via
Gc = (1 - v2)K~jE

where E is Young's modulus and v is the Poisson ratio, presents no


difficulties. For polymers E is generally not known and can vary sig-
nificantly with rate and temperature. It is therefore desirable to determine
Gc directly from the fracture test and this is done via the energy to fracture
V, as shown in Figure 7.5.

G = V
c BW0(ajW)

where 0 (ajW) is an appropriate calibration factor [7]. A major error in


this method is that part of the measured energy goes into local indentation
at the supports and load point, so that V is too high giving too Iowan
effective modulus. This is corrected by performing a subsidiary test as
indicated in Figure 7.6. A section of the bend specimen is supported by the
two rollers placed together and the load displacement diagram is measured
(and similarly for compact tension). The indentation energy Vi' up to P s%
(or Pmax) is then measured and the true value given by (V - Vi)' which is
used to find Gc . The modulus may then be determined via K~jGc and also
directly from a corrected compliance and used as a check on accuracy.
When the validity criteria are violated, one route to achieving validity is
obviously to increase size, but this can often be rather difficult because of
'i' testing 351

moulding problems and the uncertainty of maintaining homogeneity. One


option is to test at a reduced temperature, since Kc changes rather slowly
while (Jy increases, thus achieving valid results albeit at a lower temperature.
Increasing rate can have the same effect, but the problems with this
approach can be profound. There is evidence that the toughness of polymers
can be rate dependent in an absolute sense in that some of the energy
dissipation mechanisms are effective only above a certain time scale. Thus
time-temperature superposition is often not applicable and toughness
should be obtained as a function of rate, or time to fracture, directly.
Increasing rate gives rise to inertial effects in the recorded load and rather
strict conditions must be observed in order to use the static analysis. These
are embodied in an appendix to the ESIS static protocol [8], which gives a
procedure for tests up to 1 mls loading speed. The basis of the method is
that some form of damping must be added to the load point to restrict load
oscillations to within ± 10% of an apparent static curve. This curve is then
used to find Ps% and the static scheme used to find Kc and Gc. Significant
variations in toughness are noted in the range 10- 4 to 1 mis, i.e. fracture
times of lOs to 1 ms (section 7.6).

7.4 'J' testing


The origins of 'J' testing were in the need to determine plane strain
toughness values for metals when only small specimens were available.
When these were found to be invalid according to elastic fracture mechanics
criteria, a scheme was sought to determine valid values with such small
specimens. It was noted that specimens loaded mainly in bending and
containing deep notches achieved a high constraint factor even for a fully
plastic ligament. It was therefore proposed that an initiation toughness
352 Fracture mechanics

(3) (4 )

(a) (b)

1~·9.
O.2mm Aa (d)
(c)

Figure 7.7 Jtesting. (a) Energy measurement, (b) crack length measurement, (c) the
resistance curve with blunting and (d) the blunting model.

obtained under such conditions would be equivalent to that obtained in a


valid elastic test. The process of defining the initiation condition involved
measuring the resistance curve, and so such curves were obtained concur-
rently.
The method has been quite widely applied to polymers with some success,
e.g. [9,10] and an ESIS protocol has been developed for polymers [11]. This
is based on the multiple specimen method in which a set of nominally
identical specimens in either three-point bend, or compact tension, are
loaded to different displacements such that different amounts of slow, stable
crack growth occurs. The scheme is shown diagrammatically in Figure 7.7.
The specimens are loaded and then subsequently broken open after un-
loading, either at low temperature or by impact, or both, to reveal the crack
growth on the surface, as shown in Figure 7.7b. This growth is often not
obvious since other features such as plastic zones and crazes can be
mistaken for crack growth. There is also the possibility of artefacts from the
breaking open process. A useful check is to section specimens and then view
')' testing 353

the polished section when it has been partially reloaded. This can often
indicate the crack growth, which can then be checked against surface
features on breaking open. There are problems, however, which arise
because the crack tip plasticity induces local compression stresses and
associated crack closure. This gives rise to a cusp-shaped profile even on
reloading, so identifying the true crack tip is not always easy. Indeed in some
cases, and particularly those materials which craze, the uncertainties are
sometimes too great to use the method [11]. There is also the difficulty of
distinguishing between crack blunting and crack growth. As mentioned
previously for fairly flat resistance curves this is not a problem, but for very
tough materials the distinction between blunting and crack growth is not
sufficiently great to define initiation with any confidence. The form of the
data is shown in Figure 7.7c where J is computed from [11]

J = _2(_V_-_V..::..J
B(W - a)

for the three-point bend test (there is a similar expression for CT). V is the
energy up to the displacement at which the test was stopped, as shown in
Figure 7.7a and Vi is the indentation energy correction as discussed
previously. The resistance curve is shown in Figure 7.7c, together with the
blunting line which is drawn with a slope of 20"y, where O"y is the tensile yield
stress. This model has been quite successful for polymers [9J and is based
on the assumption that the crack tip is semicircular as shown in Figure 7.7d.
Thus the apparent crack growth due to blunting, Lla b , is half the crack
opening displacement (j and hence

J = (j0" y = (20" y) Lla b


The other two lines shown in Figure 7.7c are the linear and power law fits
to the measured points. The former can be used together with the blunting
line to give an initiation value J e , as shown. For this to be a sensible exercise
the slope of the line must be substantially less than the blunting line, and
for glassy polymers, which are not generally very tough, this is usually the
case. For many crystalline materials and some modified glassy polymers it
is not so.
The resistance curve is also described by the power law form as shown,
and this really precludes the notion of any initiation value. In the spirit of
the Ps% method, a characterizing value is used at a crack growth of 0.2 mm,
which is small but measurable. This is not initiation and so does not purport
to give an equivalent value to the elastic test.
An interesting version of the J test which rather simplifies the procedure
is to use an impact test to deliver the energy [12]. Once the energy required
to give complete failure is known, blows of less energy are then applied and
354 Fracture mechanics

the resulting crack length measured as before. This is not a true constant
rate test in that in each test the striker comes to rest, or bounces, but it does
give mostly high rate loading and is simple to perform. If an instrumented
striker is used, then indentation corrections can be applied as usual, but for
the very simple uninstrumented test, the potential energy is used to find J,
giving errors of about 10-20%.

7.5 ESSENTIAL WORK TESTS


Very thin materials are often difficult to test and, although both elastic Gc
and plastic J c tests can be applied, establishing initiation values and

p.&

t
w - ---ot p

(a) (b)
p.&

8(W-Z8)

(c)

Figure 7.8 The essential work test. (a) Specimen and zones, (b) load deflection and
energy and (c) determination of essential work we '
Essential work tests 355

resistance curves can present problems. The essential work test [13] has
been used to measure plane strain values [14] but its wider use is for plane
stress. Double-edge notch specimens are used and then loaded to give
complete plastic yielding in the ligament followed by complete tearing. The
load deflection curve is measured for this purpose and from it the energy U
to fail the ligament. The analysis assumes [13] that the energy required to
give failure can be partitioned into two parts. The first is the 'essential' work
we associated with the process zone and is a function only of the area
generated. This parameter would be expected to be the same as the initiation
value in a conventional test and be geometry independent. The rest of the
energy is assumed to go into the yielding of a zone whose size is propor-
tional to the ligament squared and is not expected to be geometry indepen-
dent. The two zones of energy dissipation are shown in Figure 7.8a together
with the force deflection curves in Figure 7.8b. The expression for the energy
is then written as
U= We' B(W - 2a) + PJt;,B(W - 2a)2
where Pis a shape factor and Jt;, is the plastic energy density. Thus
U
-~~
B(W -- 2a)
= W
e
+ PWp . (W - 2a)

and we would expect a linear relationship, as shown in Figure 7.8c with the
intercept giving We' The theory assumes these relationships and there is an
intrinsic problem in that generally the slopes are quite steep and the
intercepts small, so the accuracy is low. However, a protocol has been
developed by ESIS [15] which gives satisfactory results.
There is some justification for the assumptions via the linear resistance
curve:

J =J o+ (~~)L\a
where dJ/da is constant. The energy may be deduced from
ri(W- 2a)
U = 2B Jo J da

I.e.

U = JoB(W - 2a) + ( dJ)B


da "4 (W - 2a)2

and hence
356 Fracture mechanics

Load (N)

50

25

oE-------~--------~----~--·
0.15 0.30
Displacement (mm)

Figure 7.9 Load deflection diagram for a phenolic resin.

7.6 EXAMPLES OF FRACTURE DATA


The best examples of linear elastic behaviour are provided by the thermo-
set glassy polymers. In general they have Kc values in the range
0.5 - 1 MPam l/2 and moduli of about 3 GPa giving Gc values in the range
80- 300 J m - 2. Yield stresses are often difficult to define, but are generally
in the range 60-80 MPa. Critical thicknesses are thus in the range 0.1-
1 mm, so there is generally no problem in manufacturing specimens. Figure
7.9 shows a load-deflection diagram taken on a phenolic resin specimen,
where there is clear linear behaviour and a sharp load drop at fracture.
There is some initial curvature and some evidence of stick slip as the crack
propagated. Table 7.1 shows data taken from six such specimens with
dimensions B = 6mm, W = 12mm and span/width = 4. The average
KQ = 0.74 MPam l/2 and the yield stress of 65.2MPa gives a critical size of
0.32 mm, which makes the specimens valid in terms of thickness, and since
p rnax/P 5% = 1 then KQ = Kc. The variations in Kc are about ± 5% and in
Gc are ± 10%, as expected.
Thermoplastic glassy polymers generally give somewhat higher values
of toughness with poly(methyl methacrylate) (PMMA) having Kc '"
1. 7 MPa m 1/2 and Gc '" 1000 J m - 2. Polystyrenes are slightly lower with
Gc '" 400Jm- 2 while polycarbonate is higher at about 1700Jm- 2 • This
latter case results in the critical thickness increasing to about 3 mm, so care
must be taken choosing the specimen size.
Table 7.1 Fracture data for a phenolic resin

Specimen Load (N) alW KQ (MPa m- 2) VQ (mJ) U; (mJ) G(Jm- 2 ) Ell - v2 (GPa)

1 40.6 0.519 0.67 2.66 0.27 138 3.24


2 53.6 0.486 0.78 3.91 0.47 177 3.44
3 49.6 0.500 0.77 3.79 0.40 184 3.19
4 55.1 0.455 0.73 3.83 0.49 158 3.35
5 51.5 0.479 0.73 3.76 0.43 167 3.21
6 51.5 0.475 0.75 3.94 0.43 180 3.11
Average a 0.74 ± 0.04 167 ± 16 3.26 ± 0.11

aErrors are given in parentheses below each value.


358 Fracture mechanics

Load(kN)
Pmax
0.8

P S% ---------------

0.4

oL-------r_----~r_----~r_·
1.0 2.0 3.0
Displacement (mm)

Figure 7.10 Load-displacement plot for ABS (8 = 18 mm).

In glassy polymers higher toughnesses than these are usually achieved


only in modified materials, and a good example is ABS (acrylonitrile-
butadiene-styrene polymer). Figure 7.10 shows a load deflection diagram
for an 18 mm thick sample for which Pmax/P5% = 1.30, i.e. it is invalid. From
these data KQ = 2.3 MPam 1 / 2 and since E = 2.8 GPa, GQ = 1890Jm- 2 •
The critical size for this case is 8mm (cry = 40 MPa), so the values can be
regarded as valid in terms of thickness, but not so for linearity, and are thus
invalid overall. This same material was then evaluated by the J test in
18 mm and 6 mm thickness and the resistance curves are shown in Figure
7.11. The power law and linear fits to this data are given in Table 7.2.
Also given in Table 7.2 are the J O•2 values and the J c values obtained via
the blunting analysis. The power law indices are relatively low and there is
a large difference in 2cr y = 80 MPa and dJ Ida = 6 MPa, so that initiation is
quite distinct. The large degree of non-linearity observed here arises from
plasticity. This may be confirmed by noting that for the 18 mm thick
specimens the J value for 2.5% crack growth is about 5000 J m - 2 compared
with GQ of 1890 J m - 2, indicating that the non-linearity is not from crack
growth and that the GQ value is lower than that of Gc because of this. Also
shown in Table 7.2 are results from an impact J test, indicating a change in
form of the resistance curve and a decrease in J c and J 0.2 with increasing
rate.
Examples of fracture data 359

Table 7.2 Power law and linear fits to data for J test of ABS

J = A(l1a)N IdJ)
J= Jo + I\da l1a

B(mm) A N JO.2 Jo dJ/da Jc

18 7000 0.44 3450 2130 5180 2280


6 7500 0.45 3640 2260 6440 2460
6" 6820 0.56 2770 840 9470 1000
Units: J m- 2, /)"a in mm.
"Impact test.

Some data for other modified glassy polymers are given in reference 16.
A somewhat tougher ABS is reported with A = 15 500 J m - 2 and N = 0.7
so that J O•2 = 5000Jm- 2 . A toughened PPO (poly(phenylene oxide)) has
A = 12300Jm- 2 , N = 0.58 and J O. 2 = 4800Jm- 2 , a similar performance.
A toughened amorphous nylon has A = 35700Jm- 2 • N = 0.68 and
J 0.2 = 12000 J m - 2, an exceptional degree of toughness for a glassy polymer
and only slightly less than the toughness achieved in crystalline nylons.
Plane strain toughness values for glassy polymers, even when modified,
rarely rise above 5000 J m - 2, but under plane stress conditions very high

1.00 104 +-'-""'r-r-+-'~-.-f--.~~j--,..~""""'f---.-~~t~~~

oj
: :

BI\Ulting Line
1_ _._ . . _. . .~. __. .!...._..._....................J..............................,.....
......__.....1......__.............. _.--..
l i i :

~
o

..._- ....--............--..-..-.--..~.---.-...- ...-...~- ....-.··..····..·····-t·····....···..··

~: 0

.......-.....-....-...J.......-.-.--...-..l. .~. --.-..-.. .--1.--.-.. . .----..-..-1.-.. . -.. --....... . _. . . . . . . . . . . ..


, .
2.00 103 .-<>..--......... L. .-.. _-._.+. . _. __.__. . . _~.....--'""---------.
o 17Ox36xlB mm
[] BOxl2x6mm
0.00 +-"~--l-~~j......... . . . . . . .~===+::::==+=~-+
o 0.2 0.4 0.6 O.B 1.2
Crack Growth /)"a (mm)

Figure 7.11 Resistance curve for ABS.


360 Fracture mechanics

400
."
350
",/
if250
300

• .1"
c:
~ 200 .?
.1'/
::::r
15
~ 150

1100
"
c:

~
50
~
0
0 5 10 15 20 25 30
Ligament (mm)

Figure 7.12 Specific fracture energy versus specimen ligament length for PET
(8= O.2Smm; essential work = 21800Jm- 2).

values can occur. This is true of poly(ethylene terephthalate) (PET), which


is usually used in thin sections and exhibits remarkable toughness in this
form. Figure 7.12 illustrates this via the essential work test with data for
0.25 mm thick sheet. Specimens with W = 75 mm were tested and very good
linear data are obtained giving We = 21800J m - 2 and j3~ = 14.5 MJ m - 3
with a correlation coefficient of 0.998. In terms of J this would suggest

Figure 7.13 Variation of toughness with loading rate for a modified Pvc.
References 361

J 0 = 21 800 J m - 2 and dJ Ida = 58 MPa compared with 2eT, = 226 MPa for
the blunting line, i.e. a distinct difference.
As a final comment on the possible complexities encountered in modified
materials, Figure 7.13 shows data obtained on a toughened PVC at testing
speeds ranging from 20 mm s - 1 up to 20 m s - 1. All the tests were performed
on a hydraulic testing machine and the data for speeds above 1 mls - 1 used
a displacement method to find Gc [17]. Gc falls from about 9OO0Jm- 2 to
about 1000 J m - 2 at 1 m s - 1 after which it remains approximately constant.
This latter value, as we have seen, it about that expected for a glassy
polymer and indicates that the toughening effect is no longer active at the
high rates.

7.7 CONCLUSIONS
The methods that have been developed for characterizing the toughness of
polymers work well for glassy polymers in both modified and unmodified
forms. For unmodified materials, linear elastic fracture mechanics works
well and is a sound basis for comparing materials. For modified materials
the lower yield stresses give rise to more extensive plastic deformation and
schemes such as J are required but are successful. It has to be recognized,
however, that defining initiation and differentially crack growth from
blunting gives rise to uncertainties which have to be accepted as limitations
in the methods.

REFERENCES
1. Griffith, A.A. (1920) Phil. Trans. R. Soc.. London A, 221, 163-98.
2. Obrimoff, J.W. (1930) Proc. R. Soc., London A, 127, 290- 7.
3. Irwin, G.R. (1958) Fracture, Enclopaedia of Physics, Vol. 6, Springer, Berlin,
pp.551-90.
4. Shih, C.F., O'Dowd, N.P. and Kirk, M.T. (1993) in Constraint Effects in Fracture
(ed. E.M. Hackett), ASTM STPl171, pp. 2-20.
5. ASTM E399-90 (1990) Standard Test Method for Plane Strain Fracture Tough-
ness of Metallic Materials.
6. ASTM D5045-93 (1993) Plane-Strain Toughness and Strain Energy Release Rate
of Plastic Materials.
7. Pavan, A. (1996) A Linear Elastic Fracture Mechanics (LEFM) Standard for
Determining K IC and GIC at High Loading Rates, ESIS.
8. Williams, J.G. (1990) A Linear Elastic Fracture Mechanics [LEFMJ Standardfor
Determining Kc and Gc for Plastics, ESIS.
9. Hashemi, S. and Williams, lG. (1980) Polym. Eng. Sci., 26, 760-7.
10. Huang, D.D. and Williams, J.G. (1987) J. Mater. Sci., 22, 2503-8.
11. Hale, G. (1995) A Testing Protocol for Conducting J Crack Growth Resistance
Curve Tests on Plastics, ESIS.
12. Crouch, B.A. (1993) J Crack Growth Resistance Curve Tests on Plastics under
Impact Loading, ESIS.
362 Fracture mechanics

13. Chan, W.Y.F. and Williams, J.G. (1994) Polymer, 35, 1666-72.
14. Saleemi, A.S. and Nairn, lA. (1990) Polym. Eng. Sci., 30, 211.
15. Clutton, E.Q. (1994) Test Protocol for Essential Work of Fracture, ESIS.
16. Huang, D.D. (1991) ASTM STP 1114, pp. 290-305.
17. Adams, G.c., Bender, R.G., Crouch, B.A. and Williams, lG. (1990) Polym. Eng.
Sci., 30, 241-8.

[ESIS Protocols are available from lG. Williams, Imperial College of Science,
Technology and Medicine, Mechanical Engineering Department, Exhibition
Road, London, SW7 2BX, UK or from ESIS Office, c/o Materials Lab, Delft
University of Technology, PO Box 5025, 2600 GA Delft, The Netherlands.]
Rubber toughening 8
C.B. Bucknall

8.1 INTRODUCTION
Glassy polymers are by definition hard and non-crystalline. Many glassy
thermoplastics, and all thermosetting resins, also display another charac-
teristic of glasses - brittleness. The discovery that materials as unpromising
as polystyrene (PS) and epoxy resin can be toughened by the addition of
5-15% of a suitable rubber has therefore been of major importance to the
plastics industry. Indeed, rubber toughening has proved so effective that the
technology has been extended to almost all of the commercial glassy
thermoplastics, including poly(methyl methacrylate) (PMMA), poly(vinyl
chloride) (PVC), and even polycarbonate (PC), the toughest of the glassy
polymers; it has also been applied to several thermosetting resins other than
epoxies. In this chapter the prefix RT- will be used to denote rubber-
toughened polymers, except for polystyrene (PS) and poly(styrene-co-
acrylonitrile) (PSAN), where the acronyms HIPS (high impact polystyrene),
and ABS (acrylonitrile-butadiene-styrene polymer) have become firmly
established.
Toughness is the property of resisting fracture by absorbing and dis-
sipating energy. Strength, on the other hand, is the ability to resist high
stresses, and is obtained by suppressing deformation mechanisms, some-
times to the extent that the material becomes brittle under normal loading
conditions. Conversely, toughness in rubber-modified polymers is achieved
at some sacrifice of strength and stiffness. Whereas glassy polymers have
shear moduli of about 1000 MPa and Young's moduli in the range 2500-
3500 MPa, rubbers have very low shear moduli, typically between 0.1 and
1.0 MPa. The reductions in modulus brought about by adding rubber are
acceptable if they are accompanied by a substantial increase in toughness.

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
364 Rubber toughening

Before considering how this is achieved in toughened glassy plastics, it is


helpful to review the key properties of their constituents.
Attainable levels of fracture resistance in toughened plastics are deter-
mined largely by the properties of the matrix. Glassy polymers are capable
of dissipating large amounts of energy per unit volume through viscoelastic-
plastic flow, and strategies for toughening are therefore directed towards
maximizing the volume of material participating in this type of deformation.
The main problem is that stress concentrations around cracks, notches and
other geometrical features act to localize the deformation, often to the extent
of causing brittle fracture. Toughness develops only when the material
subjected to these high local stresses is able to respond by deforming and
strain hardening, a process which in polymers occurs through rotation of
the chain axes towards the direction of maximum extension, and is more
properly called 'orientation hardening', since it reflects the increasing
resistance of the network as it approaches its theoretical maximum draw
ratio Amax (Chapters 5 and 6). If orientation hardening is effective in
stabilizing the overstressed region against fracture, yielding spreads to the
surrounding material, and may eventually encompass a large fraction of the
specimen or component. The quantitative application of this principle to
yielding in polycarbonate has been demonstrated by Nimmer (1987) using
elastic-plastic finite element (FE) modelling.
Locally, orientation hardening inevitably involves a change in the true
stress acting on the highly deformed element. This is seen most clearly in
the formation of a macroscopic neck during ductile drawing of a tensile
specimen, but applies equally to the internal necking represented by crazing.
Polymers with low entanglement densities have high natural draw ratios
Amax (Donald and Kramer, 1982a,b; see also Chapters 5 and 6), and on
necking or crazing are therefore subjected to high true stresses, often in
excess of 200 MPa, which may be sufficient to cause fracture.
Fracture of polymers always involves some combination of chain disen-
tanglement and bond rupture. The critical stress required to break a given
microscopic element of a glassy polymer depends upon a number of factors,
including chain length, concentration of network points (entanglements plus
crosslinks), chain structure and loading conditions. If the chains are very
short and loading periods relatively long, failure may occur principally
through chain disentanglement and resulting flow (Trassaert and Schirrer,
1983), but this is rarely the main cause of fracture in commercial thermo-
plastics, because average chain lengths are chosen to ensure the formation
of stable entanglements. Rupture of the most highly stressed bonds therefore
plays a critical part in the fracture process, including both the formation and
failure of craze fibrils (Kramer, 1984; Sha et at., 1995).
Characterization 365

The concentration of network points determines both the natural draw


ratio and the distribution of stresses between individual chain segments. In
the ideal case, where all chains are fully aligned parallel to the tensile stress
direction (which means that the material is without entanglements), the
strength approaches the theoretical strength of the polymer chain, which is
of the order of 10-30GPa (Kelly, 1966; Vincent, 1972). In cases of more
practical interest, where the critically stressed elements are formed through
drawing of isotropic (chain entangled) glassy material, true stresses at
fracture are more typically between 0.2 and 0.8 GPa, and the bonds
undergoing rupture are carrying forces well in excess of the average. In this
connection it should be noted that polymer chains vary widely in cross-
sectional area, from 0.19nm 2 for polyethylene, through 0.33 nm 2 for poly-
carbonate to 0.74nm 2 for polystyrene (Vincent, 1972). The fracture resis-
tance of individual bonds is therefore strongly dependent upon the chemical
composition of the chains, a point of fundamental importance in any
discussion of rubber-toughened plastics, where the properties of the matrix
ultimately determine the performance of the blend.
As with glassy polymers, the intrinsic strength of rubbers is determined
by the cross-sectional area of the chains, which is usually small. Rubbers are
very compliant at low tensile strains, but exhibit very marked orientation
and network related hardening at high strains (Edwards and Vilgis, 1986):
natural rubber regularly reaches extension ratios greater than 9, at nominal
stresses above 30 MPa, which means that true stresses approach 300 MPa.
These properties may be called into play during the deformation of
toughened plastics, especially those in which the matrix is susceptible to
crazmg.
This chapter is concerned mainly with the effects of added rubber upon
the mechanical properties of glassy polymers, with particular emphasis upon
fracture resistance. Toughness is controlled not only by the basic com-
position of the polymer, but also by details of its morphology and chemical
structure, including such features as particle size distribution, crosslinking
in the rubber phase and grafting at the interface. Methods for characterizing
these parameters are therefore discussed first, before the central themes of
deformation mechanisms and fracture behaviour are addressed.

8.2 CHARACTERIZATION
Toughened polymers are complex materials which often contain three or
more different types of polymer chain, and several separate phases. The first
step in any characterization is to determine the compositions and concen-
trations of these phases. This can sometimes be done using the standard
366 Rubber toughening

methods of chemical analysis, including FTIR and NMR, but there may be
difficulties if there is a degree of miscibility between the polymers. Other
methods, including measurements of relaxation behaviour and various types
of microscopy, therefore have an essential part to play in defining the
structure of toughened plastics.
Dynamic mechanical thermal analysis (DMT A) is useful in characterizing
the relaxations of the glassy matrix, including not only the IX-relaxation,
which marks the transition from the glassy to the rubbery state, but also the
secondary relaxations below Tg , which in some cases are associated with
changes in ductility. When, as in some commercial polymers, the matrix is
formed by blending two glassy polymers that have different glass transition
temperatures, there may be two separate IX-peaks. Any mixing on the
molecular level will cause the peaks to shift towards each other, and if there
is a high level of miscibility they will combine to form a single IX-peak at an
intermediate temperature. The best known examples are blends of poly(2,6-
dimethyl-1,4-phenylene) ether (PPE) with polystyrene, which have been
shown to be miscible (MacKnight, Stoelting and Karasz, 1971; Shultz and
Gendron, 1972). This property is exploited commercially: one well estab-
lished family of toughened plastics is made by blending PPE with rubber-
toughened polystyrene (Robeson, 1984).
Although matrix composition and relaxation characteristics are impor-
tant, the main aim of DMT A tests on toughened plastics is usually to
determine the glass transition temperature of the rubber phase, and the
general shape and size of the corresponding loss peak. This is sometimes
difficult because of overlap with the f3-transition in the matrix. The rubber's
IX-peak rarely coincides with that of the original rubber, for two main
reasons: (1) crosslinking and grafting reactions, which occur during manu-
facture, shift the peak to higher temperatures; and (2) differential thermal
contraction, which occurs when the blend is cooled from the matrix ~,
generates dilatational strains in the rubber, shifting the peak to lower
temperatures.
Undoubtedly the most important method for determining the concen-
trations and morphologies of the various phases present in toughened
plastics is transmission electron microscopy (TEM). Using a modern ultra-
microtome, it is relatively easy to prepare sections between 50 and 100 nm
thick for observation by TEM. The main problem is to ensure adequate
contrast between the phases. Fortunately, many of the rubbers used for
toughening contain carbon-carbon double bonds, which react readily with
osmium tetroxide, leaving the matrix polymer unaffected (Kato, 1965, 1967).
The literature contains numerous high resolution micrographs of OS04-
stained sections, showing the structure of polymers toughened with either
Characterization 367

poly butadiene or diene based copolymers. Other rubbers present a greater


challenge. In some cases Ru0 4 provides contrast that cannot be obtained
using OS04 (Vitali and Montani, 1980; Trent, Schienbeim and Couchmann,
1983; Montezinos, Wells and Burns, 1985; Li, Ness and Leung, 1996). In
others it is necessary to pretreat either rubber or matrix to make it receptive
to a heavy metal stain. For example, electron beam irradiation of PVC
blends creates double bonds in the PVC which can be stained with OS04
(Vesely and Finch, 1988).
One basic parameter that can be determined using electron microscopy is
the volume concentration of toughening particles. This is not necessarily
equal to the concentration of added rubber, because the particles often
contain either immiscible long chains or miscible short chains of glassy
polymer, formed during a grafting reaction. Immiscible chains form hard
subinclusions, which mayor may not have the same composition as the
glassy matrix. For example, one important class of toughening additives is
made by grafting a shell of acrylate rubber around a spherical core of lightly
crosslinked PMMA, and then grafting an outer shell of PMMA onto the
rubber layer. These 'hard-soft-hard' three-layer core-shell particles are
used to toughen PMMA (Lovell et al., 1991; Archer et aI., 1993), but are
also effective in PC or PVC (Lutz and Dunkelberger, 1992), In relation to
mechanical properties, the soft rubber shell and the hard PMMA core
together constitute the rubber particle, whereas the hard outer shell is
effectively part of the matrix.
Perhaps the best known examples of particles with composite structure
are the 'salami' particles found in HIPS (Bucknall, 1977). which typically
consist of four parts PS subinclusions to one part polybutadiene. The
importance of making separate measurements of polybutadiene rubber
content Cr and particle volume fraction o/p is obvious in this case. The
problems of determining o/p from thin sections of HIPS are discussed by
Maestrini et al. (1992). In an ideally thin section, the volume fraction can be
determined directly from the observed area fraction, a relatively simple task
using an image analyser, provided that the sample is large enough to be
statistically representative. However, major problems arise when the section
thickness is not small in comparison with the average particle diameter,
because the observed volume fraction is then higher than the true volume
fraction of particles o/p. In order to calculate o/p accurately under these
circumstances, it is first necessary to determine the distribution of true
particle sizes. Again there is a problem: true diameters are equal to observed
diameters only for particles with their centres lying in the plane of the
section. Maestrini et al. were able to obtain improved estimates of both
volume fraction and mean particle size in a single HIPS polymer by
368 Rubber toughening

blending it with PS to four different dilutions, making measurements at five


section thicknesses on each blend, and applying equations based on the
principles of stereology (Underwood, 1970).
The problems of deriving distributions of true particle size from obser-
vations on thin sections are even more daunting, which explains why almost
all authors are content to quote observed particle sizes. In order to make a
satisfactory analysis, it is necessary to begin with some assumptions, guided
by the data, about the form of the true size distribution. The quality of the
results depends very much upon the reliability of these original assumptions
(Correa, 1993).
The internal morphology of the rubber particles is another important
structural feature. At their simplest, particles are plain homogeneous rubber
spheres, but in most toughened plastics the morphology is more com-
plicated. Reference has already been made to particles containing rigid
spherical cores, or distributed glassy subinclusions (salami structure).
Other variants include alternating concentric shells of rubber and glassy
polymer (onion structure) and assemblies of rod-like or thread-like rubbery
elements localized in a glassy polymer domain (the 'can of worms' morphol-
ogy). Figure 8.1 illustrates some of the commonest morphologies found
in toughened plastics. One feature that is rarely revealed by electron
microscopy is the shell of hard polymer bonded to the outer surface of the

(a)
a (b)

(c) (d) (e)


Figure 8.1 Schematic diagram showing typical morphologies of rubber particles:
(a) soft core-hard shell; (b) hard core-soft shell-hard shell; (c) onion morphology;
(d) salami structure; (e) can of worms morphology.
Toughening mechanisms - principles 369

rubber through graft copolymerization or other reactions. This shell is


designed to be compatible with the matrix polymer, and is generally
invisible under the microscope. However, it contributes to the hydro-
dynamic volume of the particle, and affects measurements of particle size
and concentration made using solutions of the toughened polymer. Several
commercial instruments determine particle size distributions in liquid sus-
pension, by detecting changes in the electrical properties of the liquid as
individual particles pass between electrodes in a capillary (Craig, Quick and
Jenkins, 1977).
Standard rubber particles tend to be spherical in the relaxed state,
becoming deformed and distorted only as a result of melt processing or solid
phase straining. However, some particles are non-spherical even in the
undeformed state. Work by Hayashi and Nishi (1990) has extended the
analysis of morphological features in toughened polymers to include the
shapes and spatial distributions of disperse phases. A valuable tool in this
type of study has been the division of space into Voronoi polygons, which
are defined as spatial regions in which all points are closer to the centre of
the enclosed particle than to the centres of any neighbouring particle
(Tanaka, Hayashi and Nishi, 1989).
The measurement of gel content and swelling index is another important
method for characterizing rubber-toughened plastics, especially HIPS and
ABS. The polymer is dissolved in a suitable solvent and centrifuged at high
speed (lOOOO-20000rpm) in order to precipitate any crosslinked material.
This procedure usually removes all of the rubber, together with any chains
of glassy polymer that have become grafted to it during polymerization.
Depending upon conditions, some of the uncrosslinked glassy polymer
originally present as subinclusions may also be trapped within the rubber
gel. Grafting is important because it affects the adhesion between rubber
particles and matrix, while the swelling index indicates the degree of
crosslinking in the rubber phase. Thus crosslinking is necessary to prevent
disruption of the particles during melt processing, but its extent must be
strictly controlled, otherwise the effectiveness of the rubber as a toughening
agent may be reduced. Methods for calculating the molecular weight
between crosslinks from swelling index data on HIPS samples have been
developed by Karam and Tien (1985).

8.3 TOUGHENING MECHANISMS - PRINCIPLES


Rubber toughening of glassy polymers involves three principal irreversible
deformation mechanisms: shear yielding, crazing and rubber particle cavita-
tion, all of which take place against a background of small strain (visco)-
elastic deformation, which is essentially reversible. At small strains, both the
370 Rubber toughening

rigid matrix and the rubber phase can to a first approximation be treated
as linearly viscoelastic, and in many cases time dependence can be neglected
altogether. On the other hand, there are also conditions, especially near the
IX-transition of the rubber phase, under which the time and temperature
dependence of elastic moduli are of critical importance (section 8.5.9). In
order to develop quantitative models for toughening, it is necessary to
understand the effects of stress and temperature on all four deformation
mechanisms.

8.3.1 VISCOELASTIC BEHAVIOUR AT SMALL STRAINS


Rubber particles affect the elastic behaviour of glassy polymers in two main
ways: they reduce stiffness, and they set up inhomogeneous distributions of
stress within the glassy matrix. Rubbers are characterized by low shear
moduli Gr and high bulk moduli K r: typically Kr = 2000 MPa for a rubber,
as compared with Km = 4000 MPa for a typical glassy polymer at 23°C
(Ferry, 1980; Tabor, 1994). Stress concentrations are produced when the
stress field includes a shear component.
Davy and Guild (1988) used FE models to calculate elastic moduli for
materials containing a random distribution of identical spheres: they first
divided the material into Voronoi cells, each containing one particle, and
obtained a particle volume fraction for each cell; they then performed FE
analysis on a range of cylindrical cells having the same particle volume
fractions as the Voronoi cells. Guild, Young and Lovell (1994) applied this
method to the calculation of moduli and stress concentrations in rubber-
toughened PMMA (RT-PMMA) with different rubber particle concen-
trations and morphologies. This study demonstrated the importance when
dealing with rubbers of choosing realistic values of K r • The usual FE
procedures, which require the programmer to specify Young's modulus E
and Poisson ratio v, can give wildly erroneous results when dealing with
rubbers, unless Vr is chosen with care. Calculations based on Vr = 0.49992
give physically realistic results, whereas at Vr = 0.5 the bulk modulus
Kr = 00.
Early studies of stress concentrations near rubber particles were based on
Goodier's equations for an isolated spherical inclusion in an infinite solid
(Goodier, 1933). Application of these equations shows that under uniaxial
tension the stress concentrations reach a maximum of 2.05 for a spherical
void, and are a little lower ( '" 1.9) in the case of a well bonded, fully relaxed
sphere of polybutadiene (Bucknall, 1969). Obviously these stress concen-
trations tend towards 1.0 when the rubber is cooled through its glass
transition or, equivalently, when the IX-transition is reached by reducing time
scales by several orders of magnitude.
Toughening mechanisms - principles 371

Because of the high bulk moduli of rubbers, dilatational stresses within


toughening particles can be significant even at small strains. They arise
because the particles are well bonded to the matrix, and forced to deform
with it. The resulting shear stresses within the particle are negligible,
but an imposed volume strain of only 0.5% acting on a typical rubber
with Kr = 2000 MPa generates a mean stress (or 'hydrostatic tension')
am = 10 MPa. Indeed, volume strains of this magnitude can be generated
simply by cooling the toughened polymer from the ~ of the matrix.
Assuming that both rubber and matrix polymer are fully relaxed at the ~
of the matrix, the volume strain and mean stress in the rubber on cooling
to temperature T are determined by the temperature interval L\ T =
(Tg - T), the rubber concentration c and the two volume coefficients of
expansion am and ar. The following expression was obtained by Boyce,
Argon and Parks (1987) for the thermally induced volume strain L\vr in the
rubber particle:

(8.1)

When the particle itself has a complex structure, as in Figure 8.1, it is first
necessary to calculate its bulk modulus and expansion coefficient, and use
the results in this equation: full details are given by Boyce, Argon and Parks
(1987). Calculations, based on the typical values L\ T = 80 K, c = 0.2,
am = 2 X 10- 4 K -1, ar = 6 X 10- 4 K -1 and Kr = 2000 MPa, give a ther-
mally induced volume strain L\vr = 1.0%. Equation 8.1 shows that the
constraint imposed by the matrix decreases with increasing rubber content,
causing a reduction in (JL\Yr/JT)(1' This effect can be seen in dynamic
mechanical experiments on ABS by Morbitzer et al. (1982). At 2% rubber,
thermally induced dilatation is sufficient to lower the loss peak from - 80
to - 90°C, but at 30% rubber the shift is halved, giving a ~ in the dilated
rubber particles of - 85°C.

8.3.2 RUBBER PARTICLE CAVITATION


Evidence for cavitation in rubber particles has been available for many
years, and there has been a growing understanding and acceptance of its
importance. At first, electron microscopists were understandably cautious
about interpreting the voids seen occasionally in ultra-thin sections, which
might easily be artefacts introduced during cutting. It was only when
fracture surfaces of RT-epoxy resins were studied carefully in the SEM that
the extent of rubber particle cavitation became clear (Bascom et al. 1975).
372 Rubber toughening

In most cases where RT-epoxy resins display significant toughness, every


particle on the fracture surface contains a prominent hemispherical void.
The apparent size of these voids is deceptive, because a hole having
one-third of the radius of the particle occupies only 4% of its volume. Each
void on the fracture surface is surrounded by a depressed crater of rubber,
a feature produced when yielding in the epoxy matrix makes the particle
dilate, so that during fracture the two halves of the stretched rubber shell
retract where they are not restrained by the boundary wall. Failure occurs
at the particle-matrix boundary only when the rubber particles are both
very large and poorly bonded (Kinloch, 1985).
Cavitated rubber particles are easy to see on fracture surfaces of RT-
epoxy, which are otherwise fairly featureless, but in other rubber-toughened
plastics, ductile drawing produces more complex fracture surface patterns,
making clear observation of cavitated particles more difficult. Other tech-
niques, especially TEM, have proved more effective in observing particle
cavitation in toughened thermoplastics. In one of the earliest studies, Breuer,
Haaf and Stabenow (1977) used a combination of TEM, optical microscopy
and low angle light scattering to demonstrate that stress whitening in
RT-PVC is due to rubber particle cavitation on well defined planes, with
normals between 26 and 35° to the tensile axis.
The conditions for void growth in rubbers have been analysed by Gent
and Wang (1991), using an energy balance criterion. Their calculations
apply to macroscopic rubber blocks, and are based on the reasonable
assumption that large samples of rubber invariably contain small voids
about 1 J.1m in diameter. However, this treatment cannot be applied directly
to toughened plastics, because rubber particles much smaller than 1 J.1m in
diameter are observed to cavitate under dilatational stress. It follows that
there is a mechanism for void nucleation in rubbers which does not depend
upon the presence of occasional defects on the 1 J.1m scale, and which is
almost certainly inherent in the behaviour of rubbers themselves, at the level
of individual chain segments. Since resistance to dilatation in rubbers arises
almost entirely from weak van der Waals interactions (Tabor, 1994), and
shear occurs easily, it is not surprising that under high triaxial tensile
stresses the distribution of polymer chains within the expanded volume of
the rubber particles become unstable, giving rise to microvoids.
Lazzeri and Bucknall (1993) and Bucknall, Karpodinis and Zhang (1994)
adapted the energy balance approach of Gent and Wang (1991) in order to
model cavitation in homogeneous rubber particles, but assumed that no
defects were present initially. When a rubber particle forms a spherical void
of radius r, the energy required is the sum of a surface energy term 41tr 2 r,
and a stretching term 21tr 3 Gr F(A,r), where r is the specific surface energy of
the rubber, typically 35mNm- 2 , Gr is its shear modulus and F(Ar) is a
Toughening mechanisms - principles 373

function representing the work done in stretching concentric spherical


membranes surrounding the growing void, up to a maximum stretch Ar.
When the radius R of the particle is held constant during cavitation, F(A r)
lies between 0.7 and 1.3.
If R is fixed, the energy required for void formation must be supplied
entirely from the elastic energy stored in the rubber. This assumption
simplifies the presentation of the model, but neglects energy contributions
from the matrix, which may be considerable (Bucknall, Karpodinis and
Zhang, 1994). However, their inclusion does not alter the basic principle of
the analysis. As noted earlier, volumetric strain energy densities in well
bonded rubber particles may be relatively high. In a typical polymeric glass
with a Poisson ratio v '" 0.4, an imposed tensile strain of 1% produces a
volume strain of about 0.2% in the matrix, and slightly more in the rubber
particles. For a rubber particle of radius R and bulk modulus K" at volume
strain Av , the volumetric strain energy Vv(O) is

(8.2)

At constant R, cavitation reduces this energy to

(8.3)

The rubber recovers towards its original density, which is reached when the
imposed volume strain Av = r3jR3, but recovery is resisted to a greater or
lesser extent by surface and elastic stretching forces near the void.
Figure 8.2 shows the relationship, for a series of different particle sizes,
between volumetric strain energy V vCr) and void radius r, in a particle with
Av = 0.005, Gr = 0.3 MPa, Kr = 2000 MPa and r = 0.03 J m - 2. For con-
venience, normalized energy V jV 0 [== V v(r)/(V v (0)] is plotted against
reduced radius r/R. The steep rise in VCr) at high r represents the stage
beyond which the imposed dilatation has been completely relaxed: with R
fixed and r increasing, the rubber is forced into compression against the
surrounding matrix. For diameters D > 0.20/lm, the minimum in the VCr)
curve is lower than V(O), which means that cavitation is thermodynamically
possible. Furthermore, the energy maximum at very small void radii offers
only a minor barrier to cavitation, which is easily crossed with the aid of
thermal energy. The condition VCr) min = V(O) therefore defines a critical
diameter for cavitation, at the specified volume strain, for the specific case
illustrated in Figure 8.2. Conversely, at fixed particle size there is a critical
volume strain AVeril for cavitation. The relationship between AVeril and Deril
is illustrated in Figure 8.3 for a series of shear moduli Gr. This relationship
374 Rubber toughening

1.4

1.2

1.0

0
::::> 0.8
:3
>-
21
Q)
0.6
c:
W
0.4
K = 2GPa
G = 0.3 MPa
r = 0.03 Jm-2
0.2
6.v(0) = 0.5 %

0.0
0.00 0.05 0.10 0.15 0.20 0.25
Hole Radius fIR

Figure 8_2 Energy of cavitated particle as a function of void radius, calculated using
equation 8.3 for a series of particle diameters.

reflects the increasing importance of the surface energy term as the particle
size decreases_
It is important to note that volume strain is generated in a rubber particle
not only by mechanically imposed stresses, but also by thermal contraction
from the matrix ~, as indicated in equation 8_1. One effect of this thermally

;?
~
----6_
5_
c: 4_
3_ _

~Q)
1 2_ _ _

01~
E Shear Modulus G in MPa
::J
"0 0.1
>

Q1 1 10
Particle Diameter (11m)
Figure 8.3 Critical volume strain for cavitation of rubber particles (log-log scale),
calculated using equation 8.3, with K, = 2 GPa, r = 0.035 J m -2.
Toughening mechanisms - principles 375

induced dilatation is to lower the ~ of the rubber (Donth, 1992). When the
rubber particles are very small, as they are in many styrene-butadiene block
copolymers, the shift can be as much as -18 K (Bates, Berney and Cohen,
1983; Boyce, Argon and Parks, 1987). This is unusually large: a limit is
imposed by cohesive failure within the rubber phase, or adhesive failure at
the particle-matrix interface. Experiments on ABS by Morbitzer and
coworkers (1976, 1982) showed that the rubber phase dynamic loss peak
splits into two under certain conditions. Particles that have not cavitated
give a peak at - 90°C, while particles that have either cavitated or
debonded, allowing the rubber phase to return to its unstressed density, give
a peak at - 80°C. The fraction of particles undergoing cavitation decreases
with increasing rubber content, because of the reduced level of constraint
imposed by the surrounding matrix on shrinkage of the particles (equation
8.1).

8.3.3 SHEAR YIELDING


Shear yielding can best be analysed using the strain energy density criterion
of von Mises, which states that yielding occurs on reaching a critical value
of effective stress 0" c:
= (0"1 - 0"2)2 + (0"2 - 0"3)2 + (0"3 - 0"1)2)1 /2 >
O"e- 2 0"0 (8.4)

where 0" l' 0"2 and 0"3 are principal stresses. In metals the critical value of 0" e
is usually assumed to be constant, and equal to the uniaxial yield stress O"y.
By contrast, the yield behaviour of polymers is strongly pressure dependent,
and is described by a modified version of the von Mises criterion:

(8.5)

where 0"0 is the yield stress under zero pressure (pure shear) and O"m is the
mean normal stress, or negative pressure.
On the basis of work by McClintock (1968), later extended by Berg
(1970), Gurson (1977a, b) developed a different modification of equation 8.4,
as follows, to represent the pressure dependence introduced into yield
behaviour by the presence of a volume fraction f of well dispersed voids:

( 1 - 2fcosh ( 30" ) )1/2


O"c = 0"0 20": +F (8.6)

Lazzeri and Bucknall (1993) proposed combining the pressure dependence


represented in equation 8.5 with Gurson's criterion in order to describe
376 Rubber toughening

yielding in a cavitated polymer, as follows:

(8.7)

This reduces to equation 8.5 when j = 0, to equation 8.6 when Jl = 0,


and to the original von Mises criterion (equation 8.4) when both j and Jl
are zero. An alternative equation which combines the intrinsic pressure
dependence of polymers with Gurson's equation has been advanced by
Steen brink (1996). Both equations give similar relationships in practice.
In applying equation 8.7 to rubber-toughened plastics, it is important to
note that the value assigned to j should not be the true volume fraction of
voids, which may be quite small, but the effective volume fraction jeff' which
is invariably larger. For example, an elastic volume strain of 0.5% in a
typical toughened polymer containing 20 vol% of rubber particles may be
converted to voids occupying only 0.5% of the total volume, so that there
is no overall change in specimen volume. However, if all of the rubber
particles have cavitated, and consequently offer negligible resistance to
further dilatation, then the effective volume fraction of voids !err = 0.2.
In isotropic materials, and under linear elastic conditions, (Je is a distor-
tional stress, which produces a change in shape at constant volume, whereas
(Jm is a dilatational stress, which produces a change in volume at constant

shape. It is convenient to plot the von Mises yield criterion and its various
modifications in terms of these two fundamental stress parameters, as shown
in Figure 8.4 for a toughened polymer containing 40 vol % of intact rubber
particles, and for the same polymer at various levels of rubber particle
cavitation. In each case the material is elastic below the curve, and plastic
(i.e. yielding) above the curve. It is clear from this diagram that the effect of
voids in depressing the yield stress is much smaller in uniaxial tension than
at a crack tip, where the stress field is strongly triaxial. This observation is
important in discussing the fracture resistance of toughened plastics, a theme
which will be developed later in this chapter.
Figure 8.4 involves an element of simplification, because the time depen-
dence of yield behaviour is not included. A more complete description of
inelastic deformation would be obtained by adding a third coordinate axis,
representing strain rate. Shear yielding in polymers is a rate process, which
is described by the Eyring equation (1936). Most experimental evidence is
obtained from tests in uniaxial tension, which is usually chosen as the stress
parameter, but from other studies it is clear that both (Je and (Jm affect strain
rate E. The equation is therefore written in a more general form (Ward,
1983):
Toughening mechanisms - principles 377

40.----------------------------------___~
uniaxial

biaxial

10 20 30 40
Mean Stress (MPa)

Figure 8.4 Yield envelopes calculated using equation 8.7, with J1 = 0.39, for a RT
polymer containing 40 vol% particles, with different levels of cavtation.

..
8 = 80 exp
(AH - yVO"e -
- RT
QO"m) (8.8)

where AH is the activation energy, y is the stress concentration factor, V and


Q are activation volumes for shear and dilatational processes respectively,
R is the gas constant and T is the temperature.
Equation 8.8 implies that strain rates are constant when the applied stress
is constant. This is rarely true. For example, Bucknall, Partridge and Ward
(1984) showed that creep in both PMMA and RT-PMMA can be fitted
approximately to the Andrade (1910) equation: the time-dependent compo-
nent of creep, given by [c:(t) - 8(0)], increases linearly with (Bt)1/3, where B
is a constant for a given test. This behaviour appears to be typical of
toughened plastics in which shear yielding is the dominant deformation
mechanism. The stress and temperature dependent parameter B obtained
from creep tests in this way is a rate coefficient, which is comparable with
the rate coefficients of chemical kinetics, on which equation 8.8 is based.
From Eyring's original derivation, it is clear that the strain rate terms Band
eo in equation 8.8 should strictly be replaced with rate coefficients. In the
studies on PMMA and RT-PMMA mentioned earlier, In B was plotted
against creep stress 0" to give a series of straight lines, in accordance with
Eyring kinetics. More often the distinction between a rate and a rate
coefficient is ignored, and the Eyring equation is used to correlate yield data
378 Rubber toughening

from tensile tests at fixed strain rates, where (Jy represents only one point on
a complex stress-strain curve.

8.3.4 DILATATION BANDS


Gurson's model (1977a, b) applies to a porous solid containing a concen-
tration of interacting microvoids. It predicts the formation of planar
dilatation bands, which combine in-plane shear with dilatation normal to
the plane, as illustrated in Figure 8.5. The angle of the band to the major
principal stress varies with void content and stress rate. Theoretically, in a
void-free isotropic solid exhibiting no pressure dependence, the angle ()
between the shear band normal and the tensile stress is 45°. The pressure
dependence of yield in solid polymers reduces () to 38°, and additional
pressure dependence due to the presence of voids may reduce it to zero
(Lazzeri and Bucknall, 1995), at which point the dilatation band bears a
superficial resemblance to a craze. There are, however, important differences
between the two types of deformation band. True crazes contain intercon-
necting voids formed through cohesive failure in the matrix, usually accom-
panied by bond rupture (Kramer, 1984), whereas dilatation bands contain
discrete voids which in the case of rubber-toughened plastics are confined
to the rubber phase.
Good examples of craze-like dilatation bands have been observed in
RT-epoxy resins, by both Vee and Pearson (1989) and Sue (1992), and in
RT-PC by Cheng et ai. (1995). The effects of rubber particle cavitation on

Figure 8.5 Schematic diagram of a dilatation band, showing combination of


in-plane shear and tension normal to the plane.
Toughening mechanisms - principles 379

band angles near rounded notch tips in RT-epoxy resin have been reported
by Jeong et al. (1994). Polarized light microscopy is particularly effective for
observing dilatation bands in rubber-toughened glassy polymers because it
shows very clearly the directions of the shear bands, and their interaction
with the cavitated rubber particles. Van der Sanden, de Kok and Meijer
(1994) have illustrated this effect in a model porous specimen consisting of
a polycarbonate film pierced by a random distribution of 3 Ilm through-
thickness holes. Indirect evidence for dilatational shear bands in RT-PVC
was earlier obtained by Breuer, Haaf and Stabenow (1977) using light
scattering.
A series of studies by Argon, Cohen and coworkers (Argon et al., 1981;
1983; Argon and Cohen, 1990) has shown that styrene-butadiene diblock
copolymers form craze-like dilatation bands, on planes normal to the tensile
direction, by repeated cavitation of the rubber domains ahead of the band
tip: Argon uses the word 'crazes' to describe these bands, but they are in fact
classic dilatation bands, formed by cavitation of either spherical or cylin-
drical domains of rubber, which typically have radii in the order of 15 nm.
Tensile tests give elongations of 5-40%, depending upon the chain length
of the PS block. Similar results were obtained by Van der Sanden, Meijer
and Lemstra (1993) in tensile tests on blends of PS with high concentrations
of rubber- PS core-shell particles.

8.3.5 MULTIPLE CRAZING


Multiple crazing was first observed in toughened plastics by Bucknall and
Smith (1965), who stretched thin sections of HIPS on the stage of a
polarizing light microscope, and found that yielding occurred by multiple
crazing in the PS matrix, accompanied by stretching of the rubber particles.
Whereas PS itself forms a small number of crazes and fractures at a
strain of '" 2 %, HIPS generates a high density of crazing, and may
reach strains of over 50%. This accounts for the observed effects of rubber
toughening in PS: stress whitening, greatly enhanced energy absorption,
substantial increases in volume and a high degree of recovery on unloading.
Later workers used TEM to observe multiple crazing in HIPS, ABS
and other toughened glassy polymers. In most cases the specimens were
deformed and unloaded before sectioning, with the result that recovery took
place both in the crazes and in the stretched rubber particles. The use of
OS04 as a stain may have accelerated this process, because it is known to
plasticize PS (Donatelli, Thomas and Sperling, 1973). Nevertheless, large
numbers of rather thin crazes can generally be seen in electron micrographs,
radiating from the surface of the larger rubber particles. Several TEM
380 Rubber toughening

Figure 8.6 Thin section of HIPS strained on the TEM stage. showing quantitative
fibrillation of salami particles and associated crazing of PS matrix. (Courtesy of R.C.
Cieslinski. Dow Chemical.)

studies on HIPS have reported that crazes initiate preferentially at particles


above '" 111m in diameter, and that initiation becomes more difficult (but
not impossible) as the particle size is reduced to O.2llm (Donald and
Kramer, 1982c; Hobbs, 1986; Keskkula, Schwarz and Paul, 1986; Okamota
et ai., 1991).
Rather more informative micrographs are obtained when sections or thin
films are stretched in situ on the stage of the TEM, and recovery is thereby
prevented (Beahan, Thomas and Bevis, 1976; Michler, 1992). As shown in
Figure 8.6, this technique reveals extensive craze-like fibrillation of the
rubber membranes within the salami particles, which is clearly associated
with crazing in the surrounding matrix. This combination of cavitation
mechanisms produces a series of continuous, approximately planar bands
running through the material. Indeed, in thick films it is sometimes difficult
to distinguish between sections of a band formed by crazing and those
consisting of fibrillated rubber.
Closer examination of the electron micrographs reveals that extension
ratios Ar in the rubber fibrils are '" 5, which means that they are carrying
significant stresses ( > 1 MPa), and are thus contributing to orientation
hardening in the cavitated band, especially when c/J p > 0.3. This craze
Toughening mechanisms - principles 381

bridging effect is very important: even a small reduction in the stress acting
on the craze wall has a large effect on the rate of thickening, which follows
Eyring kinetics, equation 8.8. Fibrillated rubber particles thus act to
stabilize the crazes that they have initiated. The fibrillation mechanism,
which is the result of constraints imposed on the rubber membranes by the
PS inclusions, ensures that almost every rubber molecule in a cavitated
particle eventually contributes to craze stabilization. By contrast, void
formation in large homogeneous rubber particles establishes a non-uniform
distribution of stress in the rubber phase, with very high circumferential
stresses near the void, where the rubber has to stretch to accommodate the
expanding cavity, and much lower streses near the particle-matrix bound-
ary. Under these conditions, dilatation takes place through tearing of the
most highly stressed regions of the rubber particles, while most of the rubber
phase makes little contribution to load bearing. Consequently, large homo-
geneous rubber particles are relatively ineffective as toughening additives in
PS (Donald and Kramer, 1982d) and in other glassy polymers that are
prone to crazing.
Multiple crazing has been observed directly by TEM not only in HIPS
but also in other rubber-toughened glassy polymers, notably ABS (Matsuo,
1969; Donald and Kramer, 1982e; Michler, 1992) and RT-PMMA (Frank
and Lehmann, 1988). In these materials, shear yielding also makes an
important contribution to deformation, especially at high temperatures and
low strain rates.
Crazes appear to be initiated in highly stressed regions of the matrix, near
the surfaces of the rubber particles, and especially at the equators of the
particles, where tensile stresses produce the highest stress concentrations
(Bucknall, 1969). Electron micrographs of lightly impacted HIPS specimens
show large numbers of crazes, all radiating outwards from the particle
surface (Okamoto et aI., 1991). It has generally been assumed that high
stress concentrations are in themselves sufficient to cause craze initiation, as
suggested by Haward and Owen (1973), but this view must now be
questioned in the light of current developments. Since shear yielding can be
initiated in toughened plastics by prior cavitation in the rubber particles, it
is at least possible that multiple crazing is also initiated by cavitated rubber
particles.
Experiments on craze-resistant toughened plastics show that the stresses
generated within the rubber particles, even under quite modest applied
stresses, can be sufficient to produce voids, and the analysis presented in
section 8.3.2 explains how this happens. Furthermore, crazes are known to
initiate at free surfaces and not in the bulk (Argon and Hannoosh, 1977)
and to propagate through a meniscus instability mechanism rather than the
382 Rubber toughening

formation of discrete voids (Argon and Salama, 1977). There is therefore a


strong case for considering rubber particle cavitation to be a necessary
precursor to multiple crazing in rubber-toughened plastics. This hypothesis
explains the well known particle size effect in HIPS, and helps to resolve a
number of other hitherto unresolved problems concerning the deformation
behaviour of toughened plastics, which are reviewed in section 8.5.
In support of the theory, recent work by Yang and Bucknall (1997) has
provided evidence that rubber particle cavitation precedes crazing in HIPS.
Specimens are subjected at 23°C to tensile prestrains of between 0.1 and
30%, annealed above the ~ of PS to heal the crazes, and retested at 23°C.
Prestrains of between about 0.25 and 5% produce a progressive reduction
in the yield stress of the annealed HIPS, but have no effect on the 'flow
stress', defined as the load minimum immediately following yield. Pre-
straining beyond 5% has no further effect on the yield stress after annealing.
Without annealing, prestraining below the yield strain, followed by un-
loading, has no effect on subsequent yield behaviour. These results are
consistent with the view that damage induced in the rubber during pre-
straining, and developed during annealing and cooling, increases the con-
centration of craze-nucleating cavitated rubber membranes when the
annealed HIPS is retested. The damage probably involves breaking selected
main chain chemical bonds.
Once initiated, crazes increase in area by propagating away from the
rubber particles, and at the same time thicken by drawing fresh material
from their walls. Lateral propagation is much the faster process. Argon,
Cohen and coworkers (Argon et al., 1990; Gebizlioglu et aI., 1990;
Piorkowska, Argon and Cohen, 1993; Spiegelberg, Argon and Cohen, 1994)
have shown that liquid diluents, including free polybutadiene of low
molecular weight (3000 g mol- 1) can act as solvent crazing agents, increas-
ing rates of craze propagation by several orders of magnitude. The effective-
ness of this mechanism decreases at high stresses and low temperatures,
because of the reduced diffusion coefficient of the liquid.
In toughened plastics, crazes are essentially planar yield zones surround-
ing a soft inclusion or cavity, a situation which has parallels in fracture
mechanics. On this basis, Donald and Kramer (1982c) have successfully
modelled early craze growth in diluted HIPS by treating rubber particles
effectively as linear cracks, and the crazes as Dugdale plastic zones. On
increasing the applied stress, or allowing more time for crazing to occur,
work is done on the specimen both to thicken and extend the crazes, and
to deform the surrounding matrix to accommodate the extra volume that
they require. As the crazes grow in area, the stiffness of the specimen
decreases, and the craze closure forces exerted by the matrix fall, reaching
zero when individual crazes lying approximately in the same plane combine
Toughening mechanisms - pnnciples 383

to form continuous fibrillated layers. This explains why creep rates in HIPS
are initially very low, and increase with time (Bucknall and Clayton, 1972;
Sjoerdsma and Heikens, 1982).
Once these crazes have extended to occupy the whole cross-section of
the specimen, they can expand only by fibrillation from the craze walls,
which occurs in parallel with fibrillation of the rubber and stretching of
the rubber fibrils. In a creep test the craze closure stresses exerted by
extended rubber fibrils may be sufficient to reduce the rate of deformation
progressively almost to zero as the total strain increases. This effect has been
observed in HIPS materials with relatively high rubber particle volume
fractions (<p p > OJ).
The observation that rubber fibrils normally reach large extensions and
high stresses suggests that they might control the final failure of the material.
It has usually been assumed that fracture of HIPS, as in PS, is precipitated
by breakdown of the crazes, but the precise mechanisms of failure in HIPS
are not well understood. Most of the evidence comes from standard tensile
tests at constant strain rate, in which the applied stress rises to a maximum
at the yield point, falls to a minimum (the flow stress) at about 5% strain,
then climbs slowly until the specimen fractures. From these tests it is not
clear whether stress or strain is the limiting factor.
This question has now been resolved by Sjoerdsma and Boyens (1994),
who tested specimens of a single HIPS polymer, in batches of 20-30 at each
of a series of fixed tensile stresses, and found that Pr, the probability of
fracture, increases with craze strain ecr (i.e. total strain minus elastic strain):
(8.9)
where the constant Q is independent of stress and temperature over the
ranges 15-18 MPa and 13-40°C. This work clearly identifies strain as the
critical factor controlling fracture in HIPS. Fracture becomes possible as
soon as the first craze forms, but becomes increasingly likely as the strain
increases. Sjoerdsma and Boyens (1994) suggested that the relationship
described by equation 8.9 arises from interactions between neighbouring
crazes, but it is more likely that it reflects the increasing probability of
failure in the rubber fibrils as they are stretched to higher strains.
In standard tensile tests, applied stresses adjust to maintain a constant
strain rate, and the rise during the later stages of the test is attributable to
the increasing stresses carried by the rubber fibrils. The maximum strain to
failure of HIPS, er(max)' is approximately proportional to the rubber particle
concentration <Pp (Buckley, 1993), and is greatly reduced when the rubber
phase is crosslinked (Soares, 1994; Bucknall et al., 1996). These observations
are consistent with the view that failure of the rubber fibrils causes final
fracture in HIPS.
384 Rubber toughening

The factors controlling tensile fracture in plain strips of rubber have been
studied extensively by Smith (1958), who found that failure data obtained
over a wide range of temperatures and strain rates could be fitted to a single
failure envelope. It is thus possible to define a maximum extension ratio
Ar(max) for the rubber fibrils, which is a function of the degree of crosslin king
(Taylor and Darin, 1955).
One further final key observation on the deformation and fracture
of HIPS comes from small X-ray scattering (SAXS). Buckley (1993) found
that 'I' en the contribution of crazing to the total volume change during
tensile impact, remained approximately constant, at about 45%, throughout
most of the test, over a wide range of rubber particle concentrations
(0.09 < ¢p < 0.36). It follows that An the average extension ratio in the
rubber fibrils, increases linearly with total strain e, until the material
fractures. Assuming that there is no significant shear yielding, and that e is
the sum of contributions from elastic deformation eel' rubber fibrillation and
associated crazing, then emax ' the maximum attainable strain for a HIPS
with rubber content cr ' is given by:

Cr(Amax - 1)
emax = eel + 1 _ 'I' (8.10)
er

Equation 8.10 represents an upper limit on the strains that can be reached
in HIPS under craze saturation conditions, when all of the rubber mem-
branes are fully fibrillated and stretched close to their limit, and the
corresponding network of crazes has been developed in the matrix, as
illustrated in Figure 8.7. The maximum possible strain is reached when
A = Amax for all particles, a condition that is likely to be approached only
when ¢p is high. It is, of course, possible that the HIPS sample will break
well short of the maximum attainable strain, through the failure of a few
overstressed rubber particles. However, the stress imposed upon a given
rubber membrane is limited by two factors: (1) while there are still
membranes of rubber within the particle that have not fibrillated, the stress
cannot rise above the level needed to produce additional fibrillation; and (2)
at relatively high particle volume fractions, the most highly strained particles
produce a marked reduction in the local rate of strain, by imposing closure
stresses on the neighbouring crazes, thereby keeping A lower than it might
otherwise be. If, on the other hand, the volume fraction of salami particles
is not high, then the material as a whole is unlikely to reach high strains. At
low ¢p the mechanisms for stabilizing crazes in regions of high strain are
ineffective, and in their absence these crazes will continue to thicken until
some of the associated rubber particles reach their strain limit and break,
while other regions of the polymer have relatively few crazes.
Toughening mechanisms - principles 385

Cavitated
Salami
Particle
Crazes Crazes

7 "\
I I

Figure 8.7 Schematic illustration of a fully fibrillated salami particle in HIPS, with
associated crazes.

Several attempts have been made to develop general criteria for crazing
that are based on stress invariants, and can therefore be compared with the
von Mises criterion for shear yielding. However, there are difficulties in
treating crazing as a continuum deformation process, because careful
experiments have shown that crazes do not form spontaneously within a
defect-free glassy polymer . Like cracks, they are almost invariably initiated
at a free surface, which may be an internal void or the external surface of
the sample. The most successful theory of craze formation in the absence of
obvious cracks or flaws was proposed by Argon and Hannoosh (1977), who
concluded that crazes initiate through the development of a porous zone
at stress concentrations near a free surface. This involves both a microshear
process, driven by the effective stress a e' and a dilatational process, driven
by am' Once a cavitated nucleus has been formed in this way, the craze
grows outwards by the Argon-Salama (1977) meniscus instability mechan-
ism, as discussed in Chapter 6. Craze growth is much better understood than
craze initiation, and is driven basically by the principal tensile stress,
essentially following Eyring kinetics (Argon et at., 1983; Kramer, 1984).

8.3.6 DISTINGUISHING BETWEEN MECHANISMS


Since yielding of rubber-toughened polymers can involve several different
deformation mechanisms, all of which can occur simultaneously, it is
important to develop experimental methods for distinguishing between
them, and for quantifying their individual contributions to the overall strain.
The most obvious starting point is microscopy, which provides detailed
information about the morphological features of deformation bands, and
their interaction with rubber particles. Electron microscopy is especially
effective in studies of crazing and rubber particle cavitation, whereas
386 Rubber toughening

polarized light microscopy is often better for identifying and characterizing


regions of shear deformation. Neither technique is really suitable for
quantifying the amount each mechanism contributes to a given deformation.
Other experimental methods have been developed for this purpose.
The earliest and most widely used of these methods is tensile dilatometry,
in which volume changes are monitored continuously in tensile bars, using
either longitudinal and lateral extensometers in tandem (Bucknall and
Clayton, 1972; Bucknall, 1977), or a water immersion dilatometer (Coumans
and Heikens, 1980). Plots of volume strain against extension then quantify
the respective contributions of dilatational and shear mechanisms as func-
tions of strain. This method has been criticized because it does not
distinguish between the two possible dilatational processes - rubber cavita-
tion and multiple crazing - which often occur simultaneously. Nevertheless,
it does provide valuable information, especially about the kinetics of
dilatation under tensile loading.
An alternative, more specific method for monitoring multiple craze
formation is now available, based on SAXS. Because crazes are essentially
arrays of cylindrical polymer fibrils with their axes lying parallel to the
tensile direction, they produce characteristic scattering patterns, from which
the concentration and diameters of the fibrils can be determined. This is the
principle of a new method developed by Bubeck et al. (1991). These authors
carried out real-time SAXS on tensile impact specimens, using high speed
cinematography to determine volume changes. They concluded that non-
craze cavitation begins in HIPS at the same time as, or a little before, craze
formation, and accounts for more than half of the observed volume increase.
In ABS the contribution of crazing to the overall deformation is even
smaller.
Similar results have been obtained by Magalhaes and Borggreve (1995)
using real-time SAXS on HIPS tensile specimens deformed at lower rates:
in their experiments, crazing accounted for only about 25% of the total
post-yield extension. Non-craze cavitation in these two studies is obviously
due to void formation in the rubber phase. The small contribution of crazing
to yielding in HIPS is a little surprising in view of the TEM evidence, but
supports the view that cavitation begins in the rubber particles, and that
crazes spread outwards from the cavitated particles. It is important to note,
however, that the energy required to generate a given volume change in a
rubber particle is almost certainly smaller than that needed to produce the
same volume change by crazing. The SAXS technique has also been used to
observe dilatational shear yielding during tensile deformation of RT-
PMMA: in this material, the voids produce a characteristic elliptical
scattering pattern, with no evidence of crazing (Lovell et al., 1993, 1994).
Cavitation diagrams 387

8.4 CAVITATION DIAGRAMS

In order to understand the deformation and fracture behaviour of rubber-


toughened polymers over a range of temperatures, strain rates, loading
conditions and specimen geometries, it is necessary to analyse the complex
interactions taking place between rubber particle cavitation, matrix yielding
and crazing. Cavitation diagrams, which were first introduced by Lazzeri
and Bucknall (1995), provide a useful basis for such an analysis. Their
function is to define conditions for shear yielding and other deforma-
tion processes, in terms of the von Mises effective stress ere and the mean
stress am.
The basic principles are illustrated in Figure 8.8, where for clarity a
separate diagram is used for each component of the analysis: shear yielding,
particle cavitation and crazing. In order to provide some perspective on the
magnitudes involved, the schematic diagrams are based approximately upon
the properties of a particular grade of RT-PMMA at 23°e.
Figure 8.8a defines the upper and lower bound envelopes for shear
yielding. Near the origin, the material is below its yield stress, and behaves
elastically. At higher stresses, between the upper and lower bound yield
envelopes, it is able to yield, but only if voids or cavitated particles are
present. Finally, above the upper bound envelope, it is unconditionally
plastic, able to yield even when no dilatation is possible. The upper and
lower bound yielding envelopes thus form a framework upon which ad-
ditional information, about particle cavitation, crazing and fracture, can be
superimposed.
Figure 8.8b defines the critical mean stresses required to cause cavitation
in rubber particles having typical properties (see caption to Figure 8.2) but
differing sizes. The data, based on critical volume strains calculated using
equation 8.3, show that reducing particle diameter causes a large shift in the
critical mean stresses at cavitation. One result of this shift is that large
particles produce cavitation before yield in uniaxial tension, whereas smaller
ones do not. The dotted line is the upper bound shear yield envelope.
Figure 8.8c quantifies the third element in the analysis: the critical
conditions for crazing and fracture. The stress field conditions governing
these phenomena are different in character from the criteria for yielding and
particle cavitation. Both the von Mises and the critical mean stress criteria
are defined uniquely in terms of the two stress invariants a e and am.
Consequently, the relevant curves in Figures 8.8a and 8.8b generate three-
dimensional yield or cavitation envelopes on rotation about the am axis
(Lazzeri and Bucknall, 1995). The same cannot be said about crazing, as
explained in section 8.3.5. In view of this difficulty, the onset of crazing is
40
.
o lo,.m 021'111 01 .m
(b)
fa
e
Q.
30
'"en
Q,I
.c
U)
Q,I
> 20
~
w
'"
Q,I
10
i'"
c
0
> 0 ~

10 20 30 40
Mean Stress (MPa )

40
(c)
fa
Q.
e 30 uniaxial biaxial

'"!!!'"
in
~ 20
."

~
W

'"
Q,I
10
'"
~ triaxial
c
0
> 0
10 20 30 40
Mean Stress (MPa)

Figure 8.8 Principles of the cavitation diagram: (a) shear yield envelopes; (b) rubber
particle cavitation; (c) onset of crazing from specimen surface (0) and from cavitated
particle (e). The stress levels chosen are purely illustrative.
Cavitation diagrams 389

defined here simply in terms of a critical tensile stress, (J llerit)' which is


equal to the craze drawing stress, acting locally on a small element of the
material.
Except in the case of pure triaxial stress «(J e = 0), it is impossible to define
a particular value of principal stress simply by specifying (J" and (Jm: further
information about the relative magnitudes of the principal stresses is
necessary in order to determine (J 1. One solution to this problem, adopted
here, is to work in reverse, by choosing stress states of interest (e.g. uniaxial
tension) and drawing relevant lines on the (Je - (Jm diagram. Points on these
lines then indicate where the major principal stress reaches the critical value,
(J l(erit)' at which crazes initiate.
Three important stress states are represented in Figure 8.8c: uniaxial
tension «(J,O,O); equi-biaxial tension «(J,(J,O) (which includes plane stress
crack tip conditions); and a state close to triaxial stress «(J,(J,0.8(J) that is
characteristic of plane strain crack tip conditions. Two points are identified
for each stress state: craze initiation from the surface of the specimen in the
absence of either particle cavitation or adventitious voids, and (at ap-
proximately half the stress) craze initiation from a large cavitated rubber
particle or equivalent void. In applying this component of the cavitation
diagram to sharply notched specimens, the very high stress concentrations
theoretically associated with notch tips present a potential problem. It can
be resolved by noting that the stresses in material adjacent to a crack tip
are initially zero, and that as load is applied to the specimen these stresses
follow the radial line on the cavitation diagram between plane stress (stress
ratio 1: 1 : 0) and plane strain (ratio 1: 1 : 0.8) until the stress singularity is
removed by some flow process. The deformation and fracture behaviour of
the specimen depend upon whether the stressing path crosses the shear yield
envelope before reaching the crazing point, or vice versa.
As noted earlier, craze initiation requires not only a minimum stress,
but also a minimum craze-opening displacement, typically of order 10 nm
(Kramer, 1984). This is easily achieved at crack tips or on specimen surfaces,
but may impose limits on the effectiveness of small particles in initiating
crazes. When the craze opening displacement is larger than about 1% of the
rubber particle radius, quite large tensile strains are necessary to initiate a
craze from a cavitated particle. Consequently, cavitation of small rubber
particles (say, 200 nm in diameter) in the elastic region of the diagram does
not usually lead to crazing. Argon, Cohen and Gebizlioglu (1985) and
Argon and Cohen (1990) used a similar 'displacement misfit' argument to
explain why small rubber particles are ineffective in toughening HIPS, but
without invoking rubber particle cavitation. In the absence of crazing, the
matrix is able to continue to deform elastically, with the stress rising, until
the stress reaches the lower yield boundary. This type of response, involving
390 Rubber toughening

~~---------------------------------,
.'

..••....... tnaxIaI

Mean Stress (MPa)

Figure 8.9 Schematic cavitation diagram showing effects of particle size on cavita-
tion.

cavitation in the elastic region, has been reported in RT-PMMA by Schirrer,


Fond and Lobbrecht (1996).
With this introduction, it is now possible to map the components of
Figure 8.8 onto a single diagram, and examine how they might interact. A
schematic map combining shear yielding and cavitation criterion is shown
in Figure 8.9 for a glassy polymer with 36 vol% of rubber particles, at three
different particle sizes. The largest particles, with diameters of 111m, cavitate
within the elastic region: unless their morphology is optimized for the
formation of strain-hardening rubber elements, they may initiate crazes at
low strains, and fail to provide the expected high level of toughness.
Whether this happens depends upon the resistance of the matrix polymer to
craze initiation from a void. For clarity, no specific values of crazing stress
are marked on Figure 8.9, but crazing may occur above or below the lower
yield boundary, depending upon the properties of the matrix and the
conditions of testing.
Under biaxial stress, smaller particles, with diameters of 0.2 11m, cavitate
between the upper and lower bound yield envelopes, thereby promoting
immediate dilatational yielding. This response is conducive to effective
toughening in plane stress regions of the crack tip, and is close to achieving
toughness in the plane strain region. On the other hand, under uniaxial
tension the 0.21lm particles do not cavitate before the upper bound yield
envelope is reached, and their effect is limited to promoting yielding by
lowering the shear yield stress: there is no dilatational yielding under these
Factors affecting deformation of toughened plastics 391

conditions. The smallest particles, with diameters of 0.1 11m, are unable to
cavitate before craze growth occurs from the crack tip, and failure is brittle.
This pattern of behaviour has been observed in RT-PVC by Wu (1992) and
by Dompas and Groeninckx (1994).
Cavitation diagrams can provide a valuable insight into the competing
effects on deformation behaviour of both external parameters (strain rate,
temperature, notch radius and stress state), and materials structure (particle
size, particle-matrix adhesion, crosslinking in the rubber and matrix com-
position). However, they are designed for only one important but limited
purpose, namely to define the limits of elastic behaviour in the matrix
polymer, and the conditions necessary to initiate shear yielding, crazing,
rubber particle cavitation or brittle fracture. This may provide some under-
standing of the subsequent deformation behaviour of the material, but a
complete analysis of this behaviour involves other considerations that are
beyond the scope of the diagram. Both aspects of toughening are covered in
section 8.5 below.

8.5 FACTORS AFFECTING DEFORMATION OF TOUGHENED


PLASTICS
The literature contains a large amount of experimental data on the defor-
mation behaviour of toughened plastics over a range of test conditions,
specimen geometries and material morphologies. As a result of this work,
some general principles have been established, but there are still topics that
are poorly understood, or controversial (e.g. interparticle spacing). The
understanding that has resulted from recognizing cavitation as a controlling
step in rubber toughening is helping to resolve some of the difficulties and
anomalies encountered in this complex field. The principal factors affecting
fracture resistance are each now reviewed in turn.

8.5.1 STRAIN RATE AND TEMPERATURE


Shear yielding and crazing are both rate processes which follow Eyring
kinetics (equation 8.8). Consequently, increases in strain rate and decreases
in test temperature have similar effects, in raising both the yield stress and
the craze growth stress. By contrast, cavitation strains are generally insen-
sitive to changes in strain rate: the exceptions occur near the a-transition in
the rubber, where Gr increases strongly with decreasing time scales, thereby
raising U vCr) in equation 8.3 (section 8.5.9). Cavitation stresses are deter-
mined not only by cavitation strains but also by the moduli of the glassy
polymer, which are themselves time dependent. This means that cavitation
392 Rubber toughening

Mean Stress (MPa)

Figure 8.10 Schematic cavitation diagram showing effects of strain rate.

stresses are somewhat more sensitive than cavitation strains to changes in


strain rate.
The effects of temperature on cavitation are a little more complicated. In
addition to its influence on the moduli of both rubber particles and glassy
matrix, temperature has a direct effect on the critical strain at cavitation,
because it determines the level of thermal contraction strain generated in the
rubber particles.
Figure 8.10 is a schematic cavitation diagram showing the effects of strain
rate on deformation mechanisms at a temperature well above the rubber ~ ,
where critical cavitation stresses can be treated as being approximately
constant. At the lower strain rate, 10 - 5 S -1, yielding occurs in uniaxial
tension at stresses well below those at which the particles can cavitate,
whereas in the crack tip region cavitation takes place above the lower yield
boundary, and is therefore followed immediately by dilatational shear
yielding. At the much higher strain rate of 10 s - 1, the particles cavitate
within the elastic region even in uniaxial tension, and the probability of
brittle fracture is increased. Point A on the diagram identifies stress
conditions under which the strain rate increases by a factor of 1 million
when all the particles cavitate. Conversely, the line to the left of point A
indicates that some shear deformation is to be expected in the 'elastic'
Factors affecting deformation of toughened plastics 393

region, albeit at very low rates, before the material has either cavitated or
reached the lower bound yield envelope for a strain rate of 10s- 1 .
The pattern of behaviour illustrated in Figure 8.10 has been observed by
Frank and Lehmann (1986) and by Schirrer, Fond and Lobbrecht (1996) in
transparent RT-PMMA under uniaxial tension. At very low strain rates and
high temperatures there is no cavitation: the material yields without either
loss of transparency or increase in volume. At intermediate strain rates and
temperatures, light scattering shows particle cavitation on inclined planes,
but the volume remains constant within experimental error (after allowing
for elastic strains). Finally, at high strain rates or low temperatures, light
scattering due to cavitation on inclined planes is rapidly obscured by
massive whitening accompanied by substantial volume increases. The prin-
cipal mechanism of deformation at room temperature is shear yielding, with
dilatational processes occurring only at high strain rates and low tem-
peratures.

8.5.2 NOTCHES
Fracture problems become most severe when the specimen or component is
relatively thick and contains sharp notches or cracks. The lines labelled
'triaxial' in Figures 8.4, 8.8c, 8.9 and 8.10 indicate the stress state in the plane
of the crack under plane strain loading, where the ratio ae/am is at a
minimum and the extent of shear yielding is therefore severely restricted.
The main purpose of adding rubber particles is to overcome this restriction.
Cavitation diagrams illustrate the way in which this may be achieved,
through multiple crazing or the formation of dilatational shear bands. At
the outer edges of the crack, the material is in plane stress, and under a state
of equi-biaxial stress (marked 'biaxial' in the diagrams) in the crack plane.
Here ae/a m is much higher than in plane strain, and shear yield zones are
relatively large, even in the absence of cavitation. A strain energy density
criterion for crack tips, which combines distortional and dilatational com-
ponents into a single expression, has been developed by Sih (1973) and
Gdoutos (1990). This approach appears to have considerable potential, but
has so far found little application in the polymer field.
Contrary to what has been suggested by some authors, rubber particle
cavitation does not necessarily convert the stress conditions at the crack tip
from plane strain to plane stress. Thickness effects associated with the
presence of plane strain zones in the polymer can persist even after yielding
has begun. This is clearly demonstrated in fracture mechanics studies on
HIPS by Yap, Mai and Cotterell (1983) and by Nikpur and Williams (1979).
There is also some evidence for a thickness effect in ABS (Newmann and
Williams, 1980). A possible reason is that lateral craze propagation is
394 Rubber toughening

restricted when the stress ratio ae/am is low, because the elastic strain
normal to the craze plane is reduced. Electron microscope evidence for
restricted craze extension near crack tips in HIPS under plane strain loading
has been published by Keskkula, Schwarz and Paul (1986).
A model for energy absorption at crack tips in toughened thermosets has
been proposed by Kinloch (1989), and shown to fit experimental data. The
main feature of the model is a term, which increases when the rubber
cavitates, in <Pr' the volume fraction of rubber. The fracture energy is also
taken to be proportional to the plastic zone size, the shear yield stress and
the maximum shear strain of the neat resin. The basic mechanisms of
toughening in RT-epoxy resins are particle cavitation and matrix shear
yielding (Kinloch, Shaw and Hunston, 1983). In highly crosslinked resins,
where shear yielding is limited, the alternative mechanisms of crack bifur-
cation, crack deflection and possibly particle bridging become more impor-
tant (Sue, 1991).

8.5.3 RUBBER CONTENT


Rubber content is an important factor in the deformation and fracture of all
rubber-toughened plastics. Low modulus rubber particles shed load onto
the stiffer matrix, thereby setting up stress concentrations and reducing the
yield stress. This reduction involves two separate effects: (1) the average
stress on the matrix increases with rubber particle volume fraction, and (2)
local elastic stress concentrations are set up around the particle even at very
low volume fractions, as discussed in section 8.3.1.
Average stresses in the matrix remain high after yielding has begun,
although they may decrease as a result of orientation hardening in the
rubber and the formation of low modulus crazes. Consequently, they affect
not only the initiation but also the propagation of deformation zones. Local
variations in stress concentration, on the other hand, tend to be reduced or
eliminated by the formation of shear bands or crazes, and are therefore more
important at the initiation stage than during subsequent propagation and
growth of deformation bands.
The importance of average as opposed to local stress concentrations in
the matrix is seen in measurements of yield stress in uniaxial compression,
where there is no contribution from crazing or particle cavitation (Kinloch,
Shaw and Hunston, 1983). Gloagen et al. (1993) found that the compressive
yield stress aye of RT-PMMA decreased almost linearly with rubber particle
concentration <Pp as follows:

(8.11)
Factors affecting deformation of toughened plastics 395

60

"'
Q.
~
en
en
50

(IJ 40
.to
(/')
(IJ
> 30
13
&
W 20
en
(IJ
en
10
~
c:
0
> 0

Mean Stress (MPa)

Figure 8.11 Schematic cavitation diagram showing effects of rubber content.

Tensile tests on HIPS and ABS show a more marked dependence of yield
stress (J y on the rubber particle volume fraction than that shown in equation
8.11, but again indicate that average stresses on the matrix, rather than local
stress concentrations, largely control the yield behaviour of toughened
plastics. Ricco et al. (1985) found that the tensile yield stress of ABS followed
the Ishai- Cohen (1968) effective area equation quite closely. while Bucknall,
Davies and Partridge (1986) found that the Ishai- Cohen equation sig-
nificantly overestimated tensile yield stresses in HIPS, but matched the
overall trend.
The effects of rubber content on the cavitation diagram are illustrated in
Figure 8.11. The upper bound shear yield envelope, which can be defined
using data from compression tests, simply shifts downwards with c/>p, but the
lower bound envelope is a function of feff' and hence of c/>p. It therefore
changes not only its location but also its shape, diverging more strongly
from the upper bound with increasing c/>p. A comparison between Figures
8.10 and 8.11 emphasizes the point that changes in yield behaviour due to
increases in strain rate or reductions in temperature can to a large degree
be offset by increasing the rubber content. It is frequently observed that the
critical rubber content required to obtain a given level of toughness
increases with increasing strain rate.
Rubber content Cr and rubber particle volume fraction c/>r both have a
strong effect upon the kinetics of deformation in toughened plastics. The
interpretation of these effects raises problems similar to those encountered
396 Rubber toughening

in chemical kinetics. In some cases, a single rate-controlling step operates


over a range of compositions, rates and temperatures, and interpretation is
straightforward. In others, more complicated kinetic schemes are necessary.
The kinetics of tensile creep in RT-PMMA at 23°e appear to be relatively
straightforward (Bucknall, Partridge and Ward, 1984). Dilatation is neg-
ligible and, as noted in section 8.3.3, creep strain increases approximately
linearly with (Bt)1/3. Plots of In B against applied stress (J for each material,
including PMMA itself, give straight lines in accordance with the Eyring
kinetics, equation 8.8. The stress concentration factor, calculated from the
slope (a In B/o(J), increases gradually with rubber content, from y = 1 at
<Pr = 0 to Y= 2.1 at <Pr = 0.36. A master curve correlating data from all
of these tests can be obtained by plotting In B against Y(J. In other words,
a single composition-dependent average stress concentration parameter
defines both the stress and composition dependence of creep behaviour.
This is not true of creep in HIPS (Bucknall, Davies and Partridge, 1986),
which also shows a linear dependence of log S upon applied stress. Eyring
curves for a series of HIPS- PS blends cannot be superimposed using a
simple stress concentration factor. The dependence of the slope (olns/o(J)
upon <Pr is stronger than would be expected from the shift in the position of
the curves along the (J-axis. A possible explanation is that creep kinetics in
HIPS are determined not only by rates of craze thickening but also by the
extent of rubber particle cavitation, which controls craze initiation. Al-
though craze thickening and growth would be expected to follow Eyring
kinetics, with relatively low average stress concentration factors as in
RT-PMMA, these processes depend upon craze initiation through rubber
particle cavitation, which is not an activated process, but is determined
simply by elastic volume strains and particle size distributions. Overall creep
rates reflect the combined effects of initiation criteria and propagation
kinetics.

8.5.4 PARTICLE-MATRIX ADHESION


Because cavitation of the rubber particles is recognized as playing an
important part in rubber toughening of plastics, some authors have argued
that rubber particles could equally well be replaced with small holes, which
might be expected to behave in a similar manner to cavitated particles, in
that they accelerate yielding and relieve constraints at crack tips.
Under certain conditions, real voids do indeed confer a degree of
toughness upon glassy polymers, basically by inducing a plane strain to
plane stress transition. Using double cantilever beam fracture mechanics
tests, Hobbs (1977) has shown that polycarbonate structural foam, con-
taining about 20 vol % voids, has a substantially higher fracture toughness
Factors affecting deformation of toughened plastics 397

than the solid polymer. Crazing occurs in the walls separating the voids,
but at a stress that is sufficiently high to promote significant shear yielding
under the less constrained crack tip conditions in a foam. Polycarbonate is
exceptional in this respect: voids weaken and embrittle most glassy poly-
mers, whereas weakly bonded rubber particles may prove very effective as
toughening agents.
Using cavitation diagrams, it is easy to see why poorly bonded rubber
particles are more effective toughening agents than voids. Because particle-
filled materials must reach a critical strain before debonding can occur,
relatively low levels of particle-matrix adhesion are sufficient to ensure a
high volume strain for debonding, and hence a satisfactory degree of
toughness, when particle sizes are small.
This can be shown using a simple energy balance calculation. Defining
failure as debonding over a complete hemisphere, to form two new surfaces
of total area 4nR2 and specific surface energy fs' and assuming for simplicity
that all energy released during debonding comes from the rubber particle,
particle, then the energy balance principle predicts debonding when:

inR3 KrA~ > 4nR2fs

or

6fs )1/2
A > ( __ (8.12)
v Kr R

Taking Kr = 2 GPa, R = lOOnm and assigning a mlmmum value of


fs = 0.035 J m - 2, equation 8.12 gives a critical volume strain of 3.2%, well
above the strain required to cavitate a particle of this size. For the data
given, this figure is an overestimate, because the calculation does not allow
for energy released from the surrounding matrix when debonding occurs,
which may be two or three times the energy stored in the particle. However,
it does show that small particles of low shear modulus are likely to cavitate
internally before they debond, even when the interface is very weak.
Interfacial failure may take place subsequently, after the surrounding
material has yielded and the particle has extended and dilated, because the
stored energy in the particle is then much higher than it was when it
cavitated. However, the important question is not whether the particles
debond or cavitate, but whether both processes can be delayed until the
material is beyond the lower bound yielding envelope, so that they produce
dilatational shear yielding rather than crazing and fracture.
The relationship between particle size and debonding can be seen in a
scanning electron micrograph due to Kinloch (1985), showing the fracture
surface of a brittle epoxy resin containing dispersed particles of a rubber
398 Rubber toughening

which was chosen deliberately to give poor adhesion to the epoxy matrix.
Particles over about 10 J..lm in diameter debonded during fracture, whereas
all smaller particles cavitated internally. A similar pattern of behaviour has
been reported by Pearson and Yee (1991) in RT-epoxy resin containing a
bimodal distribution of rubber particle sizes. These observations suggest a
possible method for measuring interfacial energy fs.
The figure of fs = 0.035 J m - 2 is based on the surface energy of poly-
styrene (Kramer, 1984) assuming no interdiffusion at the rubber-matrix
interface, and is far lower than measured G1C values, which include energy
dissipated by viscoelastic relaxation around the crack tip. In practice, the
microscopic surface energy of debonding between rubber particles and
typical glassy polymers is unlikely to be as low as 0.035 J m - 2. Adhesion
tests by Brown (1990) on PS-PMMA interfaces gave G1C = 13 J m - 2, which
is much higher than the fracture energies obtained by Robertson (1974) and
Kusy and Turner (1974) for very low molecular weight grades of PS and
PMMA.
Studies by Van der Sanden, Meijer and Lemstra (1993) have shown that
weakly bonded rubber particles can produce ductile yielding even in
polystyrene, provided that they are very small and the rubber content is high
enough. They blended PS with 200 nm diameter MBS particles, which have
rubbery poly(butadiene-styrene) cores enclosed in grafted PMMA shells,
and found that the blends were tough, even (indeed especially) at 60 wt%
MBS. The authors concluded that the toughness was due to extensive
debonding, and suggested that PS ligaments formed between debonded
particles were below a critical dimension for crazing.
This view is difficult to support in the light of their experimental evidence,
which shows yielding of 60% MBS blends at a stress of 22 MPa, followed
by shear deformation at constant volume up to about 200% extension. The
absence of a volume change clearly indicates that there is little or no
debonding, despite the suggestion that the particle-matrix adhesion is low.
Almost quantitative dilatation occurred during drawing of PSjMBS blends
with lower rubber contents and therefore higher yield stresses. Since the
maximum packing fraction of identical spheres is 0.74, it appears more likely
that at 60% concentration the PMMA-coated rubber spheres formed
continuous regions of PMMA within the glassy matrix, where shear yielding
took place preferentially. When the PMMA coating was replaced with a PS
coating, which is better bonded but more craze-prone, the PS blends were
still tough but their tensile elongations were reduced to 30%. Recent work
by Schneider (1995) has shown that polystyrene can be toughened very
effectively by adding relatively large (> 1 J..lm) particles made from natural
rubber latex, with grafted PMMA shells. No evidence of debonding was seen
in the TEM.
Factors affecting deformation of toughened plastics 399

Reasonably high levels of particle-matrix adhesion are essential in HIPS,


ABS and a number of other toughened glassy polymers which are prone to
crazing, and yield by multiple craze formation under standard test con-
ditions. Extensive failure of the rubber fibrils, either by rupture or by
debonding from the matrix, leads to immediate fracture of the material. The
required high levels of adhesion are achieved using graft and block
copolymers. In many cases, chains of the glassy polymer are grafted onto
the rubber particles during manufacture. This not only affects the strength
of the interface, but also alters the spatial distribution of the particles: the
grafted surface layer reduces the tendency to agglomerate and provides a
barrier to contact between rubber domains (Rink et ai., 1978).

8.5.5 RUBBER PARTICLE SIZE


As discussed earlier in this chapter, particle size can affect the deformation
behaviour of rubber-modified plastics in three ways, through its effects on
(1) the critical volume strain to cavitate the rubber, (2) the critical tensile
strain to initiate a craze and (3) the volume strain to debond the particles.
Wherever possible, steps are taken by the manufacturer to ensure that
rubber-matrix adhesion is good, and the choice of particle size then
depends upon the relative positions of the lower bound shear yielding
envelope and the critical crazing stress.
In toughening glassy polymers such as PC, which have high entanglement
densities and small molecular cross-sections, and which are therefore rela-
tively resistant to crazing, there are advantages in using small particles. This
ensures that both cavitation and debonding are postponed until they can
initiate immediate dilatational shear yielding, which means reaching a strain
above the lower bound yield envelope. The importance of delayed void
formation in PC has been demonstrated by Sue, Huang and Vee (1992) in
fracture mechanics tests on PCjLDPE blends. Adding 3.3 wt% of poorly
bonded LDPE to PC raised its J IC value from 2.5 to 5.7kJm- 2 , whereas a
similar concentration of MBS core-shell rubber (Parker et aI., 1990) gave
J IC = 4.6kJm- 2 • From electron microscope evidence, Sue, Huang and Vee
(1992) concluded that LDPE particles are the more effective toughening
agents because they de bond at a higher strain than that required to cavitate
MBS particles. In both blends it appears that the modifier particles play
little part in the deformation after forming voids: orientation hardening in
the PC matrix is more than sufficient to compensate for the loss of
load-bearing particles.
In less ductile polymers, such as PS, it has to be accepted that crazes will
almost certainly initiate from the specimen surface before the material
reaches the lower bound shear yield envelope, so that an alternative strategy
400 Rubber toughening

is required. If crazing cannot be avoided, then the best approach is to choose


relatively well bonded, large complex particles, which will not only promote
multiple crazing but also form bridging rubber fibrils that can provide
mechanical stability to the crazes. In PS, small rubber particles are ineffec-
tive craze initiators, and larger (> 1 J.lm) particles give better toughness
(Hobbs, 1986; Argon and Cohen, 1990; Cigna et al., 1992; Cook, Rudin and
Plumtree, 1993).
The two other commercially important glassy thermoplastics, PMMA
and PSAN, are intermediate between polycarbonate and polystyrene. Over
a range of temperatures and strain rates, both RT-PMMA and ABS deform
by a combination of shear yielding and crazing, with ABS showing the
greater tendency to craze at and below room temperature, especially at high
strain rates. Because of this dual deformation mechanism, manufacturers of
ABS often choose a bimodal distribution of particle sizes, with a prepon-
derance of small particles to promote shear yielding, but a fraction of larger
particles or particle aggregates (say 0.5 J.lm diameter), which can act as craze
initiators under critical conditions.
Particle size effects in PVC/MBS blends were studied by Dompas and
Groeninckx (1994), using tensile tests (Dompas et al., 1994a), notched Izod
impact tests (Dompas et al., 1994b), light transmission and TEM. As
predicted by equation 8.3, the critical volume strain at cavitation fell with
increasing particle size, over the range 80-300 nm. Electron microscopy
confirmed that the larger particles cavitated preferentially, and impact tests
showed that very small particles (D = 92 nm) are ineffective as toughening
agents. These effects are explained in section 8.4.
Figure 8.12 illustrates this effect in blends of rigid PVC with 10% MBS
modifier (Purcell, 1972). Two sets of notched Izod data are presented, for
specimens of differing thicknesses. For the 3.2 mm bars it is clear that plane
stress conditions are established at the crack tip, so that plastic deformation
can occur without cavitation, and impact strength is almost independent of
particle size. In the 6.4 mm bars, by contrast, the depression in yield stress
produced by small solid rubber particles is not sufficient to prevent plane
strain conditions from developing at the crack tip, and toughness is
optimized only when the particles are large enough to cavitate before crack
growth begins, causing a further reduction in yield stress. In these thicker
bars the fraction of the crack tip exhibiting plane stress fracture increases
with particle size, and the impact strength rises accordingly.
The literature on particle size effects in rubber-toughened thermosets,
notably epoxy resins, contains a number of contradictory results (Kinloch,
1989; Pearson and Vee, 1991). It is generally agreed that very large particles
(> 20 J.lm) are ineffective as toughening agents, but different trends have been
reported for medium sized particles (0.5-5.0 J.lm). Some authors find little
Factors affecting deformation of toughened plastics 401

15
~

.r::.
.B
0
C
I

E
.!2
2-
.r::. 10
C>
c
Q)
'-
U5
u(II
0-
E
u 6mm thick
0 5
~
u
Q)
.r::.
u
0
Z

O+-------.-------r------.--~
0.0 0.1 0.2 0.3
Rubber Particle Diameter (Ilm)

Figure 8.12 Effects of specimen thickness on the notched Izod impact strength of
toughened PVC. (Data of Purcell, 1972.)

dependence of toughness upon particle size, while others report significant


variations. The main reason for this lack of consensus appears to be that
factors other than particle size are also changed when the rubber particles
are formed by phase separation from the liquid resin. To overcome this
problem, Kim et al. (1996) used core-shell emulsion polymers to toughen
two grades of epoxy resins: a lightly crosslinked material cured with
piperidine, and a more highly crosslinked resin cured with diaminodiphenyl
methane (DDM). The rubbery cores varied in diameter from 0.1 to 1.0 !lm.
In the piperidine-cured resin, high levels of fracture toughness were
achieved, which increased by 30% on decreasing the particle size from 1.0
to 0.34 !lm, but fell drastically on reducing the rubbery core diameter to
0.1 !lm, at which size the particles debonded with little evidence of internal
cavitation. The degree of toughening achieved in the DDM-cured resin was
much lower, and showed little dependence upon particle size.
The effects of particle size upon toughness are sometimes presented in
terms of another parameter, the interparticle spacing, which is also a
function of the particle volume fraction ¢p. There is little convincing
402 Rubber toughening

evidence for toughness transitions in toughened glassy polymers that are


related to interparticle spacing as distinct from the two primary variables.
It is important tei note that such evidence would have to show that the
critical size was a function of c/J p •

8.5.6 PARTICLE MORPHOLOGY


Electron microscopy reveals a wide range of rubber particle morphologies
in toughened plastics, from simple homogeneous spheres to complex com-
binations of rubber and rigid inclusions, as illustrated in Figure 8.1 (Riess,
Schlienger and Marti, 1980; Sardelis, Michels and Allen, 1987; Argon et al.,
1990). These inclusions affect mechanical properties in three ways: they
increase the volume fraction of particles, and hence decrease the volume
fraction of load bearing matrix; they confine and restrict the growth of voids
within the rubber phase; and they provide anchor points for the rubber
fibrils.
Complex particle morphologies are most frequently seen in HIPS. Large
homogeneous rubber particles simply weaken the PS matrix when they
cavitate, effectively acting as spherical voids (Donald and Kramer, 1982d).
Once cavitation has begun, the polystyrene ligaments between particles
first craze and then fracture, with the rubber making little contribution
to strength. In order to overcome this problem, it is necessary to introduce
one or more inclusions (usually of polystyrene) into the rubber particles.
With this morphology, cavitation of the particles results in fibrillation of
the rubber membranes and a reinforcement of the surrounding crazes. The
rubber fibrils act as strain hardening elements in the deforming structure.
The presence of rigid inclusions necessarily raises the stiffness of the
rubber particles. However, the degree of stiffening depends upon the exact
morphology of the particle. In experiments on HIPS, Bucknall, Cote and
Partridge (1986) found that modulus and yield stress were determined
primarily by the volume fraction of particles, and that the ratio of rubber to
inclusions was only of secondary importance. By contrast, in RT-PMMA,
Archer et al. (1993) found that E and (Jy were dependent primarily on the
concentration of rubber phase, and not on total particle volume fraction.
The reasons for these differences are not properly understood.
In the absence of rigid subinclusions, growing voids tend to be spherical,
thus minimizing their surface energy. However, this is impossible in particles
containing high concentrations of glassy polymer, where voids are con-
strained to form in the thin membranes of rubber separating the inclusions.
Only when they are extremely small are these voids even approximately
spherical. Consequently, the energy required to form a given volume of
voids is much greater than that indicated in equation 8.3. To supply this
Factors affecting deformation of toughened plastics 403

extra energy, it is necessary either to increase the volume strain in the


particles and adjacent matrix, or to obtain energy from a larger volume of
material by making the particles bigger.
After cavitation, rubber particles are able to dilate, showing some of the
characteristics of true voids. The effective void content ierr, and therefore the
deformation behaviour of the material, is determined both by the fraction
of particles cavitated and by the properties of the cavitated particle. Large
homogeneous rubber particles show little resistance to continued dilatation;
but particles that are filled with rigid inclusions connected by rubber fibrils
behave quite differently, allowing deformation to initiate relatively easily in
the matrix, then slowing the process down locally by becoming stiffer
through orientational hardening of the fibrils. Starke et al. (1997) have
shown that the stresses within three-layer core-shell acrylic rubber particles
embedded in a PSAN matrix can even become high enough to disrupt the
PMMA cores.

8.5.7 CROSSLINKING OF RUBBER


Crosslinking raises the shear modulus of rubbers, making them more
difficult to cavitate, as illustrated in Figure 8.3. In the more ductile polymers,
such as PC, this may not present problems. It may even prove advan-
tageous, provided that the level of crosslinking is not excessive, because
crosslinking has a similar effect to a reduction in particle size, in that it shifts
cavitation of the rubber to higher stresses where dilatational yielding can
take place immediately. If carried too far, however, crosslin king will delay
cavitation so much that the material fails by crazing and fracture from
external surfaces before significant yielding has taken place.
Once cavitated, homogeneous rubber particles exercise only a limited
influence on the subsequent yielding and fracture of the glassy polymer,
because they are unable to form effective orientation-hardened bridges
across shear bands, crazes and cracks (section 8.5.5). Fibrillated rubber
membranes, stabilized by glassy subinclusions, are much better in this role.
Here again, crosslinking has a strong effect upon the behaviour of the
toughened polymer, because it determines the distribution of stress between
the rubber particles and the matrix, and the strain at break in the rubber
fibrils.
The effects of crosslinking in the rubber phase are seen most clearly in
HIPS and ABS, where mUltiple crazing is the main mechanism of toughen-
ing. Crosslinking of the particles raises tensile yield stresses and depresses
impact strength (Birkinshaw, Buggy and O'Niell, 1990; Birkenshaw, Buggy
and Quigley, 1993; Soares, 1994; Bucknall et aI., 1996). In order to optimize
energy absorption in these materials, the rubber should ideally have a low
404 Rubber toughening

level of crosslinking, allowing early cavitation of the particles and maximum


extension of the rubber fibrils, but it should also form a graft or block
copolymer with the glassy polymer to provide anchor points for the fibrils.
Experiments by Keskkula and Turley (1978) have shown that effective
toughening can be achieved in PS using polyisoprene that has been grafted
but not crosslinked. In this experiment the PS subinclusions of the salami
particles acted as 'physical crosslinks' in a manner similar to the glassy
phase in SBS thermoplastic elastomers. However, toughness was lost during
later melt processing, which broke the particles down into small fragments.
Excessive crosslinking of the rubber, introduced either deliberately during
manufacture or through subsequent oxidation, can lead to a marked
reduction in impact strength. A particular problem is photo-oxidation on
exposure to ultraviolet light. Degradation of the surface layer, typically to a
depth of about 0.1 mm, may result in a severe embrittlement of the material
(Bucknall, 1977). Significant levels of crosslinking are also produced simply
by prolonged heating at high temperatures. In extreme cases, the ~ of the
rubber is shifted to much higher temperatures. However, studies on both
irradiated and vulcanized HIPS (Soares, 1994; Perche, 1995) have shown
that tensile elongation and impact energy absorption are adversely affected
even when the upward shift in the rubber ~ is only a few degrees. The
rubber remains rubbery throughout these crosslinking reactions, and the
changes in toughness are due solely to its higher shear modulus.

8.5.8 MATRIX CHARACTERISTICS


The nature of the matrix is of critical importance in determining both the
maximum achievable fracture resistance and the optimum morphology
required to achieve that resistance. Generally, the matrix absorbs most of
the fracture energy, and the role of the rubber phase is the secondary one of
extending that energy absorbing capacity as far as possible. All glassy
thermoplastics are capable of reaching large extensions under appropriate
conditions, but they also have the ability to fracture at quite low strains,
with the deformation localized into one or two crazes. Thermosetting resins
are also capable of large deformations if they are lightly crosslinked, but
their ductility becomes increasingly restricted as the crosslink density
increases.
Earlier sections of this chapter have shown how cavitation diagrams
provide a basis for analysing the mechanical response of rubber-toughened
thermoplastics, and the effects of both test conditions and polymer structure
on shear yielding and crazing. To a large extent, the toughness of a
rubber-modified polymer depends upon the resistance of its matrix to
crazing, which is determined both by its entanglement density and by the
Factors affecting deformation of toughened plastics 405

strength of its chains (Vincent, 1972; Kramer, 1984). Factors that raise the
shear yield stress, including physical aging, also tend to promote crazing.
Of the polymers in common use, polystyrene is the most susceptible to
crazing. Crazing resistance in styrenic polymers is increased by copolymeriz-
ing with acrylonitrile (Kim, Keskkula and Paul, 1991), and reduced by
copolymerization with a-methyl styrene or maleic anhydride. Another route
to a craze-resistant styrenic polymer is to blend the PS with PPE. The
higher craze resistance of a miscible PS- PPE matrix is reflected in the
transition from multiple crazing to shear yielding as the ratio of PPE to
HIPS increases (Bucknall, Clayton and Keast, 1972). Toughening of this
blend requires smaller rubber particles than those used in standard HIPS.
Craze fibrils are stabilized by entanglements, and the ratio le/1 (where Ie
is the chain length between entanglements and 1 is the total contour length
of the polymer chain) is therefore of fundamental importance in determining
the fracture resistance of glassy polymers. Below a critical value of 1/le' the
fracture energy G,c falls rapidly with the molecular weight of the glassy
polymer. Under certain conditions, rubber-toughened polymers show a
similar dependence of toughness upon the molecular weight of the matrix.
The effects of chain length on the fracture of glassy polymers are seen
most clearly in fatigue. Bucknall and Faitrouni (1991) studied the fatigue
crack growth rate per cycle, da/dn, as a function of the stress intensity
amplitude 11K in three grades of PSAN of differing molecular weight, and
in three ABS polymers which were prepared by blending each PSAN with
lS% of grafted rubber particles. In standard plots of log(da/dn) against
10g(I1K), they obtained three roughly parallel 'Paris-law' lines of slope '" 4
for PSAN, showing a shift to higher 11K values with increasing molecular
weight. There was a clear separation between the three K,c values at high
crack growth rates, and also between the three threshold values of stress
intensity amplitude I1K thn which corresponded to fracture energy ampli-
tudes I1G = 30,70 and 100Jm- 2 respectively at da/dn = 10nm cycle-I.
The ABS polymers showed somewhat different behaviour. At low da/dn,
below 10 nm cycle - \ there was again a separation between values of I1K thn
corresponding to fracture energy amplitudes I1G r = 1S0, 300 and SOOJm- 2 .
However, with increasing 11K and da/dn, the three ABS curves converged
to a single point, at Kmax = 3.S MPa m -1/2 (equivalent to Gc = 6000J m -2).
These results indicate that fatigue failure in ABS occurs through craze
extension and failure from the crack tip at low 11K and very low da/dn, and
that fatigue cracks meet increasing resistance from dilatational shear yield-
ing at the crack tip as 11K increases. The rubber particles have a measurable
effect on fatigue behaviour in the threshold region, where they act as
stabilizing bridges across the growing crazes, but this effect is very local. At
high 11K, cavitated particles can initiate yielding over a deformation zone
406 Rubber toughening

extending 1 mm or more from the crack tip, so that chain disentanglement


becomes much less important, and the molecular weight dependence disap-
pears.
The fracture behaviour of thermosets is determined largely by the degree
of crosslinking in the resin phase. Plane strain compression tests show that
strains to failure in epoxy resins decrease with increasing crosslink density
(Finch, Hashemi and Kinloch, 1987). This loss of ductility is reflected in the
behaviour of the corresponding rubber-toughened thermosets. Pearson and
Yee (1989) studied a series of DGEBA (diglycidyl ether of bisphenol A)
epoxy resins cured with 4,4'-diamino-diphenyl-sulphone, and found that G1C
of the neat resin increased with Me' the molecular weight between cross-
links. On adding 10% of CTBN rubber, the fractional increase in G1C
increased rapidly with Me' Similar results were obtained by Levita (1989) in
the same system. Clearly, the effectiveness of rubber toughening is severely
reduced when the resin is so highly crosslinked that it has substantially lost
its capacity to yield.
High levels of crosslinking are necessary in thermosets in order to achieve
stiffness at elevated temperatures, a basic requirement in a number of
applications, notably high performance composites for aerospace. For
these applications, toughness is achieved by blending the strong, brittle,
thermosetting resin with a ductile, high ~ thermoplastic such as poly-
etherimide (Bucknall and Gilbert, 1989) or polysulphone (Bucknall and
Partridge, 1983; Carter, 1991), to form a co-continuous morphology. The
thermoplastic provides energy absorption and raises G1C ' without the
reduction in stiffness that always results from adding a rubber.

8.5.9 RELAXATION BEHAVIOUR OF RUBBER


To provide effective toughening, the rubber phase must have a low shear
modulus Gr , well below that of the glassy matrix Gm . Application of the
Goodier (1933) equations shows that stress concentrations in the matrix
increase, and shear yield stresses therefore decrease, as Gr/Gm is reduced to
0.1, but that in the absence of cavitation further reductions in Gr have
relatively little effect on the onset of yielding. This sets an upper limit of
'" 100 MPa on Gr. However, where cavitation of the rubber particles plays
an important part in the toughening mechanism, Gr must be much lower,
preferably below 0.5 MPa (Figure 8.3). These low moduli can be achieved
only when the rubber is fully relaxed.
Below its a-relaxation temperature, the shear modulus of a typical rubber
is about 1000 MPa, whereas at higher temperatures, Gr values of 0.1-1.0
MPa are more typical. As discussed in section 8.2, the a-transition in the
rubber phase is seen as a low temperature secondary peak in the dynamic
Factors affecting deformation of toughened plastics 407

mechanical loss curve, which at standard test frequencies v '" 1 Hz is


close to the ~ of the rubber. At higher frequencies and shorter time scales
the transition moves to higher temperatures, a phenomenon that is most
easily studied using dielectric methods. For example, Pratt and Smith (1992)
studied a toughened blend of PC with poly(ethylene terephthalate) (PET),
and observed a shift in the dielectric loss peak due to the impact modifier,
from lODC at v = 1 Hz to 70 DC at v = 10 kHz. They also found that log v
increased linearly with liT, in accordance with the Arrhenius equation,
giving an activation energy of 120 kJ mol- 1.
Shifts of this kind have a profound effect on the fracture behaviour of
toughened plastics, especially in impact, where time scales are short. The
characteristic interval between first contact and crack initiation in a standard
notched Izod or Charpy impact test on plastics is about 1 ms, and during
subsequent fracture cracks can travel at speeds of up to 800 m s - 1, close to
the velocity of the transverse wave (Dear, 1996). At these limiting speeds the
time available for a rubber particle to respond to an applied stress is in the
order of IOns. Consequently, the dynamic fracture toughness of rubber-
modified plastics remains low over a temperature interval extending some
tens of degrees above the ~ of the rubber particles (Bucknall, 1977; Julien,
1995). A recent paper by Leevers (1995) proposes a thermal decohesion
model to explain why the dynamic fracture resistance GD of ductile polymers
is so low at high crack speeds (> 100 m s - 1), which do not allow the normal
crack-shielding processes sufficient time to operate effectively.
At very low temperatures, below the ~ of the rubber, no true rubber
toughening effect can occur. Consequently, notched impact energies gen-
erally fall to the level of the parent polymer (Bucknall, 1977). Nevertheless,
Kinloch and Hunston (1987) have observed increased values of G1C in
epoxy-CTBN (carboxyl-terminated butadiene-acrylonitrile copolymer)
blends at temperatures as much as 30 K below the ~ of the modifier,
apparently contradicting the principle. A possible explanation is that the
epoxy resin is so brittle that glassy CTBN is able to increase energy
absorption by acting as a ductile thermoplastic.
The first signs of toughness are usually observed just above the ~ of the
rubber phase, where in fracture mechanics tests G1C begins to rise, and in
notched impact tests the energy absorption increases. Fracture surfaces
show an increasing area of whitening spreading from the notch. These
changes all occur before the load maximum, where the crack or notch tip is
stationary: subsequent crack propagation is rapid and brittle because GD is
low. In some cases, notably rigid PVC, the onset of toughening is limited
not by the relaxation behaviour of the rubber but by the deformation
behaviour of the matrix: PVC becomes more ductile above its fJ-relaxation,
which is centred at about - 30DC (Bucknall and Street, 1967).
408 Rubber toughening

Full toughness is achieved in rubber-modified plastics at temperatures


well above the ~ of the modifier, where the material has a high dynamic
fracture resistance, as well as a high static toughness. At these higher
temperatures, the fracture surface is fully ductile and usually heavily
whitened.

8.6 OVERVIEW
This chapter has examined the principal factors that affect fracture resis-
tance in rubber-toughened glassy polymers, with particular reference to six
commercially important and well documented glassy polymers: PS, PSAN,
PVC, PC, PMMA and epoxy resin. Since rubber-toughened plastics were
first manufactured in the late 1940s, these materials have been very exten-
sively studied, and there is now a large amount of information on them in
company reports, patents, journals and books. However, despite this large
body of experimental data, of which it has been possible to discuss only a
fraction in this review, interpretation of structure-property relationships
continues to present difficulties.
There are two main reasons for these difficulties. First, fracture itself is a
complex dynamic process, which cannot be characterized satisfactorily in
terms of a single quantity, such as an impact strength. Second, rubber-
toughened plastics respond to stress in a complex way, which varies with
their structure and with test conditions. Consequently, different test methods
rank materials differently. The evaluation of rubber-toughened plastics in
industrial research laboratories has inevitably relied upon simple, short term
tests such as the Izod test, which have become the basis for optimizing
composition and morphology. Much less emphasis has been placed upon
measurements that relate more closely to long term performance, such as
fatigue tests. A better understanding of structure-property relationships,
especially in the area of whole-life prediction, should lead to better matching
of rubber-toughened plastics to service requirements.
On the purely scientific level, rubber particle cavitation has increasingly
been recognized as a mechanism of central importance to rubber toughen-
ing, whereas earlier it was regarded as an interesting phenomenon of no
great significance. The present chapter reflects this change in emphasis, and
develops a theoretical framework for incorporating particle cavitation
quantitatively into the modelling of stress-strain behaviour. This new
approach is supported by experimental evidence. Its main importance is in
offering a set of principles that can be used both to resolve some of the
puzzling observations that have been reported in the past and to predict new
and hitherto unexpected effects.
References 409

REFERENCES
Andrade, E.N. da C. (1910) Proc. R. Soc. London A, 84, 1.
Archer, A.C., Lovell, P.A., McDonald, 1. et a/. (1993) Am. Chem. Soc. PMSE, 70,153.
Argon, A.S. and Cohen, RE. (1990) in Adv. Polym. Sci., 91-92 (ed. H.H. Kausch),
Springer-Verlag, New York, p. 301.
Argon, A.S. and Hannoosh, 1G. (1977) Phil. Mag., 36,1195.
Argon, A.S. and Salama, M.M. (1977) Phil. Mag., 36,1217.
Argon, A.S., Cohen, R.E. and Gebizlioglu, O.S. (1985) in Toughening of Plastics II,
2-4 July, London, Plastics & Rubber Institute.
Argon, A.S., Cohen, R.E., Gebizlioglu, O.S. and Schweier, c.E. (1983) in Adv. Polym.
Sci., 52-53 (ed. H.H. Kausch), Springer-Verlag, Heidelberg, p. 275.
Argon, A.S., Cohen, R.E., Jang, B.Z. and van der Sande, 1.B. (1981) J. Polym. Sci.,
19, 253.
Argon, A.S., Cohen, R.E., Gebizlioglu, O.S. et al. (1990) Macromolecules, 23, 3975.
Bascom, W.O., Cottington, R.L., Jones, R.L. and Peyser, P. (1975) J. Appl. Polym.
Sci., 19, 2545.
Bates, F.L., Berney, C.V. and Cohen, R.E. (1983) Macromolecules, 16,1101.
Beahan, P., Thomas, A. and Bevis, M. (1976) J. Mater. Sci., 11, ]207.
Berg, C.A. (1970) in Inelastic Behaviour of Solids (eds 1.R Rice and M.A. Johnson),
McGraw-Hill, New York.
Birkinshaw, c., Buggy, M. and O'Niell, M. (1990) J. Appl. Polym. Sci., 41, 1913.
Birkinshaw, c., Buggy, M. and Quigley, F. (1993) J. Appl. Polym. Sci., 48, 181.
Boyce, M.E., Argon, A.S. and Parks, D.M. (1987) Polymer, 28, 1681.
Breuer, H., Haaf, F. and Stabenow, J. (1977) 1. Macromol. Sci. B: Phys., 14, 387.
Brown, H.R (1990) 1. Mater. Sci., 25, 2791.
Bubeck, RA, Buckley, OJ., Kramer, EJ. and Brown, H.R (1991) J. Mater. Sci., 26,
6249.
Buckley, DJ. (1993) PhD Thesis, Cornell University.
Bucknall, C.B. (1969) J. Mater. (ASTM), 4 214.
Bucknall, c.B. (1977) Toughened Plastics, Applied Science Publishers, London.
Bucknall, c.B. and Clayton., O. (1972) J. Mater. Sci., 7, 202.
Bucknall, c.B. and Faitrouni, T. (1991) 8th International Conference on Defor-
mation Yield and Fracture of Polymers, Churchill College, Cambridge, Institute
of Materials, London, Paper 30.
Bucknall, C.B. and Gilbert, A.H. (1989) Polymer, 30, 213.
Bucknall, C.B. and Partridge, I.K. (1983) Polymer, 24, 639.
Bucknall, C.B. and Smith, R.R. (1965) Polymer, 6, 437.
Bucknall, C.B. and Street, O.G. (1967) Monograph No. 26, Society of Chemical
Industry, London.
Bucknall, c.B., Clayton, D. and Keast, W.E. (1972) J. Mater. Sci., 7,1443.
Bucknall, C.B., Cote, F.F.P. and Partridge, I.K. (1986) 1. Mater. Sci., 21, 301.
Bucknall, C.B., Davies, P. and Partridge, I.K. (1986) J. Mater. Sci., 21, 307.
Bucknall, c.B., Karpodinis. A. and Zhang, X.c. (1994) J. Mater. Sci., 29, 3377.
Bucknall, c.B., Partridge, I.K. and Ward, M.V. (1984) J. Mater. Sci., 19, 2064.
Bucknall, c.B., Soares, V.L.P., Yang, H.H. and Zhang, X.c. (1996) Makromol. Chem.,
Macromol. Symp., 101, 265.
Carter, J. (1991) Plast. Rubb. Compos. Process. Appl., 16, 157.
Cheng, 1., Hiltner, A., Baer, E. et al. (1995) J. App/. Polym. Sci., 55, 1691.
Cigna, G., Maestrini, c., Castellani, L. and Lomellini, P. (]992) J. App/. Polym. Sci.,
44,505.
Cook, D.G., Rudin, A. and Plumtree, A. (1993) J. Appl. Polym. Sci., 48, 75.
410 Rubber toughening

Correa, C.A. (1993) PhD Thesis, Cranfield University, UK.


Coumans, W.I. and Heikens, D. (1980) Polymer, 21, 957.
Craig, T.O., Quick, R. M. and Jenkins, T.E. (1977) J. Polym. Sci., Polym. Chem. Edn.,
15,433.
Davy, P.I. and Guild, F.I. (1988) Proc. Roy. Soc. London A, 418, 95.
Dear, J.P. (1996) Polym. Eng. Sci., 36, 1210.
Dompas, D. and Groeninckx, G. (1994) Polymer, 35, 4743.
Dompas, D., Groeninckx, G., Isogawa, M. et al. (1994a) Polymer, 35, 4750.
Dompas, D., Groeninckx, G., Isogawa, M. et al. (1994b) Polymer, 35, 4760.
Donald, A.M. and Kramer, E.I. (1982a) J. Polym. Sci.: Polym. Phys., 20, 899.
Donald, A.M. and Kramer, E.I. (1982b) Polymer, 23, 461.
Donald, A.M. and Kramer, E.I. (1982c) J. Appl. Polym. Sci., 27, 3729.
Donald, A.M. and Kramer, E.I. (1982d) J. Mater. Sci., 17, 2351.
Donald, A.M. and Kramer, E.I. (1982e) J. Mater. Sci., 17, 1765.
Donatelli, A.A., Thomas, D.A. and Sperling, L.H. (1973) ACS Polym. Prepr., 14(2),
1080.
Donth, E.J. (1992) Relaxation and Thermodynamics in Polymers: Glass Transition,
Akademie Verlag, Berlin.
Edwards, S.F. and Vilgis, T. (1986) Polymer, 27, 483.
Eyring, H. (1936) J. Chem. Phys., 4, 283.
Ferry, J.D. (1980) Viscoelastic Properties of Polymers, 3rd edn, John Wiley, New
York.
Finch, C.A., Hashemi, S. and Kinloch, A.I. (1987) Polym. Comm., 28, 322.
Frank, O. and Lehmann, J. (1986) Colloid Polym. Sci., 264, 473.
Frank, O. and Lehmann, J. (1988) 20th Europhysics Conference, Lausanne, 26-30
September.
Gdoutos, E.E. (1990) Fracture Mechanics Criteria and Application, Kluwer, Dor-
drecht, Netherlands, Ch. 6.
Gebizlioglu, O.S., Beckham, H.W., Argon, A.S. et al. (1990) Macromolecules, 23,
3975.
Gent, A.N. and Wang, C.W. (1991) J. Mater. Sci., 26, 3392.
Gloagen, J.M., Steer, P., Galliard, P. et al. (1993) Polym. Eng. Sci., 33, 748.
Goodier, J.N. (1933) J. Appl. Mech., 1, 39.
Guild, F.J., Young, R.J. and Lovell, P.A. (1994) J. Mater. Sci. Lett., 13, 10.
Gurson, A.L. (1977a) J. Eng. Mater. Technol., Trans. ASME, 99, 2.
Gurson, A.L. (1977b) [CF4 Fracture 1977, Waterloo, Canada, Vol. 2A (ed. D.M.R.
Taplin), Pergamon, Oxford, p. 357.
Haward, R.N. and Owen, R.I. (1973) J. Mater. Sci., 8, 1130.
Hayashi, T. and Nishi, T. (1990) Benibana Symposium, Yamagata University, 8-11
October.
Hobbs, S.Y. (1977) J. Appl. Phys., 48, 4052.
Hobbs, S.Y. (1986) Polym. Eng. Sci., 26, 74.
Ishai, O. and Cohen, L.I. (1968) J. Compos. Mater., 2, 302.
Jeong, H.Y., Li, X.W., Yee, A.F. and Pan, J. (1994) Mechanics Mater., 19 (1 Dec),29.
Julien, O. (1995) PhD Thesis, University of Paris VI.
Karam, H. and Tien, L. (1985) J. Appl. Polym. Sci., 30, 1969.
Kato, K. (1965) J. Electron Micros., 14, 220.
Kato, K. (1967) Polym. Eng. Sci., 7, 38.
Kelly, A. (1966) Strong Solids, Clarendon Press, Oxford.
Keskkula, H. and Turley, S.G. (1978) Polymer, 19, 797.
Keskkula, H., Schwarz, M. and Paul, D.R. (1986) Polymer, 27, 211.
References 411

Kim, H., Keskkula, H. and Paul, D.R (1991) Polymer, 32, 1447.
Kim, D.S., Cho, K., Kim, J.M. and Park, C.E. (1996) Polym. Eng. Sci., 36, 755.
Kinloch, A.I. (1985) in Polymer Blends and Mixtures (eds D.I Walsh, IS. Higgins
and A. Maconnachie), Martinus Nijhoff, Dordrecht.
Kinloch, A.I. (1989) in Rubber Toughened Plastics (ed. CK. Riew) Adv. Chern. Ser.,
222, Am. Chern. Soc., Washington, DC
Kinloch, A.I and Hunston, D.L. (1987) J. Mater. Sci. Lett., 6, 137.
Kinloch, A.I, Shaw, A.J. and Hunston, D.L. (1983) Polymer, 24, [355.
Kramer, E.J. (1984) Polym. Eng. Sci., 24, 761.
Kusy, RP. and Turner, D.T. (1974) Polymer, 15, 394.
Lazzeri, A. and Bucknall, CB. (1993) J. Mater. Sci., 28, 6799.
Lazzeri, A. and Bucknall, CB. (1995) Polymer, 36, 2895.
Leevers, P.S. (1995) Int. J. Fracture, 73, 109.
Levita, G. (1989) in Rubber Toughened Plastics (ed. CK. Riew), Adv. Chern. Ser., 222,
Am. Chern. Soc., Washington, DC.
Li, JX, Ness, IN. and Leung, W.L. (1996) J. Appl. Polym. Sci., 59, 1733.
Lovell, P.A., McDonald, I, Saunders, D.E.I et al. (1991) Plast. Rubb. Compos.
Process. Appl., 16, 37.
Lovell, P.A., Ryan, A.I., Sherratt, M.N. and Young, R.I. (1993) Am. Chern. Soc.
PMSE.,70, 155.
Lovell, P.A., Ryan, A.I., Sherratt, M.N. and Young, R.I. (1994) 9th International
Conference on Deformation Yield and Fracture of Polymers, Churchill College,
Cambridge, Institute of Materials, London, Paper 3.
Lutz, IT. and Dunkelberger, D.L. (1992) Impact Modifiers for PVC, John Wiley,
New York.
MacKnight, W.I., Stoelting, I and Karasz, F.E. (1971) Adv. Chern. Ser., 99, 29.
McClintock, F.A. (1968) J. Appl. Mech., 90, 362.
Maestrini, c., Merlotti, M., Vighi, M. and Malaguti, E. (1992) J. Mater. Sci., 27,
5994.
Magalhaes, A.M.L. and Borggreve, R.I.M. (1995) Macromolecules, 28,5841.
Matsuo, M. (1969) Polym. Eng. Sci., 9, 206.
Michler, G.H. (1992) Kunststoff-Mikromechanik, Hanser-Verlag, Munich.
Montezinos, D., Wells, B.G. and Burns, J.L. (1985) J. Polym. Sci .. Polym. Lett. Edn.,
23,43.
Morbitzer, L., Kranz, D., Humme, G. (1976) J. Appl. Polym. Sci., 20, 2691.
Morbitzer, L., Humme, G., Ott, K.H. and Zabrocki, K. (1982) Angew. Makromol.
Chern., 108, 123.
Newmann, L.V. and Williams, J.G. (1980) J. Mater. Sci., 15, 773.
Nikpur, M. and Williams, J.G. (1979) J. Mater. Sci., 14, 467.
Nimmer, R.P. (1987) Polym. Eng. Sci., 27, 263.
Okamoto, Y., Miyagi, H., Kakugo, M. and Takahashi, K. (1991) Macromolecules,
24,5639.
Parker, D.S., Sue, H.I., Huang, I and Yee, A.F. (1990) Polymer, 31, 2267.
Pearson, RA. and Yee, A.F. (1989) J. Mater. Sci., 24, 2571.
Pearson, RA. and Yee, A.F. (1991) J. Mater. Sci., 26, 3828.
Perche, N. (1995) MSc Thesis, Cranfield University.
Piorkowska, E., Argon, A.S. and Cohen, R.E. (1993) Polymer, 34, 4435.
Pratt, G.I. and Smith, M.I.A. (1992) Third International Conference on Electrical,
Optical and Acoustic Properties of Polymers, London, 16-18 September, Plas-
tics & Rubber Institute.
Purcell, T.O. (1972) Am. Chern. Soc. Polym. Prepr., 13(1), 699.
412 Rubber toughening

Ricco, T., Rink, M., Caporusso, S. and Pavan, A. (1985) International Conference on
Toughening of Plastics II, 2-4 July, Plastics & Rubber Institute, London, Paper
27.
Riess, G., Schlienger, M. and Marti, S. (1980) J. Macromol. Chem. B: Phys., 17, 355.
Rink, M., Ricco, T., Lubert, W. and Pavan, A. (1978) J. Appl. Polym. Sci., 22, 429.
Robertson, R.E. (1974) ACS Div. Org. Coat. Plast. Prepr., 34(2),229.
Robeson, L.M. (1984) Polym. Eng. Sci., 24, 587.
Sardelis, K., Michels, H.J. and Allen, G. (1987) Polymer, 28, 244.
Schirrer, R., Fond, C. and Lobbrecht, A. (1996) J. Mater. Sci., 31, 6409.
Schneider, M. (1995) PhD Thesis, Strasbourg.
Sha, Y., Hui, c.Y., Ruina, A. and Kramer, E.1. (1995) Macromolecules, 28, 2450.
Shultz, A.R. and Gendron, B.M. (1972) J. Appl. Polym. Sci., 16,461.
Sih, G.c., (1973) Eng. Fract. Mech., 5, 365.
Sjoerdsma, S.D. and Boyens, 1.P.H. (1994) Polym. Eng. Sci., 34, 86.
Sjoerdsma, S.D. and Heikens, D. (1982) J. Mater. Sci., 17, 747.
Smith, T.L. (1958) J. Polym. Sci., 32, 99.
Soares, v.L.P. (1994) PhD Thesis, Cranfield University.
Spiegelberg, S.H., Argon, A.S. and Cohen, R.E. (1994) J. Appl. Polym. Sci., 53, 1251.
Starke, 1.U., Godehardt, R., Michler, G.H. and Bucknall, c.B. (1997) J. Mater. Sci.,
32, 1855.
Steenbrink, S. (1997) J. Mech. Sci. Solids, in press.
Sue, H.J. (1991) Polym. Eng. Sci., 31, 275.
Sue, H.J. (1992) J. Mater. Sci., 27, 3098.
Sue, H.J., Huang, 1. and Yee, A.F. (1992) Polymer, 33, 4868.
Tabor, D. (1994) Polymer, 35, 2759.
Tanaka, H., Hayashi, T. and Nishi, T. (1989) J. Appl. Phys., 65, 4495.
Taylor, G.R. and Darin, S.R. (1955) J. Polym. Sci., 17, 511.
Trassaert, P. and Schirrer, R.1. (1983) J. Mater. Sci., 18,3004.
Trent, 1.S., Schienbeim, 1.1. and Couchmann, P.R. (1983) Macromolecules, 16, 589.
Underwood, E.E. (1970) Quantitative Stereology, Addison-Wesley, Reading, Mas-
sachusetts.
Van der Sanden, M.C.M., de Kok, 1.M.M. and Meijer, H.E.H. (1994) Polymer, 35,
2991.
Van der Sanden, M.C.M., Meijer, H.E.H. and Lemstra, P.J. (1993) Polymer, 34,2148.
Vesely, D. and Finch, D.S. (1988) Makromol. Chem. Macromol. Symp., 16, 329.
Vincent, P.I. (1972) Polymer, 13, 558.
Vitali, R. and Montani, E. (1980) Polymer, 21, 1220.
Ward, I.M. (1983) Mechanical Properties of Solid Polymers, 2nd edn, John Wiley,
New Yark, p. 379.
Wu, S. (1992) Polym. Int., 29, 229.
Yang, H.H. and Bucknall, c.B. (1997) 10th International Conference on Deforma-
tion Yield and Fracture of Polymers, Churchill College, Cambridge, Institute of
Materials, p. 458.
Yap, O.F., Mai, Y.W. and Cotterell, B. (1983) J. Mater. Sci., 18, 657.
Yee, A.F. and Pearson, R.A. (1989) in Fractography and Failure Mechanisms in
Polymers and Composites (ed. A.C. Roulin-Moloney), Elsevier, London.
Interfaces 9
R.A. L. Jones

9.1 INTRODUCTION
Glassy polymers are often used as one component of a heterogeneous,
composite material; examples include fibre-reinforced plastics, filled poly-
mers and multiphase polymer blends such as rubber-toughened plastics. In
all these cases the macroscopic mechanical properties of the material depend
on the microscopic characteristics of the interface between the phases. The
characterization of such interfaces is generally rather difficult, but the last
10 or 15 years has seen a tremendous growth in our knowledge. This has
been gained by the use of sophisticated new techniques, such as neutron
reflectivity, applied to well controlled model interfaces, as well as the use of
more traditional mechanical testing methods applied to interfaces for which
microscopic information was available. It is hoped that this understanding
of model systems may now start to be transferred to the more complicated
types of problems involved in the actual use of glassy polymers in hetero-
geneous systems.
In considering the structure of any kind of interface, it is important to be
clear at the outset what length scale is relevant. On an atomic or molecular
length scale, one might be interested in the arrangement of polymer
segments and the nature of interactions between these segments, either with
each other or with the components of a non-polymeric material involved in
the interface. These interactions might include covalent bonds, hydrogen
bonds, acid-base interactions or dispersive interactions. This interfacial
chemistry is not fundamentally altered by the polymeric nature of one or
more of the components, but in a sense it may be regarded as being at the
origin of interfacial properties in all classes of materials. But the long chain
nature of polymers does lead to qualitatively new phenomena which are

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
414 Interfaces

characterized by larger length scales than that which characterizes the range
of chemical interactions; these length scales fall in the range from 1 to 10 nm,
in rough terms, and the polymer physics that emerges from applying
continuum ideas at these lengths is often of dominant importance to
mechanical properties of polymer containing systems. An example that
shows this very nicely is the way the fracture toughness of glassy polymers
depends on molecular weight. For example in polystyrene, going from a
molecular weight of 10000 to 200000 increases the fracture toughness by
more than a factor of 1000 (Wool, Yuan and McGarel, 1989). The chemical
interactions, the dispersive forces between neighbouring segments, are
exactly the same, but the radius of gyration of the chain has changed from
6.5 to 29.4 nm. It is this change on an intermediate length scale that has
caused this dramatic change in properties, and so we should not be
surprised to find that interfacial properties are also largely determined by
phenomena at similar length scales.
Theories of interfacial structure in polymers are largely based on statisti-
cal mechanics, so an important question to be borne in mind is the degree
to which an interface has had the opportunity to reach equilibrium. In a
glassy system rearrangement of segments on the relevant length scales is
extremely slow, so the structure of an interface usually reflects conditions at
some point in the system's processing history when sufficient segmental
mobility was present to enable the system to approach equilibrium. The
degree to which the system will be able to approach equilibrium will depend
both on the length of time spent in a mobile state, for example in a melt
processing step, and on the kinetics of approach to equilibrium in that
mobile state. Different processes may have very different kinetics; for
example, the broadening of an interface between two incompatible polymers
would be expected to occur rather quickly in the melt, whereas the
segregation of a surface-active additive to the interface might be very much
slower. So a system will very often be in a state of equilibrium with respect
to some processes but far from equilibrium with respect to others.
Dynamic properties are also interesting directly in the glassy state;
properties like fracture toughness will often be determined by processes
which involve the motion of segments in the glassy state. Much less is
understood about these aspects of interfaces in glassy polymers, though
substantial progress had been made by applications of the kinds of molecu-
lar concept that have been so fruitful in the understanding of crazing. In the
future, computer simulations will also play an important role in unravelling
these difficult questions, though at the moment they are still limited in terms
of the size of system and time scale that are accessible.
In this chapter three broad classes of interfaces will be considered. The
first class of interface is that between two polymers (Figure 9.1a). If at any
Introduction 415

(a)

t Interfacial width

(b)

Figure 9.1 Different types of polymer -polymer interfaces: (a) polymer -polymer,
showing interpenetration: (b) polymer-solid, showing chemically bound chains and
an interphase.

time during processing the polymers have had some mobility, interpenetra-
tion of the chains will take place to produce an interface that is diffuse on
a local level, even if the polymer pair are, like most polymer pairs,
immiscible in the bulk. The width of the resulting interface can be measured
by, for example, neutron reflectivity, and it can to some extent be predicted
by theory. For many pairs of immiscible polymers the degree of mixing is
rather small, which accounts for the weakness of such unmodified polymer-
polymer interfaces. However, such interfaces may be modified by various
interfacially active components such as block and graft copolymers, which
can lead to spectacular increases in interfacial fracture toughness. The study
of the mechanisms of this toughening has not only provided a guide to
making the most efficient interfacial compatibilizers, but has also provided
much fundamental insight into the molecular origins of interfacial strength
in general.
Interfaces between glassy polymers and inorganic materials, such as the
interfaces between polymers and filler particles or reinforcing fibres, at first
would appear to be very different from polymer-polymer interfaces. Good
adhesion is often achieved by arranging strong chemical or physical bonds
between the polymer and the inorganic surface, perhaps by the use of
primers or sizing agents. However, as sketched in Figure 9.1 b, this in fact
416 Interfaces

leads to another kind of polymer-polymer interface, this time between the


free polymer and the polymer chains that are directly attached to the
interface. There has been very little systematic study of such situations; at a
microscopic level the structure of only a few very highly idealized cases has
been studied. Perhaps the most important of these model systems is formed
when a number of chains are attached by a single end to the interface,
forming a 'brush'. The degree of interpenetration between these 'brush'
chains and the free polymer can be predicted by self-consistent field theories
and measured by neutron reflection, while the effect of the brush chains on
the adhesion between polymer and non-polymer can be directly measured.
Finally, we come to what at first appears to be the simplest kind of
interface, the free surface of a glassy polymer. However, even this apparently
simplest of situations conceals a number of serious difficulties. The presence
of a surface may seriously perturb both the conformation and the dynamics
of nearby polymer chains in ways that may well have a profound influence
on properties like adhesion and friction; at present, however, such effects are
hardly understood at all.

9.2 INTERFACES BETWEEN INCOMPATIBLE POLYMERS


Two immiscible glassy polymers, when welded above their glass transition
temperatures, will develop an interface which at room temperature has a
small but significant degree of strength. A useful measure of this strength is
the value of the fracture toughness for a failure mode in which a crack
propagates along the interface without any component of in-plane shear or
out-of-plane torsion. This quantity, GIc , is the sum of all the energies
required to propagate a crack along an interface. A variety of mechanical
tests may be used to measure this quantity, though many such tests tend to
overestimate GIc , because during the test additional energy dissipation
occurs in plastic deformation that takes place in the bulk of the test piece
rather than at the interface. An example of a simple and relatively reliable
test geometry, the double beam cantilever test (Kanninen, 1973), is shown
in Figure 9.2.
For an ideally brittle material, one might expect the fracture toughness to
be equal to the so-called work of adhesion WAB (lsraelachvili, 1991), which
is expressed in terms of the interfacial tension (JAB and the two surface
tensions (J A and (JB as
(9.1)
For example, for an interface between polystyrene (PS) and poly(methyl
methacrylate) (PMMA), one would expect a work of adhesion of about
0.08 J m - 2. However, when the fracture toughness of such interfaces is
Interfaces between incompatible polymers 417

Razor Polymer beams:


blade thiclmess 'h'

Crack length 'a'


Figure 9.2 A double beam cantilever test for measuring interfaCial fracture tough-
ness. Two welded polymer bars are driven apart by a razor blade of width fJ, and the
length of the crack ahead of the blade is measured. For bars with identical thickness
h and Young's modulus E the fracture toughness is given by
3fJ2Eh 3
G 1c = -:-16-=-a--:;-4-:-:{1-+----=-0.-=-64"-;(-:-h/:-:a)"'-;}4

If the bars have different moduli, an asymmetric test must be performed where the
thickness ratio of the two bars has to be selected to keep the crack propagating at
the interface.

measured experimentally, values of Gte between 10 and 20 J m - 2 are


obtained (Brown et ai., 1993). The toughness is much greater than the
theoretical work of adhesion, indicating that mechanisms giving rise to
substantial additional energy dissipation are at work.
Possible mechanisms that could give rise to such dissipation are discussed
below. However, in general we should note that the existence of such
mechanisms must depend on the fact that chains on either side of the
interface are to some extent entangled. The fracture toughness of uncross-
linked glassy polymers in the bulk is also intimately connected with
entanglement (Chapter 6); however, we note that bulk high molecular
weight PS and PMMA have fracture toughnesses in the range 500-
1000 J m - 2. Thus we can deduce that at the interface between these two
immiscible polymers the degree of entanglement is very much less than in
the bulk materials.
We now have some understanding of the limited interpenetration that
occurs at immiscible polymer-polymer interfaces based on both theoretical
and experimental work. Some years ago Helfand (Helfand and Tagami,
1971) found an elegant analytical solution to the Edwards (1965) self-
consistent field equations to describe polymer chain conformations at a
polymer-polymer interface in the limit of high molecular weight and strong
immiscibility. He predicted that the volume fraction profile of one polymer
through the interface ¢(z) is given by

(9.2)
418 Interfaces

«
~
0.8
c
0
13CU 0.6

-
E
....
Q) 0.4
~
(5 0.2
>
0
-3 -2 -1 0 1 2 3
Distance zlw 1
Figure 9.3 Theoretical volume fraction profile through an interface between high
molecular weight immiscible polymers, as given by equation 9.2, and illustrating the
definition of the intrinsic interfacial half-width WI·

where the intrinsic interfacial half-width WI is given by

(9.3)

where a is the statistical step length of the polymer and X is the Flory-
Huggins interaction parameter. Note that for highly immiscible polymers
with a degree of polymerization N, the volume fractions of the coexisting
phases are exp( - XN) and 1 - exp( - XN) which, in the limit of high
molecular weight and strong immiscibility that we are considering here,
approach zero and one respectively. The shape of this interfacial profile is
shown in Figure 9.3.
A simple argument shows how this functional form for the interfacial
width arises, and illustrates the physical reasons why an immiscible poly-
mer-polymer interface has a finite thickness. Consider an interface which is
molecularly narrow between two immiscible polymers A and B. The
polymer chains near the interface are only able to adopt a limited number
of their possible configurations, and so have a lower entropy than corre-
sponding chains in the bulk. In order to increase this entropy, loops of
polymer A will tend to cross the interface into polymer B, and vice versa.
The size of these loops will however be limited by the fact that each such
loop will increase the total energy of the system by virtue of the unfavour-
able A-B interactions that they give rise to. The balance between an
en tropic drive to a wider interface and an energetic drive to a narrow
Interfaces between incompatible polymers 419

interface determines the equilibrium width of the interface which leads to


the lowest possible system free energy.
We can estimate that at equilibrium the unfavourable energy of interac-
tion will be of order kT If a loop is on average N loop units long, we can
write the unfavourable interaction energy as

(9.4)
thus equating this with kT shows us that the average loop size N loop ~ 1/x.
The loop itself follows random walk statistics, so we can write for the
half-width of the interface WI:

'"
"I
~ a(N loop)1/2 ~ ~1/2 (9.5)
X
which is the same functional form as the more sophisticated theory predicts.
If we substitute in typical values for a and X, we find that interfacial
half-widths should be expected to fall roughly in the range 1-20 nm. For
example, the value of X for the immiscible pair PS and PMMA is 0.037 at
150°C, on the basis of an experimental determination of the temperature
dependence of X by small angle neutron scattering from diblock copolymers
(Russell, Hjelm and Seeger, 1990). Using this value in the theoretical
expression leads us to an estimate of the interfacial half-width of 1.4 nm. If
we compare this with the: value for the distance between entanglements in
PS, which is 9 nm (Chapter 6), we are able to understand why only partial
entanglement occurs at such immiscible polymer-polymer interfaces, and
thus why such interfaces are substantially weaker than the bulk materials.
The most satisfactory method of measuring the width of such narrow
interfaces is perhaps neutron reflectivity (Russell, 1990; Stamm and
Schubert, 1995). This technique relies on the fact that the propagation of
slow neutrons, with wavelengths between 0.2 and 1.5 nm, through a uniform
material is controlled by a refractive index which is determined by the way
in which the nuclei in the material interact with the neutrons. Just as a light
wave is reflected when it encounters an interface between two regions with
differing refractive indices, so a neutron beam is specularly reflected when it
encounters an interface between two regions wih different neutron refractive
indices. The neutron refractive index is given as a function of wavelength by

(9.6)

where the sum is over all the different types of nuclei in the sample, Pi
the number density of a particular nucleus and hi is the corresponding
neutron scattering length. fln is a linear absorption coefficient for neutrons
420 Interfaces

in the material; for most materials this is very small and can be neglected.
Scattering lengths vary from nucleus to nucleus in a seemingly unsystematic
way; most usefully for polymer science, the value for deuterium is very
different to the value for hydrogen, so by using deuterium labelling one can
prepare a version of a polymer with a very different neutron refractive index
to the corresponding unlabelled species.
The neutron reflectivity - the fraction of neutrons reflected from a sample
containing one or more interfaces - depends on both the angle of incidence
f) and the wavelength of the incident neutrons through the perpendicular
component of the neutron wave vector k, defined as

21t . CJ
k =Tsmu (9.7)

Just as is the case for light, if a neutron passes from a region with high
refractive index into a region with lower refractive index, for a low enough
angle of incidence or a large enough wavelength it will be totally reflected.
As the angle of incidence is increased or the wavelength is decreased, at a
critical value of the perpendicular component of the neutron wave vector
kerit the reflectivity will start to drop off from unity. For a sharp interface
the fall-off in reflectivity is described by the classical Fresnel equations, but
if the interface is diffuse or rough, at higher values of k - corresponding to
larger angles or smaller wavelengths - the reflectivity will be decreased
below the Fresnel value by an amount that depends quite sensitively on the
width of the interface (Figure 9.4).
Two groups have measured the interface width between PS and PMMA
using neutron reflectivity (Fernandez et al., 1988; Anastasiadis et al., 1990);
the values they obtained for the half-width agreed within error and was
2.5 nm (note that different authors do not always define the width of the
interface in the same way; this value is based on the definition in equation
9.2). Thus the order of magnitude of the experimental result agrees with
theory, but the absolute value from experiment is nearly twice as large as
the prediction. This discrepancy is far larger than the experimental errors in
determining both the interfacial width and the interaction parameter x.
The origin of this discrepancy is not yet fully understood. A likely
possibility arises from the fact that the calculation of the interfacial width is
a mean-field calculation; that is to say, it replaces the actual interaction that
a single chain feels due to all the other chains by an interaction with a
mean-field, which averages out the contribution of all the other chains. Such
methods neglect the effect of fluctuations. In this case we do expect
important fluctuations, because at any interface between two liquids we
expect to find thermally excited capillary waves (Shull, Mayes and Russell,
Interfaces between incompatible polymers 421

0.1

~
.s;
:;::::: 0.01
0
Q)
;:
Q)
CC 0.001

0.0001

o 0.02 0.04 0.06 0.08 0.1

Momentum transfer Q 1A-1


Figure 9.4 Calculated neutron reflectivity curves for an interface between
deuterated polystyrene and poly(methyl methacrylate) for various values of the
interface width. The solid curve is for a sharp interface, the dashed curved has an
interfacial half-width of 12.5 A (1.25 nm), and the dotted curve has an interfacial
half-width of 25 A (2.5 nm).

1993; Semenov, 1994}. In the polymer-polymer interface case, when the


system is in the melt state capillary waves will be thermally excited at the
interface, and when the polymers are cooled down below their glass
transition temperatures these transient waves will be frozen in place. As a
result the interface, in addition to having an intrinsic diffuseness, will also
be rough on a microscopic length scale. The technique of neutron reflectivity
will be unable to distinguish between an interface that is diffuse and one that
is rough on a length scale smaller than the coherence length of the neutron
(if indeed such a distinction is ever meaningful). This extra roughness will
lead to a correction to the interfacial width which can be expected to take
the following form (Shull, Mayes and Russell, 1993):

(9.8)

where w(L} is the effective interface width as measured by a technique that


is sensitive to roughness up to a length scale L (Figure 9.5). To reconcile the
experimental value for the interface width with the predicted one it is
422 Interfaces

2 w(L)

(a) (b)
Figure 9.5 (a) An interface without capillary waves, with some intrinsic interfacial
half-width WI . (b) The interface is roughened by thermally excited capillary waves. A
technique which averages over a distance L will measure larger apparent interfacial
half-width w(L).

necessary to assume that the length L over which the roughness is averaged
is of the order of 20 Ilm.

9.3 REINFORCEMENT OF POLYMER-POLYMER INTERFACES


WITH BLOCK COPOLYMERS
Our knowledge of the microscopic structure of immiscible polymer-poly-
mer interfaces gives us a qualitative understanding of why such interfaces
are rather weak. One might ask whether this microscopic structure can be
altered in a way that makes them stronger. It turns out that certain
polymeric additives can be used which preferentially locate themselves at
polymer-polymer interfaces. The modification of the interfacial structure
thus obtained can lead in some cases to order of magnitude improvement
of the interfacial fracture toughness.
Perhaps the best understood polymeric interfacial agents are block
copolymers. If we have an interface between two immiscible polymers A and
B, it is easy to see that a block copolymer, consisting of an A segment and
a B segment which are covalently linked, is likely to find it energetically
favourable to segregate to the A- B interface, allowing both blocks to find
themselves in a chemically identical environment, thus minimizing un-
favourable interactions, albeit at the cost of a loss of entropy. This of course
is exactly the reason that a small molecule surfactant will segregate to an
oil-water interface, and block copolymers, like surfactants, can thus reduce
the interfacial tension between the two immiscible phases. This reduction in
interfacial tension may have the effect of reducing the average domain size
in a dispersion of the two immiscible polymers, just as a surfactant can help
emulsify a mixture of oil and water, but there is another purely polymeric
effect that is, if anything, more important. If the blocks are long enough,
each block may form entanglements with the polymer on its side of the
Reinforcement of polymer-polymer interfaces 423

interface, and it is the presence of these entanglements that can lead to


substantial increases in interfacial fracture toughness.
The mechanisms by which interfaces can be toughened by block
copolymers have been elucidated by a series of careful experiments on well
controlled materials (Brown, 1989, 1991a; Creton et ai., 1992; Brown, 1993;
Brown et ai., 1993; Char, Brown and Deline, 1993; Creton, Brown and
Deline, 1994; Kramer, 1994; Washiyama et ai., 1994). These experiments are
important, not only because of the light they cast on the specific mechanisms
by which block copolymers reinforce interfaces, but also because they have
led to a much more general understanding of the origins of adhesion at
polymer interfaces.
The interfacial activity of block copolymers can be observed in experi-
ments in which a bilayer is prepared from two immiscible polymers A and
B, in which one polymer film has added to it a deuterium-labelled block
copolymer of A and B. Forward recoil spectrometry (for recent reviews of
this powerful technique see Shull, 1992, and Jones, 1993) can be used to
detect the depth profile of deuterium in such a sample; on annealing one

0.10

....
Q) 0.08
E
~
0
a.
0 0.06
0
c::
0
U
-
~
Q)
E
:::J
0.04

(5 0.02
>

0.00
-2000 0 2000 4000 6000 8000
Depth fA

Figure 9.6 Distribution of a deuterated styrene-2-vinyl pyridine diblock copolymer


in a bilayer of polystyrene and poly(2-vinyl pyridine), as revealed by forward recoil
spectrometry. The dashed line is a fit to the copolymer distribution in the sample as
prepared by spin casting, while the points and solid line are the data and fit when
the sample has been annealed for 8 h at 178°C. This shows a substantial excess of
copolymer at the interface between the two homopolymers. (After Shull et al.,
1990.)
424 Interfaces

finds an accumulation of deuterium at the polymer-polymer interface


indicating that the copolymer has segregated there (Figure 9.6). By making
such measurements as a function of the amount of block copolymer present
in the initial film, one can construct adsorption isotherms for the block
copolymer (Shull et at., 1990). An example of such an isotherm is shown in
Figure 9.7; the shape of these curves can be predicted accurately using
self-consistent field theory (Shull and Kramer, 1990). Note that theory is
only able to predict the adsorbed amount up to about a 5% copolymer
content, at which the interfacial excess rises sharply. This rise is associated
with the formation of micelles, which themselves segregate both to surfaces
and interfaces. If one wishes to maximize the amount of block copolymer at
the interface in the form of free chains, rather than micelles, there is no point
in using a higher concentration of copolymers than the critical micelle
concentration (CMC); any excess of block copolymer above the CMC will
in effect be wasted. The CMC itself is a very strong function of molecular
weight, so that very long block copolymers will tend to have extremely small
CMCs. In addition to this, long block copolymers will tend to diffuse rather
slowly, so equilibrium will take a long time to be reached.

300

250

~
III 200
III
Q)

~
a; 150
·0
~
~ 100
.5
50

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.Q7
Volume fraction copolymer in PS phase

Figure 9.7 Excess of styrene-2-vinyl pyridine block copolymer at a polystyrene-


poly(2-vinyl pyridine) interface, as determined by forward recoil spectrometry. The
solid line is the prediction of self-consistent field theory. (After Shull, et at., 1990.)
Reinforcement of polymer~polymer interfaces 425

1.0 ~-
/

0.8 /
c
.2
t> 0.6
~
Q)
E
::J 0.4
(5
>
0.2

0.0
-300 -200 -100 o 100 200 300
Depth fA

Figure 9.8 Segment distributions of a styrene-(methyl methacrylate) block


copolymer at an interface between polystyrene and poly(methyl methacrylate), as
revealed by a series of neutron reflection experiments in which various parts of the
copolymer and/or one of the homopolymers was labelled with deuterium. The bold
lines are the segment density profiles for all styrene and methyl methacrylate
segments, summed over both the homopolymer and the copolymer; the solid lines
are the homopolymers, and the dotted lines the styrene and methyl methacrylate
blocks of the copolymer. (After Russell et aI., 1991.)

Forward recoil spectrometry provides a reliable measurement of the


amount of block copolymer at an interface, but its depth resolution is too
large to allow one to resolve the details of the way the block copolymer
chains are organized at the interface. This can be achieved once again by
the use of neutron reftectometry (NR). By making NR measurements on a
series of samples in which different polymers or parts of polymer are
deuterium labelled but which are otherwise identical, one can build up a
very complete and detailed picture of the block copolymer modified inter-
face (Russell et ai., 1991). Figure 9.8 shows the results obtained for a
styrene - methyl methacrylate block copolymer at a PS - PMMA interface.
Once again, one can account quantitatively for all the features of such
curves using self-consistent field theory. Let us however concentrate on
qualitative features. If one looks at the width of the interface between the
total styrene and MMA segments, one sees that it is still relatively narrow.
426 Interfaces

However, at the interface most segments are part of block copolymer chains,
and when one looks at the interface between the MMA segments of the
block copolymer and the PMMA homopolymer, revealed by deuterating
only the MMA block of the copolymer, one sees a much broader interface,
which one would expect to lead to a substantial degree of entanglement. The
interface between the styrene block of the copolymer and the PS homo-
polymer is very similar. Thus both halves of the block copolymer should be
well entangled with their respective homopolymers, and since they are
covalently linked, this effectively results in the two homopolymers being
stitched together. We would thus expect this to lead to substantial increases
in fracture toughness.
The toughening effects of block copolymers at interfaces have been
extensively studied using mechanical tests of the type shown in Figure 9.2.
This kind of experiment was pioneered by Brown (1991a); one comprehen-
sive study was carried out by Creton et al. (1992), looking at the reinforce-
ment of an interface between PS and poly(2-vinyl pyridine) (PVP) with
block copolymers of styrene and 2-vinyl pyridine. PS and PVP are highly

~
o
o
E
:::2 o o
.!.!
(!l 8
l8CI) o
s:::
..c:
100
"
"It £""
Cl 0" 0 0

.s
::l
o
~ f5' 0
::l 00

-~ "
00 o
10
1ii

~
o

.93
"
.E

0.02 0.04 0.06 0.08 0.'


Crossing ch'ains per unit area Inm-2
Figure 9.9 Fracture toughnesses of interfaces reinforced by block copolymers as a
function of effective areal density of chains crossing the interface. Triangles and
squares are for polystyrene-poly(2-vinyl pyridine) interfaces reinforced with styrene-
2-vinyl pyridine block copolymers (Cretan et al., 1992); circles are for poly(phenylene
oxide)-poly(methyl methacrylate) interfaces reinforced with styrene-methyl meth-
acrylate block copolymers (Brown, 1991a). (After Creton et al., 1992.)
Reinforcement of polymer-polymer interfaces 427

incompatible, and the fracture toughness of the unreinforced interface is


only of the order of 1 J m - 2. Figure 9.9 shows the fracture toughnesses that
can be achieved as a function of the amount of block copolymer at the
interface when a relatively high molecular weight block copolymer is used.
This shows that increases in fracture toughness of orders of magnitude are
possible with block copolymer reinforcement. Over a wide range of
copolymer coverage, the fracture toughness rises in proportion to the square
of the copolymer coverage; at lower coverage the increase in Gle is closer to
a linear relationship with the coverage, while at very high coverage a plateau
in the fracture toughness is reached.
Additional insight into the mechanisms by which block copolymers can
toughen interfaces is obtained by experiments in which the molecular
weights of the blocks are varied. Figure 9.10 shows the results of such an
experiment. For very short blocks, only modest increases in toughness are
achieved; once the block length exceeds the entanglement length, the
toughness increases rapidly, directly confirming the importance of entangle-
ments for interfacial toughness.
From these kinds of experiments, supplemented by surface analysis of the
fracture surfaces to determine which side of the joint the block copolymer

1200

1000

S 800
CD

b..... entanglements
:::::::-
600
~
......
£
C!)
400

200

0 200 400 600 800 1000


PVP degree of polymerisation
Figure 9.10 Fracture toughness per reinforcing chain for a polystyrene-poly(2-vinyl
pyridine) interface reinforced with styrene-2-vinyl pyridine block copolymers, as a
function of the degree of polymerization of the 2-vinyl pyridine block. (After Creton
et al., 1992.)
428 Interfaces

(a)

(b)

(c)

Figure 9.11 Possible modes of failure of a block copolymer reinforced interface:


(a) pull-out. (b) scission and (c) crazing.

ends up after failure of the joint, it has proved possible to identify three main
mechnisms of failure. For block copolymers in which one block is relatively
short, failure may occur by the pulling out of one of the blocks from its
homopolymer (Figure 9.l1a). The chain will feel some frictional force that
resists its being pulled out, which is presumably proportional to the block
length N b if the block is not completely entangled. The fracture toughness
Reinforcement of polymer-polymer interfaces 429

can then be estimated as the work done against this frictional force as the
chain is pulled completely out, which leads to the prediction

(9.9)

where ~ is the number of block copolymer chains per unit area. In general
one should expect the frictional force to be a function of the velocity at
which the chain is pulled out, but the nature of this dependence is unknown,
and experimentally the fracture toughness in this regime does not seem to
be a strong function of velocity over the range of velocities at which
experiments are carried out.
On the other hand, if both blocks of the copolymer are relatively well
anchored, the force on an individual chain may exceed the force required to
break the backbone of the chain, and failure will occur by scission (Figure
9.llb). If this force is ib' then the stress the interface can sustain is

(9.10)

The fracture toughness associated with this chain scission mechanism is


more difficult to estimate, but it must be small because the size of any zone
of plastic deformation at the crack tip must itself be smalL of the order of
the size of the monomer.
Both the pull-out and scission mechanisms can only occur when the
coverage of block copolymers is relatively low. As the coverage increases,
the stress at the interface will exceed the stress at which the material on one
or other side of the interface will craze (Figure 9.11c). The formation of a
craze at the interface will be associated with a large increase in the fracture
toughness, because of the large volume of material that undergoes plastic
deformation in a craze. In fact, it is now becoming clear that crazing is the
only mechanism of adhesive failure at glassy polymer-polymer interfaces
that can lead to usefully large values of the interfacial fracture toughness.
The fracture toughness of an interface which fails by crazing has been
calculated by Brown (1991b). Here is a highly simplified version of his
argument.
Through most of the craze the stress has a constant value acraze; this
crazing stress is essentially a material constant for a given type of polymer.
This stress is carried primarily by the main craze fibrils which run perpen-
dicular to the craze-bulk interface. However, in addition to the main craze
fibrils, cross-tie fibrils connect the main craze fibrils laterally, and these
permit some stress transfer in this lateral direction (Figure 9.12). This means
that at the crack tip there is a stress concentration, which causes the
breakdown of the last load-bearing fibril and the growth of the crack. At
the crudest level we can model the craze as an elastic continuum, and write
430 Interfaces

crac tip - -HO:.

craze fibril
d
Figure 9.12 Geometry of a craze breaking down to form a crack . (After Sha. Hui
and Ruina. 1995.)

the stress as a function of distance x from the end of the crack as

a(x) = acraze (~h)1/2 (9.11)

where h is the width of the craze. If the spacing between craze fibrils is d,
this means that tqe stress on the last load-bearing craze fibril is

afibril '" a craze (dh)1/2 (9.12)

The craze will break down when the stress on the last load-bearing fibril
reaches a value

(9.13)

where ib is the force to break or pull out a single reinforcing chain;


reinforcing chains are present at an areal density of l: (in fact this is an
Reinforcement of polymer-polymer interfaces 431

overestimate as some of the chains are broken or pulled out when the craze
forms). This allows us to write an expression for the craze width h:
LZJbZd
h=-z- (9.14)
(J craze

In the Dugdale model of craze micromechanics, we can estimate the fracture


toughness as
(9.15)
where be' the critical crack opening displacement, is related to h via
(9.16)

where Vf is the volume fraction of fibrils. Thus, finally, we are led to an


expression for the fracture toughness:

GIe L2Jb2d(1 - Vf )
~ (9.17)
(J craze

This correctly accounts for the observed quadratic dependence of fracture


toughness on the areal density of reinforcing chains. A more sophisticated
version of the same argument would treat the craze as an anisotropic elastic
continuum (Brown, 1991b), while for weak interfaces the craze is rather
narrow, and the approximation of treating the crazed material as a continu-
um breaks down. For weak interfaces a discrete treatment of craze break-
down is now available (Sha, Hui and Ruina, 1995); in these circumstances
the quadratic dependence of fracture toughness on areal coverage breaks
down and the fracture toughness rapidly drops off as the coverage is
reduced.
One might also ask what limits the attainable degree of fracture toughness
for very large coverages of block copolymer. If the coverage exceeds that
corresponding to half a period of the lamellar structure that the block
copolymer would form in the bulk, then additional copolymer will not be
able to fit in at the homopolymer interface but will instead start to form a
new lamella (Figure 9.13). Thus this additional block copolymer will not
add any more reinforcement to the interface, and indeed it may start to
decrease the interfacial fracture toughness, because there may be less
entanglement between the blocks of the copolymer than with the blocks and
the respective homopolymers. In this case failure will occur not at the
original interface, but within the lamella.
Armed with this understanding of the interfacial activity of block
copolymers and the mechanisms by which block copolymer-reinforced
interfaces fail, it should in principle be possible to design a block copolymer
432 Interfaces

crack
path

Figure 9.13 An interface saturated with block copolymer. Failure occurs not at the
original interface, but within the block copolymer lamella, so any further block
copolymer added will not increase the fracture toughness of the interface. (After
Creton, Brown and Deline, 1994.)

molecule which at a certain interfacial coverage would give optimum


reinforcement to an interface. However, it might be argued that such a
technique would be unlikely to find widespread use, because of the expense
of such carefully synthesized molecules. However, there are other methods
of reinforcing interfaces which use the same basic principles as the block
copolymers but with less specialized materials.
Two examples of reinforcing strategies that do not use block copolymers
are the use of reactive grafting at interfaces and the use of random
copolymers. In both of these methods, the same basic principle of reinforce-
ment operates in block copolymers: a molecule straddles the interface in
such a way that a single, covalently bonded molecule is well entangled with
the homopolymers on both sides of the interface.
In a reactively grafted interface, copolymers are effectively created in situ
at a polymer-polymer interface (Hobbs, Bopp and Watkins, 1983); at least
some of each of the polymers have complementary reactive groups which,
when they meet at the interface, can produce a covalent bond between
polymer strands from each side of the interface (Figure 9.14). This method
of controlling interfacial properties is effective and often used in practice,
though there have been few studies at the microscopic level of the way such
Reinforcement of polymer- polymer interfaces 433

(a) (b)
Figure 9.14 A reactively grafted interface. Some polymers on each side of the
interface have reactive groups (a); these can react at the interface to form an in situ
graft copolymer (b)

interfaces are modified by the reaction. Fleischer, Morales and Koberstein


(1994) studied the modification of interfacial tension between poly(dimethyl
siloxane) (PDMS) and polybutadiene (PBD) when some of the PDMS was
terminated by an amine group and some of the PBD had a carboxy
termination. Complexation of the two reactive groups at the interface
resulted in an appreciable lowering of the interfacial tension. Norton et al.
(1995) studied the grafting of polystyrene with a carboxy end-group at an
interface between polystyrene and a thermosetting epoxy resin. The unmodi-
fied interface had a fracture toughness of 4 J m - 2. If relatively short
polystyrene chains with functional end-groups were introduced to the
system, with molecular weights below the entanglement molecular weight,
very little enhancement of fracture toughness was observed. Surface analysis
of the failure surface conjirmed that the polystyrene chains had become
anchored to the epoxy network, but that they had been pulled out of the
polystyrene homopolymer. However, for functionalized polystyrene chains
longer than the entanglement molecular weight, substantial increases in
fracture toughness were observed, with maximum fracture toughnesses of
more than 100 J m - 2. For these long functionalized chains a sharp transi-
tion was noted as a function of coverage of grafted chains; for low coverages
the interface failed by scission of the functionalized chains and the fracture
toughnesses were rather low, but when sufficient grafted chains were present
to allow the interfacial stress to reach the crazing stress for polystyrene, the
interface failed by crazing and very large increases in fracture toughness
were observed.
Our final example of an interfacial agent that can cause significant
increases in fracture toughness is perhaps the most surprising; these are
random copolymers. Brown et al. (1993) discovered that a random styrene-
methyl methacrylate copolymer could increase the fracture toughness of a
ps- PMMA interface up to a value of about 80 J m - 2. This is of course a
434 Interfaces

less dramatic increase in strength than is possible with diblock copolymers,


but it is still a respectable value, given the potential overwhelming economic
advantage of random copolymers over diblocks. Subsequent work (Dai
et al., 1994) has demonstrated that the best reinforcing effect is obtained
when the copolymer is most symmetric in its composition. The suggestion
is that the copolymer chain makes multiple crossings of the interface
(Yeung, Balazs and Jasnow, 1992), with each loop being well entangled with
the homopolymer. Any substrand of the random copolymer will have a
composition that differs from the average purely as a result of the statistical
nature of the copolymer; in an A-B random copolymer some sub strands
will be richer in component A than average, and will tend to form loops into
the A homopolymer, while other substrands will be richer in component B
and will form loops into the B homopolymer. The result is an interface that
is stitched together by multiple crossings in a way that is highly effective at
enhancing its fracture toughness.
A final point to note from this work is that although the strength of a
random copolymer reinforced interface is sharply maximized for a symmet-
ric copolymer composition, the interfacial width as a function of copolymer
composition varies relatively smoothly. This highlights the importance of
long loops, which may be present at rather low concentrations, in conferring
strength to polymer-polymer interfaces.

9.4 GRAFTED CHAINS AT POLYMER-SOLID INTERFACES


It has long been suspected that the region of polymer matrix closest to
the interface with a solid reinforcing particle or fibre has properties which
are different to those of bulk polymer; however, it is difficult to probe by
experiment the details of the properties of this so-called 'interphase'.
Recently, however, some progress has been made in the context of some
highly idealized model systems.
We can distinguish between at least three aspects of the interaction of a
polymer with a solid material. First, there is a simple steric effect; the
presence of an impenetrable hard wall prevents the polymer molecules from
taking up certain configurations. There may also be changes in density due
to packing constraints, and these density changes may in themselves lead to
the polymer segments nearest the wall having different dynamic properties
to segments in the bulk. Some of these effects will be discussed in the next
section.
Second, specific chemical or physical interactions may take place between
the polymer segments and the wall, including chemical bonds (often sought
out by the use of primers and sizing agents), hydrogen bonds and acid-base
interactions. The details of this kind of interfacial chemistry is beyond the
Grafted chains at polymer-solid interfaces 435

scope of this chapter. Obviously the polymer segments affected by interac-


tions will have very different properties to those segments in the bulk.
In addition to those segments that are directly affected by such chemical
interactions, the connectivity of the chain means that non-interacting
segments attached to the same chain as interacting segments are also
strongly affected by the interface, even though they may be quite far away
from it. The structure of the interface between those chains that have some
segments chemically interacting with the wall and those chains which are
completely free is likely to be of great importance in determining the overall
properties of the polymer-solid interface. One might have extremely strong
bonding between the solid and a layer of grafted chains, but if the interface
between these grafted chains and the free chains of the bulk is weak, it is
this that will limit the strength of the interface as a whole.
One situation in which the conformation of chains grafted to a polymer-
solid interface has been studied in some detail is the idealized system in
which monodisperse chains have been attached by one end to a model
planar interface. One can label these end-grafted chains with deuterium, and
use ion beam analysis techniques and neutron reflectivity to study the
density of end-grafted chains and their conformation in some detail.
Attachment of the chains may be either chemical or physical. In one early
study, polystyrene chains terminated by a short butadiene block with a
silane group on the end were used (Jones et ai., 1992); in the experiments a
film consisting of a mixture of deuterium-labelled end-terminated polymers
and a normal polystyrene matrix were annealed above the glass transition
temperature, under which conditions segregation of the functionalized
polymers to the interface occurred. Similar experiments have been carried
out with other end-groups which interact via physical or chemical interac-
tions with the solid interface; most strikingly, it turns out that even a single
carboxylic acid group on the end of a long polystyrene molecule is sufficient
to lead to a substantial degree of segregation.
These results emphasize the extreme sensitivity of polymer interface
properties to small perturbations. This arises because the rather small
entropy of mixing between polymers can be overcome by any interaction
with a wall, as long as its energy is comparable to or larger than kT(Figure
9.15). This is a very relaxed condition, which implies that all sorts of
adventitious chemical groups along a polymer chain could lead to the
formation of a grafted layer.
Given that a grafted layer at a polymer-solid interface may be relatively
easily achieved, we may ask what determines the nature of the interface
between the free and the grafted chains, and in particular how much
interpenetration there is between them. The key factors determining
how much interpenetration there is seem to be the molecular weights of
436 Interfaces

(a) (b)

Figure 9.15 A free polymer (a) has an end group with an energy of interaction with
the polymer matrix of Bpal ' The energy of interaction of the end group with the
interface is Bint ; if the net change in energy in putting the end group on the wall,
(B int - Bpol), outweighs the loss of translational entropy due to the localization of the
chain at the wall, of order kT, then adsorption will take place (b).

the free and grafted chains, and the nature of the thermodynamic inter-
action between them. In terms of a free energy balance, there is a competi-
tion between the stretching energy of the grafted chains, the free energy of
mixing between free and grafted polymers, and the free energy penalty
required to distort the matrix chains. If the free chains are short compared
to the grafted chains, a substantial mixing energy gain will offset the
stretching energy and we would expect that the brush layer would be
penetrated and swelled by the free chains, resulting in a rather extended
grafted layer with a diffuse interface between grafted and free chains. On the
other hand, if the free chains are long the gain in free energy due to mixing
in the brush will be smaller and the energy cost of distorting the matrix
chains to fit inside the grafted layer will be larger, so the grafted layer will
be more compact and with a relatively sharp interface with the free chains.
These qualitative arguments may be made quantitative using self-consistent
field methods (Shull, 1991).
Neutron reflectivity experiments allow one to check the predictions of
such theories in some detail. For example, Clarke and coworkers (Clarke et
aI., 1995) used neutron reflectivity to study the conformation of polystyrene
chains labelled with deuterium, chemically grafted by one end to the thin
native silicon oxide coating on a silicon wafer. With a matrix of normal
polystyrene of different molecular weights, these trends are well displayed.
Figure 9.16 shows the resulting volume fraction profiles of the grafted
chains. For high molecular weight matrices the profiles are all rather similar,
but when the molecular weight of the matrix drops well below the molecular
weight of the brush, the brush chains start to extend out into the matrix.
These volume fraction profiles can be quantitatively accounted for by the
self-consistent field theory.
If the matrix chains are chemically different from the grafted chains, the
degree of interpenetration must depend on the nature of the chemical
Grafted chains at polymer-solid interfaces 437

-156kPS
- - ·52kPS
en ·········28.5kPS
a.. 7kPS
'0
c:::
o
"
""
-
~
Q)
E
0.4
. ,
"., ""
:l
g "'< ." " .

0.2 .... ."


"'"
-.. ....~
.."':-,. ........
. ....
-'-""'"
o
o 50 100 150 200 250 300
distance from substrate / A
Figure 9.16 Volume fraction profiles of a chemically end-grafted polystyrene chain
as a function of the matrix molecular weight, as revealed by neutron reflectometry.
(After Clarke et aI., 1995.)

interaction between the grafted and free chains. If the two polymers are
miscible, then a favourable free energy of mixing will offset a larger degree
of stretching of the grafted chains, resulting in an extended, diffuse grafted
layer. On the other hand, if the grafted and free chains have an unfavourable
thermodynamic interaction that would make them immiscible in the bulk,
there will be very little interpenetration between the grafted layer and the
matrix, and the interface between them should resemble the interface
between the corresponding homopolymers,
Clarke and coworkers (Clark et al., 1995) also studied this situation;
Figure 9.17 shows the volume fraction profiles obtained for grafted PS
chains with a matrix of poly(vinyl methyl ether) (PVME), which is miscible
with polystyrene at room temperature, and with poly butadiene and PMMA,
both of which are immiscible with PS. Also shown, for comparison, is a
profile for a PS matrix. Qualitatively the trends are entirely as expected; the
grafted layer with the PVME matrix is highly extended, with a very diffuse
interface, while with the PBD and PMMA matrices the grafted layer is
much more compact, with a rather narrow interface with the free chains.
As one might expect, the case of a chemically identical matrix is inter-
mediate between these two extremes. In the case of the PVME matrix, fair
agreement is achieved between the experimental profile and the prediction
of self-consistent field theory; the agreement is not nearly as good as for the
438 Interfaces

1.00

-PBD
0.80
- - -PMMA
- - -29kPS
C/) ........ -PVME
a.. - -ppo-ps
"C
t:: 0.60
o

-~
Q)
E 0.40
:::J .....
~ ,
0.20
'-- .. ~ "
-'"''.~. --.~ .. -.~
0.00
o 50 100 150 200 250 300 350

distance from substrate / A


Figure 9.17 Volume fraction profiles of a chemically end-grafted polystyrene chain
for different matrix polymers, as revealed by neutron reflectometry_ Polybutadiene
(PBD) and poly(methyl methacrylate) (PMMA) are immiscible with polystyrene in the
bulk, polystyrene (PS) is chemically identical with the grafted chains, while poly(vinyl
methyl ether) (PVME) and a blend of poly(phenylene oxide) and polystyrene (ppo-ps)
are miscible with polystyrene in the bulk. (From Clarke et al., 1995.)

PS matrix but this perhaps is not unexpected given the greater uncertainties
in the parameters used as input to the theory and the fact that the PVME
used was polydisperse. However, there seems to be an unexplained discrep-
ancy between theory and experiment for matrix polymers with an unfavour-
able interaction; the interface between the grafted layer and the free chains
seems to be systematically broader than that measured or predicted between
homopolymers, although the trend is in the correct direction in the sense
that the more immiscible polymer pair has the narrower interface. The
source of this discrepancy is as yet unexplained.
The mechanisms by which such a grafted layer will toughen an interface
between a glassy polymer and a non-polymeric material are expected to be
closely analogous to the mechanisms by which polymer-polymer interfaces
are reinforced by block copolymers. However, rather less work has
been done to elucidate these mechanisms at a molecular level of detail,
largely because the required, accurate fracture toughness measurements are
rather difficult to carry out (Smith, Kramer and Mills, 1994). Smith and
coworkers have carried out perhaps the most extensive study, on the effect
Chain conformation and dynamics in glassy polymers 439

of polystyrene-poly(vinyl pyridine) block copolymers on the adhesion


between a layer of polystyrene and a glass slide (Smith, Kramer and Mills,
1994).
In these experiments, adhesion between polystyrene and a soda-lime glass
slide was modified by the use of a block copolymer of poly(2-vinyl pyridine)
and polystyrene. The fracture toughness of an unmodified glass-polystyrene
interface was found to be about 1 J m - 2. This very low value indicates that
little, if any, plastic deformation occurs in the polystyrene ahead of a crack
tip. However, the addition of the block copolymer leads to an increase in
the fracture toughness, with maximum values of the order 40 J m - 2 being
obtained. The origin of this increase at the chemical level is believed to be
a strong hydrogen bond between silanol groups on the glass surface and the
pyridine groups of the block copolymer; if the glass is made hydrophobic
with a self-assembled monolayer of chlorodimethyl-octadecylsilane, no sig-
nificant increases in fracture toughnesses are obtained.
These kinds of experiments, involving planar interfaces and specially
synthesized, very well characterized molecules, are a long way from the real
situation in a composite material. However, these experiments have at least
pointed out a possible route that might lead to the design of interfaces from
first principles.

9.5 CHAIN CONFORMATION AND DYNAMICS IN GLASSY


POLYMERS NEAR INTERFACES
In this section I return to some of the fundamental questions about the
nature of an interface in a glassy polymer which although simple to pose,
are still difficult to answer. If we have a polymer glass next to a wall or a
free surface with which no part of the polymer has any preferential inter-
actions, how are the conformations of the chains located near the surface
altered, and how does the density of the polymer vary? Moving from static
to dynamic properties, how are the motions of polymer segments close to
the wall or surface altered? These questions have proved to be very difficult
to deal with theoretically, and experiments have been few and usually rather
indirect. Perhaps the greatest progress in this area has been made by
computer simulation.
The density profile of a polymer melt near a free surface is expected to
decrease smoothly from its melt value to zero over a distance of a few tenths
of a nanometre. This density profile, together with the surface tension, can
be calculated using the classical mean-field approach of van der Waals
(Poser and Sanchez, 1979; Rowlinson and Wid om, 1982) if the pressure-
volume-temperature (PVT) properties of the melt are known or can be
modelled. These PVT properties are not strongly affected by the connect-
440 Interfaces

ivity of the polymer chains; a consequence of this is that neither the


characteristic width of the polymer-vacuum interface, nor the surface
tension, is a strong function of molecular weight for polymers. Such
molecular weight dependence as does exist in the surface tension can be
accounted for in terms of the known variations in bulk properties of the
corresponding polymers (Sauer and Dee, 1994).
It is far from obvious how this kind of equilibrium statistical mechanics
approach could be applied to a non-equilibrium system such as a polymer
glass. Nonetheless, measurements by neutron and X-ray reflectivity reveal
that the surface of a glassy polymer such as polystyrene has a roughness or
diffuseness of the order of 0.6 nm (Russell, 1990), which is close to what
would be expected for a melt. A very similar picture emerges from a
molecular dynamics simulation of the free surface of glassy atactic poly-
propylene (Mansfield and Theodorou, 1991b), which yielded an interfacial
thickness of about 0.7 nm.
The situation close not to a free surface but an impenetrable wall is rather
different. The presence of the wall imposes constraints on packing which can
lead to strongly oscillatory density profiles, as seen in a number of computer
simulations of polymer melts (e.g. Bitsanis and Hadzioannou, 1990; Kumar,
Vacatello and Yo on, 1990), and predicted by theory (Yet hi raj et ai., 1994).
A simulation of the interface of glassy atactic polypropylene with a graphite
surface (Mansfield and Theodorou, 1991a) showed a density maximum at
the wall followed by a shallower minimum, with bulk density recovered after
about 1 nm. The absence of oscillations here could be attributed to the
intrinsic disorder of the atactic chain. Although for the melt it is clear that
the length scale characterizing the recovery of the bulk density is smaller
than the radius of gyration of the chains, a recent simulation indicates that
the same may not be true for glassy polymers. Baschnagel and Binder (1995)
show that while at high temperatures the decay length for the wall-induced
density perturbation is of the order of the segment size, as the melt is
supercooled to a temperature close to the glass transition this length be-
comes larger even than the polymer radius of gyration.
Experimental evidence for these density variations near walls is sparse. In
small molecule liquids it is well known that such layering effects manifest
themselves by producing an oscillatory force between mica surfaces im-
mersed in fluid and brought together closer than about 5 nm (lsraelachvili,
1991). Similar oscillations have been observed for rather short chain
poly(dimethyl siloxane) molecules (Horn and Israelachvili, 1988); however,
any irregularity in either surface or molecule is found to suppress these
oscillations, which leads one to suspect that in practical polymer situations
these rather subtle effects may not be terribly relevant.
Chain conformation and dynamics in glassy polymers 441

Some experiments have been carried out to study the density profiles of
thin glassy polymer films (Fernandez et ai., 1990; Wu et ai., 1994) by X-ray
or neutron reflectivity. In both studies rather a thick layer ( ~ 4 nm) of low
density material was detected at the surface of the films immediately after
they were prepared by spin-casting. On annealing this diffuse layer disap-
peared, suggesting that it was a non-equilibrium product of the sample
preparation technique, but in the former study some films showed a density
gradient increasing towards the surface. These observations remain unex-
plained.
Moving from average properties such as the density profile to the details
of chain conformation, many computer simulations have now shown that
near both a hard, non-interacting wall and a free surface, one expects the
conformation of polymer chains to be perturbed; in a melt the chains near
a wall are oriented parallel to it and have a 'pancake' -like conformation.
This flattening of the chain conformation persists over a distance away
from the wall comparable to the radius of gyration of the chains. Few of
these studies have specifically addressed polymers in the glassy state.
Among those that have are a molecular mechanics simulation of the
surface of glassy atactic polypropylene (Mansfield and Theodorou, 1990)
and a molecular dynamics simulation of a siinilar system (Mansfield and
Theodorou, 1991b; Theodorou, 1992). Baschnagel and Binder (1995) have
explicitly considered the effect of temperature on conformational properties
in a glass-forming polymer as the system is supercooled.
Figure 9.18a shows one result of this simulation: the values of the radius
of gyration of polymer chains as a function of the distance from the wall of
their centre of mass. This clearly shows that while chains at the centre of the
film are isotropic, towards the walls the parallel component of the radius of
gyration increases while the perpendicular component decreases, indicating
that the chains assume anisotropic, flattened conformations near the wall.
Bulk-like behaviour is recovered at distances greater than about twice the
radius of gyration from the wall. Figure 9.18b shows similar data at a much
lower temperature, displaying similar qualitative features but greatly exag-
gerated in extent. Here the influence of the wall seems to extend farther into
the bulk than it does in the high temperature case; another length scale
characterizing the decay of the perturbation due to the walls has appeared
which is longer than the radius of gyration. The theoretical origin of this
length is not yet understood.
Very few experimental studies of conformations of polymer chains near
surfaces have been possible so far, due to the extreme difficulty of obtaining
such information. Pioneering experiments have been carried out by Russell
and coworkers using grazing incidence X-ray diffraction (Factor, Russell
442 Interfaces

20

o o

o o

o Xv "XX 0
'0 W 'W
X
X

X
X

X
X

X
x
o
o 5 10 15 20 25 30 35 40
Distance
30

o o

o o
o o
00

00
DOn, n

XX
X
X
Xx
X
X X

X Xx
X

'v
o
o 5 10 15 20 25 30 35 40
Distance
Figure 9.18 Component of the radius of gyration parallel to (0), and perpendicular
to ( x) a hard, non-adsorbing wall in a dynamic Monte Carlo simulation using the
bond fluctuation lattice model. (a) Results for a relatively high temperature, corre-
sponding to a polymer melt; (b) the corresponding results for a strongly supercooled
melt. (After Baschnagel and Binder, 1995,)
Chain conformation and dynamics in glassy polymers 443

and Toney, 1991, 1993, 1994); in this technique X-rays derived from a
synchrotron are incident on a thin film of polymer at less than the critical
angle for total reflection. An exponentially damped evanescent X-ray field
penetrates a short distance below the surface and scattering provides
structural information specific to this near surface region. In a polyimide
film, specifically poly(pyromellitic dianhydride oxydianiline) or PMDA-
ODA, it was found that there was a markedly enhanced degree of structural
order near the surface compared to the bulk. However, this tendency is only
observed for polymers which have some tendency to order in the bulk; it is
not seen in flexible polymers such as polystyrene.
One particular conformational perturbation that may be induced by the
presence of a wall has received a lot of attention; that is the possibility that
chain ends may be preferentially segregated to such an interface. The
argument is that, on purely en tropic grounds, chain ends would be preferred
at an interface because this minimizes the loss in conformational entropy.
Such an effect has been observed in many computer simulations, including
those cited above that specifically address glassy polymers (Mansfield and
Theodorou, 1991b; Theodorou, 1992; Baschnagel and Binder, 1995); the
distance over which chain ends are enhanced seems to behave in the same
way as the distance over which the density is perturbed. The experimental
situation is not clear, partly for reasons that we discussed in the section
above; chemically different end groups may strongly segregate to surfaces
and interfaces for enthalpic reasons in addition to these entropic factors, and
in practice it may be difficult to disentangle the two effects, as one usually
needs some kind of label in order to be able to distinguish the chain end.
Certainly segregation of end groups to surfaces has been unequivocally
observed; in addition to the examples given above of end-group segregation
to polymer-solid interfaces, one could mention a study in which polystyrene
with an end group of perfluorooctyl dimethyl chlorosilane was shown to
strongly segregate to the surface (Affrossman et al., 1994). A minimally
perturbing end group would be a short deuterated block, though even here
one should be aware of a small isotopic difference in surface energy between
normal and deuterated polymers which itself can drive surface segregation
(Jones et al., 1989). End-group segregation in a triblock copolymer of
normal and deuterated polystyrene was observed in one study (Affrossman
et al., 1993); another study on a very similar system (Zhao et al., 1993) could
only put an upper limit on the amount of segregation; their data were
consistent with a degree of end segregation of no more than a factor of two.
Just as the conformations of polymer chains are modified by the presence
of an interface, one might expect the dynamics of the chain also to be
affected. Such effects may well be important in understanding adhesion,
but unfortunately virtually no direct experiments to probe differences in
444 Interfaces

mobility between chains in the bulk and those near a surface are available.
Apart from a few more indirect measurements, discussed below, what
understanding we have of this important subject comes from computer
simulations.
A closely related question concerns the way the glass transition tempera-
ture of a polymer is affected either by the presence of a surface or an
interface, or by the mere fact of being in a thin film rather than bulk form.
An indication that the glass transition behaviour of thin films may not
necessarily be the same as for bulk polymers can be found from simple
dewetting experiments. A thin polystyrene film will dewet an untreated
silicon wafer substrate at equilibrium; however, a continuous film can be
made by spin casting, and of course dewetting cannot take place while the
film is in its glassy state. If such a film is now heated, the temperature at
which dewetting does occur is found to decrease as the film is made thinner
(Reiter, 1993, 1994). The conclusion must be that in this experiment, thinner
films have a higher mobility than bulk polymer; however, it is not at once
clear whether this is a confinement effect, possibly due to the fact that the
films are thinner than the unperturbed dimensions of the polymer chains, or
a reflection of the importance of the surface and interface in such thin films.
Direct measurements of the effect of size on the glass transition tempera-
tures of thin polymer films are now becoming available. The glass transition
temperature is marked by, among other things, a discontinuity in expansiv-
ity, so a convenient way of measuring it in a thin film is to follow the film
thickness as the temperature is increased or decreased. This can be conve-
niently done by ellipsometry or X-ray or neutron reflectivity. Initial results
(Keddie, Jones and Cory, 1994) were obtained for thin polystyrene films
heated in air on silicon wafers etched to remove the native oxide. Results
are shown in Figure 9.19; at thicknesses below about 100 nm the glass
transition temperature drops below its bulk value, with the transition
temperature depression exceeding 20°C for films of a few tens of nanometres.
Significantly, the effect seems to be essentially independent of molecular
weight, suggesting that the distortion of chain configurations in very thin
films is not the major cause of the ~ depression.
Further results have been obtained for other polymers and other film
substrates. For example, PMMA shows a decrease in ~ very similar to that
of polystyrene on silicon if the experiments are carried out on a gold-
coated silicon substrate (Keddie, Jones and Cory 1995), while if native
oxide-coated silicon is used there is a very small increase in ~ (Keddie,
Jones and Cory, 1995; Wu, v. Zanten and Orts, 1995). What seems to be
happening is that each interface in a thin film has its own effect on the
mobility of nearby chains; where there is a strong interaction between
polymer segments and the interface, this results in a decrease in mobility
Chain conformation and dynamics in glassy polymers 445

[
380
g o
375
;<>~
~
~

e
::J
370
(I)

~
Co
E 365
(I)
I
l-

I
i
c: 360
,g
'iii 355
c:
...
«I
I-
0 ,6, MW=120 000

CIJ
CIJ
350
o MW=500 800
a
«I
345 0 o MW=2 900000
340
10
Film Thickness (nm) 100
Figure 9.19 Glass transition temperatures of thin polystyrene films on a silicon
substrate, determined by measuring the thermal expansivity of the films using
ellipsometry. Data are shown for three molecular weights. (After Keddie, Jones and
Cory, 1994.)

which can offset the influence of the free surface in increasing the mobility
of nearby segments.
At present there is essentially no theory available to describe either
interface effects or finite size effects on glass transitions in these polymer thin
films; this reflects the poor level of fundamental understanding of the glass
transition phenomenon in general. A number of approaches emphasize the
idea of a length scale; of dynamic origin, which characterizes the size of the
region in a glass-forming liquid over which motion must be cooperative
(Adam and Gibbs, 1965; Donth, 1992). As the temperature decreases, it is
supposed that this dynamic correlation length increases and ultimately
diverges at an idealized glass transition. An experimental glass transition
arises because the system's relaxation time T depends on the length scale;
like T "" ;z where z is a (positive) dynamic exponent; as the temperature is
lowered the correlation length increases until the relaxation time is so large
that the system cannot equilibrate on experimental time scales, and an
experimental kinetic glass transition temperature emerges naturally. For
example, in a computer simulation of a two-dimensional glass in a finite
geometry with periodic boundary conditions (Ray and Binder, 1994), at a
given temperature the diffusion coefficient is larger as the system size gets
smaller. This is because the correlation length for the finite system cannot
446 Interfaces

exceed the system size, so in the finite system there is an upper bound on
the relaxation time and a consequent increase in the diffusion coefficient at
low temperature.
Of course, periodic boundary conditions do not correspond to an experi-
mentally realizable system; real confined fluids always have walls present. If
the motion of polymer segments near the wall is severely inhibited we would
expect the glass transition temperature to increase as the system size is
decreased (Sappelt and JackIe, 1993). Such an effect has been seen in a recent
computer simulation of a colloidal fluid confined between hard walls (Fehr
and Lowen, 1995). Experiments on confined small molecule and oligomeric
liquids are inconclusive; some show the expected increase in glass transition
temperature (Schuller et at., 1994) while others show a decrease in glass
transition temperature (Jackson and McKenna, 1991; Zhang, Liu and Jonas,
1992). The reasons for these current apparent contradictions remain unex-
plained, but there must be a suspicion that the details of the interaction
between the fluid and the wall must be important. The situation for high
polymers is even less clear because of the difficulty of creating experimental
samples with a well defined degree of confinement. Two situations that have
been studied are semicrystalline polymers, in which amorphous regions are
confined between lamellar crystals, and block copolymer microphases. For
example, in poly(ethylene terepthalate) a slight increase in the ~ of the
amorphous phase has been reported as the spacing between the crystalline
lamellae is decreased (Schick and Donth, 1991), while in styrene-butadiene-
styrene block copolymers a depression of the ~ of the styrene phase has
been observed (Bares, 1975). In the latter case, however, one needs to bear
in mind that in addition to the effect of confinement, a non-negligible
amount of the styrene is intimately mixed with butadiene segments in the
interfaces between the phases .
. Let us now return to the effect of the free surface on a glassy polymer.
Computer simulation shows that polymer segments near the surface of a
glass are more mobile than those in the bulk (Mansfield and Theodorou,
1991b); Figure 9.20 shows this near-surface mobility enhancement for both
the local motion of chain segments and the centre of mass motion of chains
in a molecular dynamics simulation of a glassy polymer. This suggests a
tempting interpretation for the observed depression of glass transition
temperature in thin films without strong interactions with a substrate.
Rather in analogy with the phenomenon of surface melting, one might
suppose that below the bulk glass transition temperature a region near the
surface might be effectively liquid-like, with the size of the region increasing
as the temperature was increased towards the bulk transition temperature.
This sort of surface liquid-like layer would lead to some interesting experi-
mental consequences; for example, if the extent of the liquid-like layer was
Chain conformation and dynamics in glassy polymers 447

4 i j

/ 26 A- 00
3.5

Free surface 19.5 - 26


3

2.5

1.5

0.5

o
o 10 20 30 40 50 60
Time (ps)

8 I j

6 \'26A--
5 Free surface
4

3
19.5-26A
2 13-19.5A
--"'. ·6.5-13A

0-6.5A

o
o 10 20 30 40 50 60
Time (ps)
Figure 9.20 Mean squared displacement of atoms (a) and chain centres of mass (b)
as a function of time in a molecular dynamics simulation of a thin film of glassy
polymer. The data are shown grouped in five regions of the film as labelled, with the
origin at the centre of the film. (After Mansfield and Theodorou, 1991 b.)
448 Interfaces

large enough to allow significant entanglement to take place if two glassy


polymers were brought into contact, then even below the bulk glass
transition temperature significant autoadhesion might be expected to occur.
In fact, experiments do show (Breach, Donald and Jones, unpublished data)
that interfacial fracture toughnesses of about 20 J m - 2 can be achieved for
autoadhesion of polystyrene 15° below the bulk glass transition tempera-
ture; it is difficult to understand how this kind of value could be achieved
without substantial interpenetration having taken place at the segment level.
The presence of a more mobile surface layer may also have implications for
the understanding of molecular mechanisms of crazing and craze break-
down, as discussed in Chapter 6. Another technologically important effect
that has been attributed to the presence of a surface layer of enhanced
mobility is the alignment of molecules at the surface of rubbed polyimides.
Alignment layers for liquid crystals in devices are created by a rather gentle
process of rubbing at a temperature well below the glass transition tempera-
ture of the polyimide; grazing angle X-ray diffraction reveals that this results
in alignment of the chains near the surface (Toney et aI., 1995).
Despite this circumstantial evidence for increased mobility near the
surface of a glassy polymer, direct evidence is much sparser. Patterns of
deformation obtained at the surface of polystyrene on interaction with an
atomic force microscope have been interpreted as suggesting that even at
room temperature the surface of polystyrene is effectively rubbery (Meyers,
DeKoven and Seitz, 1992). On the other hand, low energy positronium
annihilation lifetime spectroscopy provides a direct, surface sensitive
probe of free volume, allowing a direct measure of the glass transition
temperature of the top 10 nm of a polystyrene sample (Xie et al., 1995); no
depression of the glass transition temperature was observed. Until a direct,
surface sensitive probe of polymer chain mobility is available, this fascinat-
ing and important question must remain open.

REFERENCES
Adam, G. and Gibbs, J.H. (1965) J. Chern. Phys., 43, 139.
Affrossman, S., Hartshorne, M., Jerome, R. et al. (1993) Macromolecules, 26,6251.
Affrossman, S., Hartshorne, M., Kifi', T. et al. (1994) Macromolecules, 27, 1588.
Anastasiadis, S.P., Russell, T.P., Satija, S.K. and Majkrzak, C.F. (1990) J. Chern.
Phys., 92, 5677.
Bares, J. (1975) Macromolecules, 8, 244.
Baschnagel, J. and Binder, K. (1995) Macromolecules, 28,6808.
Bitsanis, I.A. and Hadzioannou, G. (1990) J. Chern. Phys., 92, 3827.
Brown, H.R. (1989) Macromolecules, 22,2859.
Brown, H.R. (1991a) Annu. Rev. Mater. Sci., 21, 463.
Brown, H.R. (1991b) Macromolecules, 24, 2752.
References 449

Brown, H.R. (1993) Macromolecules, 26, 1666.


Brown, H.R., Char, K., Deline, V.R. and Green, P.F. (1993) MllCromolecules, 26,
4155.
Char, K., Brown, H.R. and Deline, V.R. (1993) Macromolecules, 26. 4164.
Clarke, CJ., Jones, R.A.L., Edwards, J. and Penfold, J. (1995) ,Wacromolecules, 28,
2048.
Creton, e., Brown, H.R. and Deline, V.R. (1994) Macromolecules, 27. 1774.
Creton, e., Kramer, EJ., Hui, e.-Y. and Brown, H.R. (1992) Macromolecules, 25,
3075.
Dai, e.-A., Dair, BJ., Dai, K.H. et al. (1994) Phys. Rev. Lett., 73, 2472.
Donth, E.-J. (1992) Relaxation and Thermodynamics in Polymers. Glass Transition,
Akademie Verlag, Berlin.
Edwards, S.F. (1965) Proc. Phys. Soc. London, 85, 613.
Factor, B.J., Russell, T.P. and Toney, M.F. (1991) Phys. Rev. Lett.. 66, 1181.
Factor, BJ., Russell, T.P. and Toney, M.F. (1993) Macromolecule,. 26, 2847.
Factor, BJ., Russell, T.P. and Toney, M.F. (1994) Faraday Discuss., 98, 319.
Fehr, T. and Lowen, H. (1995) Phvs. Rev. E, 52, 4016.
Fernandez, M.L., Higgins, J.S., 'Penfold, J. and Shackleton, C. (1990) Polym.
Commun., 31, 127.
Fernandez, M.L., Higgins, J.S., Penfold, J. et al. (1988) Polymer, 29, 1923.
Fleischer, e.A., Morales, A.R. and Koberstein, J.T. (1994) Macromolecules, 27, 379.
Helfand, E. and Tagami, Y. (1971) J. Chem. Phys., 56,3592.
Hobbs, S.Y., Bopp, R.e. and Watkins, V.H. (1983) Polym. Eng. Sci., 23, 380.
Horn, R.G. and Israelachvili, J.N. (1988) Macromolecules, 21, 2836.
Israelachvili, J. (1991) I ntermolecular and Surj(lce Forces, Academic Press, San Diego.
Jackson, e.L. and McKenna, G.B. (1991) J. Non-Cryst. Solids, 131-133, 221.
Jones, R.A.L. (1993) in Polymer Surfaces and Interfaces II (ed, WJ. Feast, H.S.
Munroe and R.W. Richards), John Wiley, New York, p.71.
Jones, RAL., Kramer, E.J., Rafailovich. M.H. et al. (1989) Phys. ReI'. Lett., 62,280.
Jones, RAL., Kramer, EJ., Norton, L.J. et al. (1992) Macromolecules, 25, 2359.
Kanninen, M.F. (1973) Int. J. Fracture, 9,83.
Keddie, J.L., Jones, R.A.L. and Cory, R.A. (1994) Europhys. Lett .. 27, 59.
Keddie, J.L., Jones, RAL. and Cory, R.A. (1995) Faraday Discuss., 98, 219.
Kramer, EJ. (1994) Faraday Discuss., 98, 31.
Kumar, S.K., Vacate 11o, M. and Yoon, D.Y. (1990) Macromolecules, 23, 2189.
Mansfieid, K.F. and Theodorou, D.N. (1990) Macromolecules, 23. 4430.
Mansfield, K.F. and Theodorou, D.N. (1991a) Macromolecules, 24, 4295.
Mansfield, K.F. and Theodorou, D.N. (1991b) Macromolecules, 24,6283.
Meyers, G.F., DeKoven, B.M. and Seitz, J.T. (1992) Langmuir, 8, 2330.
Norton, L.J., Smigolova, V., Pralle, M.ll. et al. (1995) Macromolecules, 28, 1999.
Poser, e.I. and Sanchez, I.e. (1979) J. Colloid Interface Sci., 69, 539.
Ray, P. and Binder. K. (1994) Europhys. Lett.. 27, 53.
Reiter, G. (1993) Europhys. Lett.. 23, 579.
Reiter, G. (1994) Macromolecules, 27,3046.
Rowlinson, J.S. and Widom, B. (1982) Molecular Theory of Capillarity, Clarendon
Press, Oxford.
Russell, T.P. (1990) Mater. Sci. Rep., 5,171.
Russell, T.P., Hjelm, R.P. and Seeger, P.A. (1990) Macromolecules, 23, 890.
Russell, T.P., Anastasiadis, S.H., Menelle, A. et al. (1991) Macromolecules, 24, 1575.
Sappelt, D. and JackIe, J. (1993) 1. Phys. A: Math. Gen. Phys., 26, 7325.
Sauer, B.B. and Dee, G.T. (1994) J. Colloid Interface Sci., 162, 25.
Schick, e. and Donth, E. (1991) Phys. Scripta, 43, 423.
450 Interfaces

Schuller, J., Mel'nichenko, Y.B., Richert, R. and Fischer, E.W. (1994) Phys. Rev.
Lett., 73,2224.
Semenov, A.N. (1994) Macromolecules, 27, 2732.
Sha, Y., Hui, e.y. and Ruina, A. (1995) Macromolecules, 28, 2450.
Shull, K.R. (1991) J. Chern. Phys., 94, 5723.
Shull, K.R. (1992) in Physics of Polymer Surfaces and Interfaces (ed. I.e. Sanchez),
Butterworth-Heinemann, Boston, p. 203.
Shull, K.R. and Kramer, E.J. (1990) Macromolecules, 23, 4769. .
Shull, K.R., Mayes, A.M. and Russell, T.P. (1993) Macromolecules, 26,3929.
Shull, K.R., Kramer, E.J., Hadziioannou, G. and Tang, W. (1990) Macromolecules,
23,4780.
Smith, 1.W., Kramer, E.J. and Mills, PJ. (1994) J Polym. Sci. B: Polym. Phys., 32,
1731.
Smith, 1.W., Kramer, E.J., Xiao, F. et al. (1993) J. Mater. Sci., 28, 4234.
Stamm, M. and Schubert, D.W. (1995) Annu. Rev. Mater. Sci., 25, 325.
Theodorou, D.N. (1992), in Physics of Polymer Surfaces and Interfaces (ed. I.e.
Sanchez), Butterworth-Heinemann, Boston, p. 139.
Toney, M.F., Russell, T.P., Logan, 1.A. et al. (1995) Nature, 374, 709.
Washiyama, 1., Kramer, EJ., Creton, e.F. and Hui, e.-y. (1994) Macromolecules, 27,
2019.
Wool, R.P., Yuan, B.-L. and McGarel, OJ. (1989) Polym. Eng. Sci., 29, 1340.
Wu, W., v. Zanten, 1.H. and arts, WJ. (1995) Macromolecules, 28, 771.
Wu, W.L., arts, W.J., v. Zanten, J.H. and Fanconi, B.M. (1994) J. Polym. Sci. B:
Polym. Phys., 32, 2475.
Xie, L., DeMaggio, G.B., Frieze, W.E. et al. (1995) Phys. Rev. Lett., 74, 4947.
Yethiraj, A., Kumar, S., Hariharan, A. and Schweizer, K.S. (1994) J. Chern. Phys.,
100,4691.
Yeung, C., Balazs, A.e. and Jasnow, D. (1992) Macromolecules, 25, 1357.
Zhang, J., Liu, G. and Jonas, 1. (1992) J. Phys. Chern., 96, 3478.
Zhao, W., Zhao, X., Rafailovich, M.H. et al. (1993) Macromolecules, 26,561.
Morphology of block
copolymers 10
A.I Ryan and I. W Hamley

10.1 INTRODUCTION
Block copolymers are unique macromolecules that enable the structural and
processing properties of distinct polymers to be combined by linking
polymer chains in a variety of architectures. The block copolymer chapter
in the first edition of The Physics of Glassy Polymers (Folkes and Keller,
1973) was timely and signalled the importance of the subject. There have
been a number of advances in the intervening years, particularly in the area
of theory and morphology. The mechanical properties of solid polymers
were treated in detail in the first edition and this work will not be repeated
here. The phase behaviour of block copolymers in the melt has been the
subject of great interest lately, because a number of new morphologies have
been discovered which supplement the classical lamellar, hexagonal-packed
cylinder and body-centred cubic micelle phases that have been known for
some time. The thermodynamics of block copolymer melts prior to the
discovery of these complex phases was reviewed by Bates and Fredrickson
(1990), who have also recently reviewed the dynamics of block copolymer
melts (Fredrickson and Bates, 1996).
In this chapter we discuss the physics of block copolymers containing
a glassy component. Examples of glass-forming components of block
copolymers for which we will present results include polystyrene (PS;
~ '" 100°C), poly(methyl methacrylate) (~ '" 120°C), poly(vinyl pyridine)
(PVP; Tg '" 100 C) (Gedcle, 1995), and poly(vinyl cyclohexane) (~ '" 148°C)
D

(Gehlsen and Bates, 1993).


The architecture of block copolymers can be controlled by the synthesis
procedure, and it is possible to prepare diblock, triblock, multi block, star
and graft copolymers. These are illustrated in Figure 10.1. In this chapter

R. N. Haward et al. (eds.), The Physics of Glassy Polymers


© Chapman & Hall 1997
452 Morphology of block copolymers

AS cyclIC d

ASAU Od<

(AB)n 51a'

ABCuobIock
ABC 51ar

AB 'n mulhb

Figure 10.1 Block copolymer architedures.

we focus on linear block copolymers, i.e. diblocks, triblocks and linear


multiblocks. The method of choice for synthesis of diblock and triblock
copolymers is living polymerization of anionically reactive polymers, which
ensures a narrow molecular weight distribution. The first anionic block
copolymerizations were achieved as early as the 1950s (Szwarc, Levy and
Milkovich, 1956). Industrially, block copolymers such as triblock co-
polymers of polystyrene-polybutadiene- polystyrene (SBS), which are used
as synthetic rubbers, are prepared by anionic polymerization (Aggarwal,
1985). Multiblock copolymers can also be synthesized anionically (Smith et
al., 1994) but are more generally prepared by condensation polymerization
(Abouzahr and Wilkes, 1985).
Microphase separation theory 453

10.2 MICROPHASE SEPARATION THEORY


In the melt, block copolymers can self-assemble into a variety of ordered
structures via the process of microphase separation. Microphase separation
is driven by the enthalpy of demixing of the constituent components of the
block copolymers, whilst macro phase separation is prevented by the chemi-
cal connectivity of the blocks. This enthalpy is proportional to the Flory-
Huggins segmental interaction parameter X, which is found to be inversely
proportional to temperature and is usually parameterized as X = AfT + B
where A, B are system-dependent constants and T is the temperature.
Microphase separation leads to ordered structures with periods of several
Rg, where Rg is the copolymer radius of gyration, and the entropic penalty
associated with the chain stretching is proportional to the degree of
polmerization N. The product XN that expresses the enthalpic-entropic
balance is then used to parameterize block copolymer phase behaviour,
along with the composition of the copolymer. For a diblock copolymer, the
volume fraction of one component, f, controls which ordered structures are
accessed beneath the order-disorder transition (ODT).
Near the ODT, the composition profile of ordered microstructures is
approximately sinusoidal. The phase behaviour in this regime, where the
blocks are weakly segregated, can then be modelled using Landau theory,
where the mean field free energy is expanded with reference to the average
composition profile. The order parameter for A-B block copolymers may
be defined as (Leibler, 1980)

where t{t(r) is the local density of monomer A and f is the average density.
The phase diagram for weakly segregated diblocks was first computed
within the Landau mean field approximation by Leibler (1980). The
resulting phase diagram for diblock copolymers is shown in Figure 10.2. The
theory predicts that micro phase separation occurs to a body-centred cubic
structure for all compositions except f = i, where a direct second-order
transition to a lamellar structure is predicted. First-order transitions to
hexagonal-packed cylinder (hex.) and lamellar (lam.) phases are expected on
further lowering the temperature for asymmetric diblocks. Leibler's theory
was extended to triblock and (A-B)m star copolymers by Mayes and Olvera
de la Cruz (1989), and to alternating multiblocks and polygraft copolymers
by Dobrynin and Erukhimovich (1993). Order-disorder transitions and
spinodals were computed for linear multi block copolymers by Fredrickson,
Milner and Leibler (1992).
454 Morphology of block copolymers

(0) (b)

22 22

18 18
z
><

14 14

Disordered Disordered
100 0.2 0.4 0.6 0.8 100 0.2 0.4 0.6 0.8 1
(a) f (b) f

100 Qlm3m

80
H L H
xN 60

40

20

DIS
a ~~~~~~~~~~~~~

a 0.2 0.4 0.6 O.B


(c) f
Figure 10.2 Theoretical block copolymer phase diagrams in the weak segregation
limit. (a) Mean field theory (Leibler, 1980). (b) Composition fluctuation theory
(Fredrickson and Helfand, 1987) for a reduced degree of polymerization,
ill = a 6 v- 2 Nwhere a and vare the statistical segment length and volume respectively
(Bates, Rosedale and Fredrickson, 1990). [(a) and (b) reproduced from 1. Chern. Phys.,
with permissionj. (c) SCFT phase diagram for a symmetric block copolymer illustra-
ting the regions of /aJd cubic morphology, between the lamellar and hexagonal
phases. (Reproduced from Matsen and Bates, 1996, with permission.)
Microphase separation theory 455

As noted by Leibler (1980), allowance for composition fluctuations


changes the mean field prediction of a second-order phase transition for a
symmetric diblock to a first-order transition. Fredrickson and Helfand
(1987) confirmed this, and also computed phase diagrams for asymmetric
diblocks using Brazovskii's theory (1975) for weakly ordered fluctuating
systems. An example of such a phase diagram is also shown in Figure 10.2.
In contrast to the mean field prediction, direct transitions to the hexagonal-
packed cylinder or lamellar phases are possible for asymmetric diblocks.
The region of stability of the body-centred cubic phase is moved to
compositions away from f = l The Fredrickson-Helfand theory also
predicts that the order-disorder transition occurs at a larger value of XN
than predicted by mean field theory, the correction for a symmetric diblock
scaling as N- 1 / 3 . The first experimental confirmation of the importance of
composition fluctuations in a block copolymer melt was reported by Bates
et al. (1988), and numerous further studies have confirmed many of the
theoretical predictions.
The theory for strongly segregated block copolymers was developed
before Leibler's theory by Helfand and coworker (Helfand, 1975; Helfand
and Wasserman, 1976, 1982). Self-consistent field theory was used to
calculate the configurational statistics of block copolymer chains (Helfand,
1975). The theory was simplified by the introduction of the narrow inter-
phase approximation, which assumes that the boundary between A and
B domains is narrow compared to the domain width (Helfand and
Wasserman, 1976). Phase boundaries are predicted to depend only on
copolymer composition f (Helfand and Wasserman, 1982) and experimen-
tal work on strongly segregated block copolymers supports this prediction
(Hasegawa et al., 1987). In this latter work, a novel 'ordered bicontinuous
double diamond (OBDD)' structure was also identified, which is not
predicted by strong segregation theory (Olmsted and Milner, 1994;
Likhtman and Semenov, 1994). We will return to a discussion of the correct
identification of bicontinuous cubic phases in block copolymers in section
10.4.
The weak and strong segregation limit theories for block copolymer melts
have recently been unified by Matsen and coworkers (Matsen and Schick,
1994; Matsen and Bates, 1996). This approach involves numerical solution
of self-consistent field equations, without approximations such as the
narrow interphase approximation or unit cell approximations. However,
calculation of a complete phase diagram is computationally intensive. This
theory predicts an J a3d phase to be stable in the weak segregation regime,
between the classical lam. and hex. phases predicted by Leibler (1980) as
illustrated in Figure 1O.2c (Matsen and Bates, 1996). A bicontinuous 'gyroid'
morphology belonging to the Ja3d space group was recently observed in this
456 Morphology of block copolymers

region of the phase diagram for diblocks (Forster et at., 1994; Hajduk et at.,
1994) although the stability window for polystyrene-polyisoprene (PS-PI)
diblocks extends to the ODT (Hajduk et aI., 1994; Khandpur et aI., 1995).
In contrast, mean field theory suggests a triple point for IiJd phase
boundaries above the ODT (Matsen and Schick, 1994), although when
allowance is made for fluctuations, a direct transition is predicted (Podneks
and Hamley, 1996). The hexagonal-modulated lamellar and hexagonal
perforated layer structures recently observed for diblocks near the ODT
(Disko et aI., 1993; Hamley et at., 1993, 1994) have not been accounted for
theoretically, although some attempts have been made (Fredrickson, 1991;
Olvera de la Cruz, Mayes and Swift, 1992; Hamley and Bates, 1994).
A phenomenological theory for the glass transition in diblock copolymers
was recently presented by Dobrynin (1995). It was shown that the interac-
tion between composition fluctuations, due to the loss of mobility of A
monomers, results in the elimination of microphase separation. Instead,
there is predicted to be a third-order phase transition from the disordered
phase into a glassy disordered phase. The calculated phase diagram also
contains a narrow region of a glassy lamellar phase, between the microphase
separated lamellar and glassy disordered phases (Dobrynin, 1995).

10.3 TECHNIQUES USED TO STUDY MORPHOLOGY


The most direct method for the investigation of block copolymer morphol-
ogy is transmission electron microscopy (TEM). In this technique a thin
section of block copolymer or a solvent cast film is first microtomed. For
soft samples, ultrathin sections are obtained at low temperatures (typically
-100°C), via cryo-ultramicrotomy. Contrast between rubbery and glassy
components in the section is then achieved by exposing the sample to a
staining substance which will enhance the contrast between the microphases.
For block copolymers containing dienes, osmium tetraoxide vapour, which
selectively stains the rubbery component (Kato, 1965, 1967), is used. In
polyolefin diblock copolymers containing polyethylene, it has recently been
shown that ruthenium tetraoxide can be used as a staining agent, and the
amorphous component [e.g. polyethylethylene or poly(ethylene-propylene)]
is selectively stained due to reduced dilfusivity of Ru0 4 in the semicrystal-
line microdomains (Khandpur, Macosko and Bates, 1995). In block
copolymers containing polyvinylpyridine, methyl iodide is used as a stain as
the iodine forms a complex with the pyridine ring (Schulz et aI., 1996). The
disadvantages of TEM are that misidentification of morphology is possible
based on inspection of a projection of only a small region of the sample. For
example, the 'OBDD' cubic phase was identified in star block copolymers
via TEM images showing a 'wagon wheel' structure (Thomas et at., 1986;
Techniques used to study morphology 457

Alward et al., 1986). Subsequent work showed that this projection could
also be obtained from the 'gyroid' bicontinuous structure, and this latter
morphology was confirmed by small angle X-ray scattering (Forster et aI.,
1994; Hajduk et al., 1994).
Small angle X-ray and neutron scattering are ideal for investigation of
block copolymer morphology, because these experiments probe length
scales typical of those characteristic of block copolymer microstructures, i.e.
1-100 nm. In contrast to TEM, the structure of the sample averaged over
the macroscopic size of the beam is probed. Scattering data are presented
either as two dimensional maps of iso-intensity contours as a function of the
scattering vector q or as linear plots of radially integrated intensity, I(q),
versus q where the scattering vector is defined

4n .
Iql = q = -ysmO

where A is the wavelength of the incident radiation and 0 is half the


scattering angle. q has dimensions of reciprocal length and allows data from
a variety of sources (and wavelengths) to be compared. The relative
positions of a sufficient number of reflections arising from microstructural
periodicities enable unambiguous identification of morphology. Further
information can be obtained by preparing oriented specimens, and obtain-
ing diffraction patterns for different orientations. For example, in an
oriented lamellar phase with the beam incident parallel to the layers, Bragg
reflections at q*, 2q*, 3q*, ... , where q* is the position of the first-order
reflection, are observed along a direction parallel to the layer normal.
Small angle X-ray scattering (SAXS) is appropriate where the electron
density contrast between blocks is sufficient for the polymer to diffract
X-rays (Balta-Calleja and Vonk, 1989). This is often possible with an
intense source of X-rays, such as a rotating anode generator or a synchro-
tron source. Use of the latter enables in situ studies of structure development
and kinetics (Singh et al., 1993; Stiihn, Vilesov and Zachmann, 1994;
Bras et aI., 1995; Rangarajan et al., 1995; Ryan et al., 1995). Small angle
neutron scattering (SANS) is valuable for studies of polymer structure
(Higgins and Benoit, 1994) because of the opportunity for contrast variation
via isotope labelling. Typically hydrogen atoms are selectively replaced
by deuterium; this changes the scattering contrast and can be used to
obtain local information on chain conformation in block copolymer melts
(Hadziioannou et al., 1982) or intramicellar structure in block copolymer
solutions (Mortensen and Pedersen, 1993), for example. Unfortunately, it
has been shown that deuterium labelling can change the thermodynamic
interactions in polymer blends (Bates, Keith and McWhan, 1987), and this
458 Morphology of block copolymers

is also likely to be the case for block copolymers. Neutron scattering has
also been extensively used to enhance the scattering contrast in polyolefin
diblock copolymers by Bates and coworkers (Bates et al., 1988; Almdal et
al., 1990; Bates, Rosedale and Fredrickson, 1990; Koppi, Tirrell and Bates,
1993).
The surface morphology of block copolymers can be investigated using
atomic force microscopy (AFM). The advantage of this technique is that
special preparation techniques (such as staining) are not required to obtain
contrast between blocks. Island and hole formation in lamellar diblock
copolymers was observed using AFM in thin block copolymer films, when
the sample thickness is incommensurate with the layer spacing (Coulon et
al., 1990; Collin et al., 1992). The terracing of droplets of spun-cast block
copolymers was also observed using this technique (Ausserre, Raghunathan
and Maaloum, 1993; Carvalho and Thomas, 1994). A flip in orientation of
hexagonal-packed cyclinders from parallel to perpendicular with respect to
the substrate was observed using tapping mode AFM in an SBS triblock
was observed by van Dijk and van den Berg (1995).
The structure of block copolymer films normal to the surface has been
investigated using neutron and X-ray reflectivity. The early work was
reviewed by Russell (1990). Recent examples from other groups include a
study of looping in a lamellar triblock copolymer film (de Jeu et al., 1993),
entropy-driven surface segregation in block copolymer melts (Sikka et al.,
1993) and surface-induced ordering in asymmetric block copolymers form-
ing a hexagonal phase (Liu et aI., 1994). In neutron reflectivity, one of the
components is deuterated to enhance the scattering contrast between
different domains, for example lamellar domains which usually form parallel
to the substrate in thin spun-cast films.

10.4 MORPHOLOGY

10.4.1 DffiLOCKS

(a) Classical Morphologies


For a block copolymer to be useful as a thermoplastic elastomer, the
ordered structure must act as a thermally labile crosslink. Thermoplastic
elastomers have a soft block which has a subambient 1'g and a hard block
which has a 1'g or Tm well above ambient. Generally the soft block is the
majority component so the morphology comprises a matrix of rubbery
chains with domains of glassy or crystalline hard block which act as physical
crosslinks and reinforcing filler. The first commercially available thermo-
plastic elastomers were based on SBS triblock copolymers and contained
15-40% polystyrene, consequently much experimental work was (and still
Morphology 459

is) carried out on these readily available polymers. They have spherical or
hex. morphologies of PS spheres or rods embedded in a polybutadiene (PB)
matrix, and good examples of micrographs of these polymers are contained
in the first edition of this book. The literature on block copolymer
microstructure is vast and we do not aim to be comprehensive. We will limit
ourselves to the equilibrium (or at least near-equilibrium) morphologies.
For example, a wide range of morphology may be found from one block
copolymer if solvent casting techniques are adopted because preferential
solvents allow metastable structures to be trapped by vitrification. The
reader is pointed to the review by Folkes (1985) for a discussion of this
subject.

ImJm laJd HPL LAM

Cylloders Bkloo... ouous Lamelloe

40 'nJmi I HEX
LAM

, \ i I

\., \!\:
" : l "
i
i
\ \: i
\ . I
\ /
\ I
\ /
"., .//

10
.'.~
'_._ _.-"
...
~./

Oi ordered

o~~--~~--~--~~--~~--~~
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Figure 10.3 Experimental phase diagram for low molar mass PS - PI diblock
copolymers. (Reproduced from Khandpur et aI. , 1995, with permission.)
460 Morphology of block copolymers

(a)

10 ' •
1 •

T" 015

.•
0. 1 Iq ~

0.05
2 J ..\

-.Jl-A ~ ., wi 0
0 0.05 0. 1 O. IS O.:! 0 .25

(b) q/

Figure 10.4 (a) The lamellar microstructure of a PS- PI diblock with fps = 0.64 and
Mn = 39 k shown by TEM where the PI is stained dark. (Reproduced from Khandpur
et al.. 1995. with permission.) (b) SAXS powder pattern from the same structure.
Morphology 461

The equilibrium phase diagram of the low molar mass PS-PI system is
shown in Figure 10.3 (Khandpur et aI., 1995) and we will use it as a guide
to the various microstructures (although some of the TEM, SAXS and
SANS examples will be from elsewhere). In order to determine this phase
diagram, a great deal of experimental work was required. First, the tempera-
ture dependence of X was determined by measuring the ODT of a series of
lamellar polymers by rheology. The microstructure of the copolymers was
determined by a combination of SAXS, SANS and TEM. The SAXS and
SANS experiments could be conducted on samples at elevated temperatures,
but for TEM imaging the material had to be annealed in the melt, at a given
temperature, and then quenched so that the S-domains became glassy prior
to sections being cut by cryo-ultramicrotome.
The most simple morphology is lamellar and the electron micrograph of
a polydomain, lamellar, PS-PI (fps = 0.64) block copolymer is shown in
Figure 10Aa (Khandpur et at., 1995). That the lamellae are in domains with
long range order is obvious from the micrograph. The SAXS pattern, which
takes an ensemble average over the scattering volume ("" 1 mm 3 ), is iso-
tropic and comprises a series of Debye-Scherrer rings in the position ratio
1 : 2: 3 : 4. Figure WAb shows the powder SAXS pattern and the long period
(d = 21t/q*) is seen to be 233 A (23.31 nm). The individual lamellar thickness
can be readily calculated from the volume fractions.
Lamellar microstructures can also be seen in Figure 10.5a, which shows
a TEM micrograph of a polydomain in a PS-PVP system (fps = 004) that
has been annealed in the melt. The micrograph contains a number of
disclinations emphasizing the relationship between the structures formed in
block copolymers and liquid crystals. The SANS patterns in Figure 10.5b
show the scattering from a shear oriented sample of this polymer in the three
orthogonal orientations defined by the shear in Figure 1O.Sc. When the
neutron beam is orthogonal to the lamellae there is very little scattering, but
when the neutron beam is parallel to the lamellae there are strongly oriented
SANS patterns with relections at q* and 2q* from the grating-like structure
of the lamellar polymer.
It is possible to assess the block copolymer microstructure from just the
SAXS powder pattern if enough reflections are sampled for the pattern to
be indexed unambiguously, but this is often the exception rather than the
rule. The combination of TEM and scattering from oriented samples makes
structure solution relatively simple!
The first example of the use of scattering from a shear oriented block
copolymer was given by Keller, Pedemonte and Willmouth (1970) and
reported in the block copolymer chapter in the first edition of this book
(Folkes and Keller, 1973). Their excellent micrographs from the hexagonal
462 Morphology of block copolymers

(a)

q' t
01

01» I
'VA-' 0
U o
~I»

~I
~IO ~I» .0 oas 010
(b)
q/A-' qy

(c)

Figure 10.5 (a) TEM micrograph from a PS - PVP diblock (fps = 0.40. Nn = 209).
annealed at 1400 ( for 6 h then quenched in liquid nitrogen. Iodine was used to
selectively stain the PVP. (b) SANS pattern from the same phase after preshearing at
140°C. (Reproduced from Schulz et al.. 1996. with permission.) (c) Schematic
illustration of the morphology showing the shear direction (double-headed arrow)
and the coordinate system for the neutron scattering in (b).
Morphology 463

(a)

~
106

~
.'
'"
::>
>.
10'

~ \ '';: 3q.
:e.
r
10'
,
"
7q·
E
\1; ~
1000
~\
100

0 O.OS 0.1 0.15


(b) I
qI

Figure 10.6 (a) TEM micrographs from a hexagonal morphology in Kraton TR 1102,
a styrene- butadiene- styrene block copolymer with 4>ps - 0.24 (Source: Folkes and
Keller, 1973.) (b) One-dimensional SAXS plot from a sample of Kraton TR1102
(Source: A.J . Ryan, unpublished data.)

phase of Kraton TRll02 (PS- PI - PS <pps = 0.24) are reproduced here (in
Figures 1O.6a and 1O.6b). When viewed end-on the hexagonal arrangement
of the rods is obvious; however, when viewed from the side the material
takes on the appearance of a lamellar structure. The powder SAXS pattern
from the same grade of polymer (Figure 10.6b) shows the classical higher
order reflections at q*, 3 1/ 2 q* , 7 1 / 2 q* and 9 1 / 2 q*. There should also be a
464 Morphology of block copolymers

(a)

(l10-0-------,

'VA" 0.00-

o 0 0

(b) -010 -0 OS 0 O~ 010

q/A"

#~
;;<$'
•••
f-----~

(c)
••
•••
Figure 10.7 (a) TEM for a PS - PVP diblock (fps = 0.35, Nn = 203) obtained after
annealing at 1400 ( for more than 6 h. The PVP-rich region appears dark. (Repro-
duced from Schulz et al., 1996, with permission .) (b) SANS patterns from a
presheared specimen under the same conditions. (Reproduced from Schulz et al.,
1996, with permission.) (c) Schematic illustration of the morphology showing the
shear direction (double-headed arrow) and the coordinate system for the neutron
scattering in (b).
Morphology 465

80.0 ""

1M

10.0 ""

S.O _

0.0 ""

1M
(b)

Figure 10 .8 AFM micrographs of thin films of a Kraton D-11 02 triblock copolymer


with <Pps ~ 0.24. (b) An enlargement of the square region in (a), showing the
coexistence of regions with parallel and perpendicular cylinders of PS. (Reproduced
from van Dijk and van den Berg, 1995, with permission .)
466 Morphology of block copolymers

IO'~
~ 1 10

10"
1 1
1 ;l1
10'· e
¢
- 101 8
cr: 10· >-
I-
§
1, 0.01 ~
'0
if
~
10'
0.001

10'

10'~
(a) 0 0.05 0.1 0. 15 0.210 '

(b)

(d)
Morphology 467

reflection at 4 1/ 2 q* but this is coincident with a minimum in the form factor


of cylinders and is systematically absent. The combination of TEM and
SAXS is needed to confirm the structure. It should be noted that this is a
commercial thermoplastic elastomer that has been in production for more
than 30 years. The length scales from TEM and SAXS that correspond to
a system of rods of diameter ~ 9.6 nm which are separated by ~ 28 nm, are
in excellent agreement, even though the SAXS and TEM are taken from
batches of polymer produced more than 20 years apart.
The polydomain structure which exists in unoriented hexagonal block
copolymers may be seen in Figure 10.7a where the TEM of a PS-PVP
copolymer (fps = 0.70; Schulz et ai., 1996) shows two grains which are
imaged parallel and perpendicular to the rods. The long range order that
can be induced by shearing is obvious in the SANS patterns in Figure 10.7b.
The central pattern shows the hexagonal symmetry obtained when the
neutron beam is parallel to the rods (and the shear direction), the two outer
patterns show only the interplanar separation of the rods. Figure 1O.7c
shows the direction in which the shear orientation was applied and the
coordinates of the neutron beam.
An AFM study of the behaviour of thin films of a hexagonal block
copolymer has been published. A change in orientation of hexagonal-packed
cylinders from parallel to perpendicular with respect to the substrate was
observed in SBS triblock by van Dijk and van den Berg (1995). The low
resolution AFM image in Figure 10.8a shows a collection of domains in a
thin film of a PS- PI - PS triblock. A higher resolution image of the area in
the box in Figure 1O.8a is presented in Figure 10.8b, where it is now obvious
that the sample comprises regions where the PS rods are parallel to the film
surface and holes where the PS rods are orthogonal to the film surface.
Unlike the behaviour observed in lamellar block copolymers (Ausserre,
Raghunathan and Maaloum, 1993; Carvalho and Thomas, 1994), there
appears to be no terracing at the polymer-air interface.
Spherical domains are found in diblock and triblock copolymers that are
asymmetric. The most commonly observed structure is a disordered liquid

Figure 10.9 (a) SAXS pattern for an styrene-isoprene-styrene triblock TR 1107,


</>ps - 0.18, in a phase with disordered spheres, compared to the scattering from a
single 100 A (10 nm) sphere (Source: A.J. Ryan, unpublished data.) (b) A TEM
micrograph from an 12 S Y-shaped polymer with fps = 0.17, Mw = 100 k oriented by
compression showing fourfold symmetry due to a [100] projection and (c) three-
fold hexagonal symmetry due to a [111] projection of the cubiC array. (d) SAXS
pattern from the 12S Y-shaped polymer with fps = 0.17. The pattern shows fourfold
symmetry from an approximate [100] zone axis of a Bee structure ((b)-(d) Repro-
duced from Pochan et at., 1996, with permission.)
468 Morphology of block copolymers

of spheres, and a typical scattering pattern is given in Figure 10.9a for


Kraton TR1102. Micrographs (Folkes, 1985) often show spherical particles
with no long range order and the scattering pattern has a peak due to the
interparticle structure factor and a series of minima due to the spherical
shape factor. A good example of the microstructure observed in the TEM
for a well ordered spherical structure is given in Figures 1O.9b and 10.9c
which shows the fourfold symmetry due to a [100] projection and the
threefold hexagonal symmetry due to a [111] projection of the cubic array.
This TEM is obtained for a Y-shaped polymer that has two PI arms and a
PS arm (Pochan et al., 1996) with ips = 0.17. Long range order may be
introduced in this material by solvent casting and in Figure 10.9d the
two-dimensional SAXS pattern shows seven orders of reflection, indicative
of such long range order, with the fourfold symmetry of an approximate
[100] zone axis.

(b) Complex Phases - Gyroid and Perforated Layers


In recent years so-called complex phases, such as the bicontinuous gyroid
and perforated layer structures have been identified. Their history is a classic
example of the danger of interpreting structures on the basis of TEM
alone.
There have been many misassignments of structures as being ordered
bicontinuous double diamond (OBDD, space group Pn3m) which is based
on a tetrahedral arrangement of channels when in fact they were Gyroid
(Hajduk et al., 1995) which has IaJd symmetry and is based on a tripod
arrangement of channels. For example, in 1986 the OBDD bicontinuous
cubic phase was identified in star block copolymers via TEM images
showing a 'wagon wheel' structure (Alward et al., 1986; Thomas et al., 1986;
Hasegawa et al., 1987). This type of TEM pattern had been previously
published by Aggarwal (1985) for a diblock copolymer and was subsequent-
ly observed in a whole series of block copolymer-homopolymer blends
(Winey, Thomas and Fetters, 1991, 1992a,b).
Figure 10.10 shows the threefold (wagon wheel) and fpurfold stuctures
obtained in the TEM (Forster et al.) for the bicontinuous cubic structure in
a PS-PI diblock (fps = 0.39). TEM micrographs like these have been
modelled as if they were OBDD (space group Pn3m; Thomas and Lescanec,
1994) and the reason is that the density projections of [111] and [100] for
the IaJd and the Pn3m structures are very similar.
The SANS pattern (Figure 10.11a) found for the shear-oriented sample of
PS-PI (fps = 0.39) has ten spots consistent with the {211} and {220}
reflections from the I aJd structure and their angular spread indicates that
Morphology 469

(a)

(b)

Figure 10.10 Two projections of the la3d phase obtained by TEM on a (PS - PI)
diblock (fps = 0.39, Mn = 32 k), selectively stained with OS04' (a) The fourfold
symmetry due to a [100] projection and (b) threefold hexagonal symmetry due a
[111] projection of the cubic array which shows the distinctive 'wagon wheel'
morphology. (Reproduced from Forster et aI., 1994, with permission.)
470 Morphology of block copolymers

oG
~
0 0
QQO
(a)

1000.0

100.0

10.0
211

1.0

!l
'a
:>

i
~
no
§
0
0.00 0.02 0.04 0 .06 0.08 0. 10 0. 12 0. 14
(b) q/ A1

0.1

0.08 44. /83 , 14


,.,
I"
0.08 '2'
.:c 420
"a-
0 . 04 220

0 . 02
2' ,
(c)
0
0 2 3 4 6 e 7 8
(h' + ~ +t')' "
Morphology 471

the [111] direction is coincident with the shear. This structure assigment is
confirmed by the one-dimensional SAXS data in Figure 1O.l1b which show
> 10 reflections which can be indexed to the Ia3d spacegroup by plotting
the square root of the sums of the squares of the Miller indices (hkl)
versus the peak position. Figure 10.11c illustrates the indexing of this cubic
pattern through a plot of (h 2 + k 2 + 12)1/2 versus q, which is linear.
Once the space group of the bicontinuous cubic structure was established
then TEM images could be used to their best. Figures HU2a and 10.l2b
show the TEM of the [111] projection and the [100] projection of the Ia3d
gyroid phase; the insets on the left and in the centre show the optical
transform of the image and the indexed diffraction pattern, while simula-
tions of the TEM projections are given on the right (Hajduk et al., 1995).
Thus the structure responsible for the 'wagon wheel' TEM micrographs was
resolved to be the gyroid structure and a three-dimensional representation
of the unit cell is given in Figure 10.12c. The three-dimensional structure of
the gyroid has recently been imaged by tomographic reconstruction from a
tilt series in the TEM by Spontak et al. (1996).
The existence of a second class of complex phases, the modulated and
perforated layers, has largely been explored by Bates and coworkers
(Hamley et ai., 1993, 1994; Forster et ai., 1994; Khandpur et al., 1995; Schulz
et ai., 1996), where SANS and TEM have been used to investigate shear
oriented structures. SANS analysis shows strong layer spacings, with higher
order reflections, and weaker off-meridional reflections from the layer
perforations.
A typical SANS pattern from a hexagonal perforated layer (HPL) phase
is shown in Figure 1O.13b for a PS- PVP (fps = 0.65) diblock which has all
the typical features of a perforated layer structure (Schulz et ai., 1996). The
perforations in the lamellar structure can be seen in the TEM micrograph
in Figure 10.13a. In order to obtain unambiguous scattering patterns, shear
oriented samples must be used because radially averaging the scattering,
either numerically or by using a polydomain sample, removes the spatial
information. Typically the inter-perforation spacing and lamellar length
scales are similar but not equal (Hamley et ai., 1994) and the second-
order reflection in the phase assigned as HPL often occurs at 1.88 q*, in
contrast to 2 q* for lamellar systems (Hamley et al., 1994; Khandpur et ai.,
1995).

Figure 10.11 (a) SANS pattern from a PS-PI diblock (fps = 0.39 and Mn = 32 k) in
the la3d phase. The polymer was shear oriented for 10 min at a shear rate 2.25 S-1
at 145°( and then heated. (Reproduced from Forster et aI., 1994, with permission.)
(b) SAXS powder pattern from the same phase. (c) Indexing of the peak positions to
the cubic space group (Source: I.w. Hamley and A.J. Ryan, unpublished data, 1993).
472 Morphology of block copolymers

(a)

(b)

(c)
Morphology 473

The orientation and arrangement of the perforations can be imaged


directly in PS-PI diblocks (Khandpur et aI., 1995) and Figure 10.14
illustrates the hexagonal nature of the PI perforations through a PS lamella
in a PS-PI diblock with ips = 0.68. The perforated layer structures have
been studied theoretically (Frederickson, 1991; Hamley and Bates, 1994;
Matsen and Schick, 1994) but have never been found to be the lowest energy
morphology. Recent, unpublished experimental work indicates that the
HPL is a metastable state stabilized in some systems by shear.
The versatility of diblock copolymers, in terms of the types of interactions
that can be introduced, is demonstrated by the tetragonal morphology
observed by TEM in a poly[2-(3-cholsteryl-oxycarbonyloxy)ethyl methac-
rylate]-block-styrene-block-[2-(3-cholsteryl-oxycarbon yloxy)ethyl meth-
acrylate] triblock (Fischer, 1994). Such tetragonal structures are uncommon
in block copolymers and only occur under special circumstances for
example block copolymers containing liquid crystal components or for
polymers with an unusual molecular architecture such as ABC triblock
copolymers.

(c) Dislocations and Grain Boundaries

That block copolymers form structures with polycrystalline order is well


established. Indeed, the subject is now so mature that the nature of grain
boundaries and dislocations is now being investigated, and non-equilibrium
metastable structures can be studied. TEM is the technique of choice for the
study of non-equilibrium structures as the local structure can be directly
imaged from a small volume. Small angle scattering techniques are inappro-
priate here as they take an ensemble average over a large volume.
There are two kinds of grain boundaries found in lamellar block
copolymers (Gido et al., 1993a; Gido and Thomas 1994a,b): twist grain
boundaries and tilt grain boundaries. The geometry of grain boundaries is
illustrated in Figure 10.15 (Gido et al., 1993a). The reference state is an
aligned bicrystal with no boundary. A tilt boundary is produced by rotating
the grains with respect to one another about an axis which is parallel to the

Figure 10.12 TEM images from a PS-PI diblock (fps = 0.33, Mn = 27 k). (a) Approxi-
mate threefold symmetry is evident from the [111] projection of the cubic structure.
Inset are the FFT and indexed diffraction pattern. The inset on the right is a simulated
[111] projection of the constant thickness gyroid structure. (b) Another region of the
same sample, showing fourfold symmetry, with a diffraction pattern, its indexation
and a simulated [1 DO] projection of the gyroid structure inset. The minority
component (PS) appears light in the micrographs and simulations (c) A three-
dimensional representation of the unit cell of the gyroid. (Reproduced from Hajduk
et at., 1995, with permission.)
474 Morphology of block copolymers

(a)

0
0.

q/l-l 0.

.0.

.0.1
(b) .0.10 .o.OS .0.00 0.05 0.10
q/l-l
Figure 10.13 (a) TEM image from a PS-PVP diblock with fps = 0.65 and Nn = 196
after disordering. quenching and annealing at 1400 ( for 18 h. The dark-stained
region corresponds to PVP. (b) Two-dimensional SANS pattern from a presheared
sample of the same diblock in the same phase (Reproduced from Schulz et al.. 1996.
with permission .)

plane of the grain boundary (Figure lO.lSb). A twist boundary is produced


by rotating the grains with respect to one another about an axis which is
perpendicular to the plane of the grain boundary (Figure lO.lSc).
The geometry of tilt grain boundaries is shown in Figure 10.16 (Gido and
Thomas, 1994a) where the distinction is made between symmetric tilt
boundaries, un symmetric tilt boundaries and the special case of pure
Morphology 475

Figure 10.14 TEM image of the HPL phase in a PS - PI diblock with frs = 0.68 and
Mn = 40 k after annealing for 6 h at 145°( followed by quenching in liquid nitrogen.
The PI appears dark. The inset on the left shows a section with the sample parallel
to the microscope stage, the one on the right shows a magnification of the
perpendicular orientation shown in the main figure. (Reproduced from Khandpur et
al., 1995, with permission.)

T-geometry. Generally three types of tilt boundary are observed. Chevron


grain boundaries occur at low degrees of tilt where the lamellae bend
continuously. Omega grain boundaries occur at higher degrees of tilt where
the lamellae bend continuously; they are a variant of the chevron grain
boundary in which one lamella in the boundary region approximates to a
Greek capital n as an overlayer to the other lamellae, and which has a
hemispherical cap. There is a discontinuous grain boundary, called the
T-junction, where lamellar planes on one side of the boundary terminate.
A TEM micrograph from a styrene- butadiene block copolymer with
ips = 0.5, which exhibits omega, chevron and T -junction grain boundaries
is shown in Figure 10.17. In the centre of the micrograph is a 125° symmetric
tilt boundary. At this very high tilt angle, the protrusion of the omega
boundary becomes elongated into an extra lamellar half-plane. Towards the
top of the micrograph this omega boundary splits into two lower angle
boundaries, an 89° omega boundary and a 28° chevron boundary. Together
these two lower angle boundaries achieve approximately the same tilt
reorientation as prior to the split. In the lower right of the micrograph there
is a pure T-geometry boundary. Moving from right to left along the
boundary, it becomes more symmetric and a transition occurs into the
chevron morphology. As the transition occurs some lamellae terminate at
476 Morphology of block copolymers

(a)

(e)
Figure 10.15 Tilt and twist grain boundaries in lamellar block copolymers. (a) The
reference state is an aligned bicrystal with no boundary. (b) A tilt boundary is
produced by rotating the grains with respect to one another about an axis which is
parallel to the plane of the grain boundary. (c) A twist boundary is produced by
rotating the grains with respect to one another about an axis perpendicular to the
grain boundary. The angle Q( measures the twist of the lattice. (Reproduced from
Gido et at., 1993a, with permission.)

the T -junction whilst others bend across like a chevron. The tilt grain
boundaries have been investigated in detail for the lamellar case and only
limited information is available about hexagonal and cubic structures
(Carvalho, Lescanec and Thomas, 1995).
Two types of twist grain boundaries have been observed in lamellar block
copolymers (Gido et al., 1993a,b) in which microphase separation of the two
blocks was maintained in the grain boundary region by intermaterial
dividing surfaces which approximate minimal surfaces (Thomas et al., 1988).
The geometry of these interfaces was demonstrated by comparing experi-
mental TEM images with ray-tracing computer simulations (Gido and
Thomas, 1994a,b). The two morphologies observed were found to have
intermaterial dividing surfaces that approximate either Scherk's first (doubly
Morphology 477

(a ) (b)
Boundary
Plane

Sym metry t--=.:;::::::;;;;:::.:.:q


Pl ane

(c ) (d)
Figure 10.16 Tilt grain boundary geometry in lamellar block copolymers. (a) Sym-
metric tilt boundary. The tilt angle e is measured between the perpendiculars to the
two sets of lamellar planes. (b) By comparison to (a), it can be seen that only the
component of tilt measured in the plane that is simultaneously perpendicular to both
sets of lamellar planes on both sides of the boundary is observable. The tilt
misorientation does not produce a grain boundary. (c) When the placement of the
boundary plane is no longer symmetrical, the deviation from symmetry is quantified
by the angle ({J between the boundary plane and the plane of symmetry which
bisects e. (d) The pure T-geometry occurs when e = 90° and (p = 45°. (Redrawn
from Gido and Thomas, 1994b.)

periodic) surface or a section of the right helicoid. The helicoid was only
found at low twist angles, less than 15°, whereas the Scherk surfaces
were found over the entire range of twist. As the twist angle approaches 0°,
the Scherk surface grain boundary morphology is transformed into a single
screw dislocation that has an intermaterial dividing surface with the
geometry of a single helicoid, which is illustrated in Figure 10.18.
478 Morphology of block copolymers

Figure 10.17 TEM of an approximately symmetric PS- PI diblock (Mw = 156 k)


showing a number of tilt boundaries. A 125° symmetric tilt boundary is labelled E.
At this very high tilt angle, the protrusion of the omega boundary becomes
elongated into an extra lamellar half-plane. Towards the top of the micrograph, this
omega boundary splits into two lower angle boundaries: an 89° omega boundary,
labelled n, and a 28° chevron boundary, labelled (1 . A T-junction boundary
is labelled T. Proceeding left along this boundary, a transition occurs to the chevron
morphology, labelled C2. The region of gradual transition between the T-junction
and the chevron is labelled TIC. (Reproduced from Gido and Thomas, 1994b, with
permission.)

(d) Blends of Diblocks and Homopolymers


There are two regimes of behaviour in blends of diblock copolymers and
homopolymers. In one regime mixing takes place on the molecular level,
where N homo < Nblock' that is where the homopolymer molecular weight
is less than that of the corresponding block in the block copolymer,
even though the system is micro phase separated. In the other regime
demixing (macrophase separation) takes place on the molecular level, where
Morphology 479

(d)

400 nm

Figure 10.18 TEM tilt series of 0° twist Scherk surface in a blend of 60 vol% PS - PI
diblock (Mw = 156 k, fps = 0.48), 20 vol% homo-PS (17 k) and 20 vol% homo-PB
(21 k). (a - c) TEM tilt series spanning 50° of a projection angle in a plane normal to
the screw axis. The projected image of the dislocation core is invariant with tilt in
this series. (d) Computer simulation of a 0° twist Scherk surface projecting normal
to the z-axis (Reproduced from Gido et aI., 1993a, with permission .)

N homo> N block ' that is where the homopolymer molecular weight is more
than that of the block copolymer, and there are large domains of pure
homopolymer and domains of microphase separated block copolymer.
In the N homo < N block regime the homopolymer acts as a preferential
solvent. Addition of a low molecular weight homopolymer to a block
copolymer causes dilation of the subphase and a progression across the
equilibrium morphologies to increasing volume fractions. Thus a PS- PI
diblock, of low PS content, which forms a spherical morphology at equilib-
rium will successively from PS rods, then lamellae on addition of increasing
amounts of low molecular weight homo-PS. Continued addition of homo-
PS will lead to the formation of PI rods and then PI spheres, which can be
either ordered on a lattice or disordered. These types of system have been
studied in detail, both experimentally and theoretically, and predictions of
the morphology are available if the characteristics (X(T, <p) and the statistical
segment length) of the polymers are known. Winey and coworkers (Winey,
Thomas and Fetters, 1992a,b; Winey et al., 1994) studied styrene- isoprene
and styrene- butadiene systems by SAXS and TEM where N homo < NbloCk.
Of particular usefulness were the isothermal morphology diagrams
which could be used to describe the phase structure obtained when blend-
ing. Figure 10.19 illustrates a typical constant copolymer composition
480 Morphology of block copolymers

2.0.-----:------::=---------------,
f =44 - 51 vol% PS

0.0 L -_ _-L-_ _-L-_ _- L_ _-l.._ _---L_ _---l

40 50 60 70 80 90 100
Overall PS volume Percent in Blend

Figure 10.19 Constant copolymer composition morphology diagram for a 49 k


PS-PI (fps = 0.51) blended with homo-PS. L, DD, C and S refer to the observed
structures of lamellae, double diamonds, cylinders and spheres. (Redrawn from
Winey, Thomas and Fetters, 1992a.)

morphology diagram for a 49k PS-PI (fps = 0.51) blended with homo-PS.
One important feature is that as the homo-PS is added, the morphology
transforms from lamellar to a bicontinuous cubic phase (labelled DD), to
hexagonal rods, to cubic spheres and then disordered micelles only if
NhomjNblock < 0.5. As the molecular weight of the homopolymer is in-
creased, the transformations directly from hexagonal rods to disordered
micelles and from lamellae to disordered micelles are possible. Similar
observations were also made by Tanaka, Hasegawa and Hashimoto (1991)
and Hashimoto, Tanaka and Hasegawa (1990) using a combination of
TEM and SAXS. It is interesting to note that Winey et al (1994) observed
the OBDD morphology in blends by a combined TEM and SAXS; current
opinion (E.L. Thomas, personal communication, 1996) is that the bicontinu-
ous cubic morphology obtained is in fact the gyroid and the previous
assignment (based on SAXS data from a Kratky camera) was in error.
The spatial distribution of the homopolymer in a block copolymer-
homopolymer blend has been studied by a combination of SAXS and SANS
by Koizumi, Hasegawa and Hashimoto (1994b) and Lee et al. (1994).
Different chemical systems were used but the common feature was that
deuterated PS was dissolved in PS-PI block copolymers. Both studies
found that when N homo < Nblock' the homo-PS dissolved in the PS domains
and the PI domain thickness was left unaltered and that the homo-PS
is localized in the middle of the PS domain. This implies that the block
Morphology 481

(d)

Figure 10.20 TEM micrograph from a mixture containing 20 wt% PS - PI diblock


copolymer (fr,s = 0.83, Mn = 82 k) in PS homopolymer (Mn = 577 k). (a) Lens-like
domains in the mixture cast from toluene under normal pressure. PI cylinders packed
hexagonally are clearly visible in (b) and (c). (d) Disordered micelles in the blend cast
from dichloromethane under low pressure. (Reproduced from Koizumi, Hasegawa
and Hashimoto, 1994a, with permission.)
482 Morphology of block copolymers

copolymer acts like a swelling polymer brush when homopolymer is being


solubilized.
Experimental work on the N homo» N block regime has been rather sparse
in comparison. An early study by Eastmond and Phillips (1979) demon-
strated that demixing occurs and that vesicles are formed. The morphology
obtained is highly dependent on the way the material is prepared. Koizumi,
Hasegawa and Hashimoto (1994a) have shown that mixtures processed
from solution are highly dependent on the solvent evaporation rate. For
example when a PS- PI diblock, which forms PI rods at equilibrium, was
mixed with high molecular weight PS, two morphologies could be formed
depending on the solvent evaporation rate. When the solvent was removed
quickly, a fractal-like distribution of spherical microdomains was obtained
whereas when the solvent was removed slowly, lens-like macrodomains of
hexagonal block copolymer were obtained; these two structures are illus-
trated in Figure 10.20. A similar result is obtained for a lamellar-forming
PS- PI where large onion-like lamellar structures are formed (Figure 1O.21a)
when the evaporation rate is slow and small vesicles are obtained when the
solvent evaporation rate is high (Figure 10.21b).
Solution casting is the route preferred by experimentalists and the
structures observed are interesting. It is difficult, however, to develop
universal structure-morphology relations as the molecular and processing
variables are enormous. As a general rule, when N homo» N block the mor-
phology formed will be two-phase, one phase being the homopolymer and
the other phase being the block copolymer with its equilibrium morphology.
The size of the macrophase structure is kinetically controlled.

10.4.2 MULTIBLOCKS VERSUS DIBLOCKS


Commercially, the most important class of block copolymers are the
multiblocks made by direct polycondensation (Legge, Holden and
Schroeder, 1987). This class of materials includes polyurethanes, polyesters
and nylon block copolymers. The materials commonly referred to as
polyurethanes (PU) [which include poly(ester-urethane)s, poly(ether-
urethane)s, poly(ether-urea)s and poly(ether-isocyanurate)s], elastomeric
polyesters [poly(ether-ester)s] and elastomeric polyamides [poly(ether-am-
ide)s] are (HmS)n type block copolymers consisting of alternating blocks of
hard, crystalline or glassy segments (H) and soft, rubbery segments (S).
Because of the polymerization mechanism they generally have a distribution
in the hard segment length, the polydispersity of the soft segment being set
by the original oligomer. Because of the distribution in both overall
molecular weight and block length, (HmS)n copolymers do not tend to form
structures with long range order.
A reasonable determination of the morphology of polyurethanes by direct
imaging has proved experimentally evasive due to the size scales involved.
Morphology 483

Hl02 / 570 (40/60) PS· ..............

IlomopolYMyrene

111 m

(a)

(b)
Figure 10.21 (a) TEM from a 40/60 wt% mixture of a PS- PI diblock and a PS
homopolymer. The PS - PI diblock has fps = 0.50 and Mn = 100 k and the
homopolymer has Mn = 577 k. Lamellar microdomains formed by the diblock are
packed in an onion-type ves icle. (b) TEM from a 20/80 wt% mixture of the same
polymers, cast from a 5% solution in toluene by rapid evaporation under normal
pressure. (Reproduced from Koizumi, Hasegawa and Hashimoto, 1994a, with
permission.)
484 Morphology of block copolymers

(a)
10' 0

~
'2 0
:::I

:s
~ 10"
:;
......
§
10'

10'

(b) o O.OS 0. 1 O. IS 0.2

Figure 10.22 (a) TEM of a copolyisocyanurate-urea showing a typical microphase-


separated structure with no long range order. The copolymer has 66 vol% poly-
isocyanurate and 34 vol% poly(ether-isocyanurate). (b) SAXS pattern of the same
polymer showing a single peak corresponding to the dominant length scale of 110 A.
(Reproduced from Ryan. Standford and Tao. 1993. with permission.)
Morphology 485

However, Chen-Tsai et al. (1986) and Li et al. (1988) have imaged regular
structures with size scales of ~ 10 nm in well characterized polybutadiene-
based PU materials, and this corresponds with the approximate radius of
gyration of the soft segment chains. The morphology of the poly(ether-
urea)s and poly(ether-urethane)s is less well defined. TEM shows that phase
separation exists on the size scale < 50 nm (Ryan, Stanford and Still, 1988).
Willkomm et al. (1988) have shown using SAXS that similar polyureas have
a phase structure with periodicities of the order of 10 to 30 nm. More
recently, Tang et al. have imaged rod-like structures in a poly(ether-
urethane) by an element specific SEM technique (Tang et ai., 1994).
Typical TEM micrographs of a crosslinked (HmS)n copolymer is given in
Figure 10.22a for a poly(ether-isocyanurate) that contains 66% glassy
polyisocyanurate. The material has been annealed close to Ig so that it is as
close to equilibrium as possible and the micrograph shows a system that has
good electron density contrast but no long range order. This structure gives
rise to the SAXS pattern in Figure 1O.22b which has just one peak. Similar
SAXS patterns have been observed for polyurethane elastomers (Abouzahr
and Wilkes, 1985) and polyurethane foams (Armistead, Turner and Wilkes,
1988; Creswick et al., 1989), and can be analysed in terms of the correlation
function in order to extract more information about the morphology from
the scattering pattern (Abouzahr and Wilkes, 1985). The interpretation of
morphology in the multiblock polymers discussed above is further compli-
cated by the fact that they are often highly branched or even crosslinked.
The structure observed in multi blocks made by direct polycondensation
indicates that long range order is not attainable even though this would be
the equilibrium state. The microphase separation of polyurethanes and
poly(ether-urea)s has been analysed (Ryan, Standford and Still, 1990) in
terms of Leibler's block copolymer theory and satisfactory agreement with
the predictions for the degree of microphase separation were obtained.
Studies of diblock copolymers show that the ordering process is two-stage
with an initial formation of the morphological units (spheres, rods and
lamellae) followed by a cooperative reorganization into a structure with
long range order (Singh et al., 1993; Stiihn, Vilesov and Zachmann, 1994).
It could be that the first process is all that occurs in multiblocks, and the
morphology gets stuck at this stage as the cooperativity required to achieve
long range order, from chains which may reside in up to ten domains, is not
available. The combination of block length polydispersity and a lack of long
range order would certainly account for the broad scattering peaks obtained
from SAXS and SANS observed for multi blocks made by direct polycon-
densation. In agreement with these observations, the mean field theory of
Fredrickson, Milner and Leibler (1992) predicts that long range order is
supressed in microphase separated multiblock copolymers due to a broad
distribution of monomer sequence lengths. An alternative explanation is
486 Morphology of block copolymers

(AB}n

---
-
24

(I)

(II)
Morphology 487
--------------------------------------------
that the structure observed in multiblocks are composition fluctuations
pinned by vitrification of the glassy microphase. This is most certainly the
case for block copolymers where the glassy phase has a high ~ (Ryan,
Stanford and Still, 1988, 1990; Ryan, Stanford and Tao, 1993).
Multiblock copolymers have been made by anionic polymerization so
that the effect of block number could be systematically studied. In an elegant
TEM and SAXS study, Smith and coworkers (Spontak, Smith and Ashraf,
1993; Smith et al., 1994) showed that even for styrene-isoprene (SI)n
multiblock polymers made by anionic polymerization, long range order was
reduced for n = 4. The effect of block number is illustrated in Figure
1O.23i(a)-(d) by the TEM micrographs of (SI)n multiblocks. As n increases,
the lamellae are observed to be thinner and the long range order decreases.
The reasons for these two morphological changes are illustrated in Figure
10.23ii(a)-(f) which compares the conformations available to the
copolymers.

10.4.3 ABC TRIBLOCKS

Given the morphological complexity of AB diblock and ABA triblock


copolymers, it might be expected that the phase behaviour of ABC triblocks
would be even more rich, and indeed this has been confirmed by recent
experiments from a number of groups. From a practical viewpoint, ABC
triblocks can also act as compatibilizers in blends of A and C homopolymers
(Auschra and Stadler, 1993). In addition to the composition of the
copolymer, an important driving force for structure formation in these
polymers is the relative strength of incompatibilities between the compo-
nents, and this has been explored by synthesis of chemically distinct block
copolymers.
In the earliest studies, so-called 'core-shell' morphologies were identified,
where the core of the minority end component is separated from the other
end-block by a shell of the mid-block. Spherical microdomains surrounded
by a shell of the mid-block have been observed using TEM for a number of
polymers (Riess, Schlienger and Marti, 1989; Arai et al., 1980). Mogi et al.

Figure 10.23 (i) TEM of a series (a-d) of (AB)n block copolymers with 4>ps = 0.5
showing contraction of domain spacing and reduction in long range order as n is
increased. (ii) A shematic illustration of the molecular conformations linear block
copolymers may adopt within the AB lamellar morphology. Only one conformation
is available to the diblock copolymer (a) whereas an ABA may assume two, one
bridged (b) and one looped (c). An (AB)3 block copolymer is capable of 16 different
conformations; the (c) limiting cases of fully bridged (d) and fully looped (e) are
shown along with an example of an intermediate case (f). (Reproduced from Smith
et aI., 1994, with permission.)
488 Morphology of block copolymers

Figure 10.24 TEM from a poly(2-vinyl pyridine)- polyisoprene- polystyrene ABC


triblock copolymer, showing a hexagonal cylinder 'core- shell structure' . (Gido et aI.,
1993b). The light, grey and black regions correspond to the PS, PVP and PI
respectively in sections (a) perpendicular and (b) parallel to the rod direction.
(Reproduced from Gido et aI. , 1993b, with permission.) A schematic of the structure
is also shown in (c).

(1992a) studied a series of polyisoprene- polystyrene- poly(2-vinyl pyridine)


(PI - PS- P2VP) triblocks where the volume fraction of the PS middle block
was varied from 0.3 to 0.8, whilst the relative volume fractions of the
end-blocks were kept equal. On increasing the PS volume fraction a
core-shell lamellar structure (composition 1: 1: 1), a tricontinuous structure
(1 : 2: 1), a cylindrical phase with separate PS and P2VP rods (1: 4: 1) and
a spherical morphology (1: 8 : 1) were observed. It was suggested that the
most efficient packing of the two types of rods leads to a tetragonal
morphology. The tricontinuous structure in this system was analysed in
detail in a separate paper (Mogi et at., 1992b), and assigned an 'ordered
tricontinuous double diamond' morphology, by analogy with the 'ordered
bicontinuous double diamond' structure that was then thought to exist in
AB block copolymers.
Morphology 489

Illustrating the importance of the relatiave strengths of incompatibility


between blocks, an ABC triblock made of the same components as in the
polymers studied by Mogi et al. (1992a,b), but with the sequence changed,
was investigated by Gido et al. (1993b). TEM micrographs from a PVP-
PI - PS triblock with equal volume fractions are shown in Figure 10.24.
They clearly reveal a cylindrical 'core-shell' structure for this sample with
composition 1: 1 : 1, with a PVP core. This structure forms even though a
PI - PS-PVP triblock with the same composition formed a lamellar core-
shell morphology (Mogi et al., 1992a); the molecular weights of the
polymers differs by a factor of two, but this does not change the microstruc-
ture if the blocks are strongly segregated. This morphology change was
rationalized by Stadler et al. (1995) on the basis that a lamellar morphology
is expected for a PI-PS-PVP symmetric triblock (cPA = cPs = cPd because
the interfacial tension )!PI-PS '" )!pspvp. However, for PVP-PI - PS,
)!PI-PVP »)!PI-PS, so that the system will tend to create a large PS-PI
interface which allows a reduction in stretching of PS (and PI) chains and
a small PI - PVP contact area to reduce interactions between these highly
immiscible polymers. As a result of this imbalance between A- Band B-C
interactions, a core-shell morphology with a PVP core forms.
A number of remarkable new morphologies have recently been discovered
by Stadler and coworkers in polystyrene-polybutadiene-poly(methyl meth-
acrylate) (PS- PB-PMMA) triblocks, and their hydrogenated analogues,
polystyrene-pol y( eth ylene-co-bu tylene )- po Iy( meth yI methacrylate) (PS-
PEB-PMMA). The common features of the polymers exhibiting this
complex phase behaviour are that the mid-block is the minority component
and that the incompatibility between the outer blocks is much weaker than
the incompatibility of each of these blocks with respect to the mid-block.
Selective staining with OS04 was used to stain the PB domain in PS-PB-
PMMA triblocks, whereas Ru0 4 stains both PS and PB. In contrast,
sufficient contrast between blocks was achieved using Ru0 4 for PS-PEB-
PMMA samples.
A structure consisting of PEB cylinders at the interface between PS and
PMMA lamellae was observed for a triblock with volume fractions
cPps = 0.48, cPPB = 0.20 and cPPMMA = 0.32 (Figure 10.25; Auschra and
Stadler, 1993). A morphology with PB spheres at the PS-PMMA interface
that was observed using TEM for a triblock with cPps = 0.47, cPPB = 0.075
and cPPMMA = 0.455 (Beckmann, Auschra and Stadler, 1994) is shown in
Figure 10.26. In contrast, when the mid-block of the triblock is hydro-
genated to give a PS-PEB- PMMA triblock with the same composition, a
morphology with PEB rings around PS cylinders in a matrix of PMMA is
observed (Auschra and Stadler, 1993; Figure 10.27). Incorporation of the
PEB block changes the PS- PMMA morphology, as well as the domain
490 Morphology of block copolymers

(a)

(b)
Morphology 491

projection 1

000 loool

(c)

Figure 10.25 TEM and schematic structure of the rods between lamellae morphol-
ogy in a polystyrene- poly(ethylene-co-butylene)- poly(methyl methacrylate) (PS -
PEB- PMMA) triblock copolymer with CPps = 0.48, CPPEB = 0.20 and CPPMMA = 0.32.
The PS lamellae appear grey and the PMMA lamellae are light. Two micrographs
(a, b) and a schematic (c) of the morphology are shown to aid interpretion.
(Reproduced from Beckmann, Auschra and Stadler, 1994, with permission.)
492 Morphology of block copolymers

(a)

'.J.
,
t •

.~.

(b)

Figure 10.26 TEM and schematic structure of spheres at lamellar interfaces for a
polystyrene-polybutadiene- poly(methyl methacrylate) (PS - PB- PMMA) triblock with
<pps = 0.47, <PPB = 0.075 and <PPMMA = 0.455. The PS lamellae appear grey and the
PMMA lamellae are light. A micrograph (a) and a schematic (b) of the morphology
are shown to aid interpretion . (Reproduced from Beckmann, Auschra and Stadler,
1994, with permission.)
Morphology 493

(a)

C
~B
- - --r--A

(b)

Figure 10.27 Schematic of (a) 'rings around rods' and (b) 'helices around rods'
morphologies obtained in (PS - PB - PMMA) and (PS- PEB -- PMMA) triblock
copolymers. (Source: Auschra and Stadler, 1993.)

structure adopted by the mid-block, because the A- C and B-C interaction


parameters are changed, as well as the B segment length. An even more
complex morphology was observed in a PS- PB- PMMA triblock with
<pps = 0.26, <PPB = 0.14 and <PPMMA = 0.60 (Krappe, Stadler and Voigt-Mar-
tin, 1995). On the basis of TEM, it was concluded that this copolymer forms
a structure with helical rings of PB around PS cylinders (Figure 10.27).
These observations of complex phases formed by mid-block segregation
at the AC interface have been accounted for by Stadler et al. (1995), who
extended the Semenov strong segregation theory (Semenov, 1985) to the case
of ABC triblocks. Using this approach, a phase diagram was constructed for
494 Morphology of block copolymers

the copolymers studied by Stadler and coworkers, which have a mid-block


composition 4>B < 0.4, which is in good agreement with the experimental
results (Stadler et al., 1995), although core-shell structures and the structure
with helical rings around rods was not considered.

10.5 SUMMARY AND CONCLUSIONS


The physics of block copolymers provides a rich seam of morphology in
which complex, well ordered structures can be found. When one or more of
the micro phases is a glassy polymer the structure is stabilized, compared
to the melt, and the microstructure formed can then be utilized. Current
applications of block copolymers containing a glassy block make use of the
bulk mechanical properties in thermoplastic elastomers and polymer (spe-
cifically polyurethane) foams. Block copolymers find other uses as highly
functional, added value products such as compatibilizers in polymer blends,
viscosity modifiers in lubricating oils and as food additives to control
texture, which are beyond the scope of this chapter but are reviewed in the
book by Legge, Holden and Schroeder (1987) and by Fredrickson and Bates
(1996). The future exploitation of block copolymers lies in their established
application as structural materials, but more importantly as functional
materials in diverse areas such as drug delivery systems and temperature
sensors. Currently, the ability to control anisotropy and surface composition
through self-assembly is not well utilized. The anisotropy of block copoly-
mers on the 10 nm scale could enable their use as chemically selective
membranes or catalyst supports, especially if one of the microphases could
be removed chemically to form a nano-porous solid. If one (or more) of the
microphases could be doped, then block copolymers could be used as
electronic or photonic materials, and the potential for the production of
nano-capacitors and nano-wires is obvious from their microstructure. Final-
ly, we will find block copolymers used in the increasingly important
emerging technologies of microelectronics and medicine where high per-
formance and surface specificity are more important than the cost of the
polymer.

REFERENCES
Abouzahr, S. and Wilkes, G.L. (1985) in Processing, Structure and Properties of Block
Copolymers (ed. M.J. Folkes), Elsevier, London, pp. 165-207.
Aggarwal, S.L. (1985) in Processing, Structure and Properties of Block Copolymers
(ed. M.J. Folkes), Elsevier, London, pp. 1-28.
Almdal, K., Rosedale, J.H., Bates, F.S. et al. (1990) Phys. Rev. Lett., 65,1112.
Alward, D.B., Kinning, D.J., Thomas, E.L. and Fetters, L.J. (1986) Macromolecules,
19, 215.
References 495

Arai, K., Kotaka, T., Kitano, Y. and Yoshimura, K. (1980) Macromolecules, 13,1670.
Armistead, 1.P., Turner, R.B. and Wilkes, G.L. (1988) J. Appl. Polym. Sci., 35, 601.
Auschra, e. and Stadler, R. (1993) Macromolecules, 26,2171.
Auserre, D., Raghunathan, V.A. and Maaloum, M. (1993) J. Phys. France II, 3, 1485.
Balta-Calleja, F.J. and Yonk, e.G. (1989) X-ray Scattering of Synthetic Polymers,
Elsevier, London.
Bates, F.S. and Fredrickson, G.H. (1990) Annu. Rev. Mater. Sci., 41, 525.
Bates, F.S., Keith, H.D. and McWhan, D.B. (1987) Macromolecules, 20,3065.
Bates, F.s., Rosedale, J.H. and Fredrickson, G.H. (1990) J. Chem. Phys. 92, 6255.
Bates, F.S., Rosedale, 1.H., Fredrickson, G.H. and Glinka, C.J. (1988) Phys. Rev.
Lett., 61, 2229.
Beckmann, J., Auschra, e. and Stadler, R. (1994) Macromol. Rapid Commun., 15, 67.
Bras, W., Derbyshire, G.E., Bogg, D. et al. (1995) Science, 267, 996.
Brazovskii, S.A. (1975) Zh. Eksp. Teor. Fiz., 68, 175 (English translation: Sov. Phys.
JETP, 41, 85).
Carvalho, B.L. and Thomas, E.L. (1994) Phys. Rev. Lett., 73, 3321.
Carvalho, B.L., Lescanec, R.L. and Thomas, E.L. (1995) Macromol. Symp., 98, 1131.
Chen-Tsai, e.H.Y., Thomas, E.L., MacKnight, W.J. and Schneider, N.S. (1986)
Polymer, 27, 659.
Collin, B., Chatenay, D., Coulon, G. et al. (1992) Macromolecules, 25, 1621.
Coulon, G., Collin, B., Ausserre, D. et al. (1990) J. Phys. (Paris), 51, 2801.
Creswick, M.W., Lee, K.D., Turner, R.B. and Huber, L.M. (1989) J. Elast. Plast., 21,
179.
de Jeu, W.H., Lambooy, P., Hamley, I.W. et al. (1993) J. Phys. France II, 3, 139.
Disko, M.M., Liang, K.S., Behal, S.K. et al. (1993) Macromolecules, 26, 2983.
Dobrynin, A.V. (1995) J. Phys. France II,S, 657.
Dobrynin, A.Y. and Erukhimovich, I.Y. (1993) Macromolecules, 26, 276.
Eastmond, G.e. and Phillips, D.G. (1979) Polymer, 20, 1501.
Fischer, H. (1994) Polymer, 35, 3787.
Folkes, M.1. (ed.) (1985) Processing, Structure and Properties of Block Copolymers,
Elsevier, London.
Folkes, M.1. and Keller, A. (1973) in The Physics of Glassy Polymers (ed. R.N.
Haward), Applied Science, London.
Forster, S., Khandpur, A.K., Zhao, 1. et al. (1994) Macromolecules, 27,6922.
Fredrickson, a.H. (1991) Macromolecules, 24, 3456.
Fredrickson, a.H. and Bates, F.S. (1996) Annu. Rev. Mater. Sci., 26, 501.
Fredrickson, G.H. and Helfand, E. (1987) J. Chem. Phys., 87, 697.
Fredrickson, G.H., Milner, S.T. and Leibler, L. (1992) Macromolecules, 25,6341.
Gedde, U.W. (1995) Polymer Physics, Chapman & Hall, London.
Gehlsen, M.D. and Bates, F.S. (1993) Macromolecules, 26, 4122.
Gido, S.P. and Thomas, E.L. (l994a) Macromolecules, 27,849.
Gido, S.P. and Thomas, E.L. (1994b) Macromolecules, 27,6137.
Gido, S.P., Schwark, D.W.,. Thomas, E.L. and do Carmo Goncalves, M. (1993a)
Macromolecules, 26, 2636.
Gido, S.P., Gunther, 1., Thomas, E.L. and Hoffman, D. (1993b) Macromolecules, 26,
4506.
Hadziioannou, G., Picot, e., Skoulios, A. et al. (1982) Macromolecules, 15, 263.
Hajduk, D.A., Harper, P.E., Gruner, S.M. et al. (1994) Macromolecules, 27, 4063.
Hajduk, D.A., Harper, P.E., Gruner, S.M. et al. (1995) Macromolecules, 28, 2570.
Hamley, I.W. and Bates, F.S. (1994) J. Chem. Phys., 100, 6813.
Hamley, I.W., Koppi, K.A., Rosedale, 1.H. et al. (1993) Macromolecules, 26, 5959.
Hamley, I.W., Gehlsen, M.D., Khandpur, A.K. et al. (1994) J. Phys. France II, 4,
2161.
496 Morphology of block copolymers

Hasegawa, H., Tanaka, H., Yamasaki, K. and Hashimoto, T. (1987) Macromolecules,


20, 1651.
Hashimoto, T., Tanaka, H. and Hasegawa, H. (1990) Macromolecules, 23, 4378.
Helfand, E. (1975) Macromolecules, 8, 553.
Helfand, E. and Wasserman, Z.R. (1976) Macromolecules, 9,879.
Helfand, E. and Wasserman, Z.R. (1982) in Developments in Block Copolymers,
Vol. 1 (ed.1. Goodman), Applied Science, New York, pp. 99-125. .
Higgins, IS. and Benoit, H.C. (1994) Polymers and Neutron Scattering, Clarendon
Press, Oxford.
Kato, K. (1965) J. Electron Microscopy (Japan), 14, 220.
Kato, K. (1967) Polym. Eng. Sci., 7, 38.
Keller, A., Pedemonte, E. and Willmouth, F.M. (1970) Nature, 225, 538.
Khandpur, A.K., Macosko, C.W. and Bates, F.S. (1995) J. Polym. Sci. B, 33, 247.
Khandpur, A.K., Forster, S., Bates, F.S. et al. (1995) Macromolecules, 28,8796.
Koizumi, S., Hasegawa, H. and Hashimoto, T. (1994a) Macromolecules, 27, 6532.
Koizumi, S., Hasegawa, H. and Hashimoto, T. (1994b) Macromolecules, 27, 7893.
Koppi, K.A., Tirrell, M. and Bates, F.S. (1993) Phys. Rev. Lett., 70, 1449.
Krappe, u., Stadler, R. and Voigt-Martin, I. (1995) Macromolecules, 28, 4558.
Lee, S.H., Koberstein, IT., Quan, X. et al. (1994) Macromolecules, 27, 3199.
Legge, N.R., Holden, G. and Schroeder, H.E. (Eds) (1987) Thermoplastic Elastomers:
A Comprehensive Review, Hanser, New York.
Leibler, L. (1980) Macromolecules, 16, 1302.
Li, c., Goodman, S.L., Albrecht, R.M. and Cooper, S.L. (1988) Macromolecules, 21,
2367.
Likhtman, A.E. and Semenov, A.N. (1994) Macromolecules, 27,3103.
Liu, Y., Zhao, W., Zheng, X. et al. (1994) Macromolecules, 27, 4000.
Matsen, M.W. and Bates, F.S. (1996) Macromolecules, 29, 1091.
Matsen, M.W. and Schick, M. (1994) Phys. Rev. Lett., 72, 2660.
Mayes, A.M. and Olvera de la Cruz, M. (1989) J. Chern. Phys., 91, 7228.
Mogi, Y., Kotsuji, H., Kaneko, Y. et al. (1992a) Macromolecules, 25,5408.
Mogi, Y., Mori, K., Matsushita, Y. and Noda, I. (1992b) Macromolecules, 25,5412.
Mortensen, K. and Pedersen, J.S. (1993) Macromolecules, 26, 805.
Olmsted, P.D. and Milner, S.T. (1994) Phys. Rev. Lett., 74, 829.
Olvera de la Cruz, M., Mayes, A.M. and Swift, B.W. (1992) Macromolecules, 25, 944.
Pochan, D.I., Gido, S.P., Pispas, S. et al. (1996) Macromolecules, 29,5091.
Podneks, V.E. and Hamley, I.W. (1996) JETP Lett., 64, 619; Pisma Zh. Eksp. Teor.
Fiz., 64, 564.
Rangarajan, R., Register, R.A., Adamson, D.H. et al. (1995) Macromolecules, 28,
1422.
Riess, G., Schlienger, M. and Marti, S. (1989) J. Macromol. Sci. B, Phys., 17, 355.
Russell, T.P. (1990) Mater. Sci. Rep., 5, 171.
Ryan, A.I., Stanford, J.L. and Still, R.H. (1988) Br. Polym. J., 20, 77.
Ryan, A.J., Stanford, IL. and Still, R.H. (1990) Plast. Rubb. Proc. Appl., 13, 99.
Ryan, A.I., Stanford, IL. and Tao, X.Q. (1993) Polymer, 34, 4021.
Ryan, A.I., Hamley, I.W., Bras, W. and Bates, F.S. (1995) Macromolecules, 28, 3860.
Schulz, M.F., Khandpur, A.K., Bates, F.S. et al. (1996) Macromolecules, 29, 2857.
Semenov, A.N. (1985) Sov. Phys. JETP, 61, 733.
Sikka, M., Sin~, N., Karim, A. et al. (1993) Phys. Rev. Lett., 70, 307.
Singh, M.A., Harkless, c.R., Nagler, S.E. et al. (1993) Phys. Rev. B, 47, 8425.
Smith, S.D., Spontak, R.I., Satkowski, M.M. et al. (1994) Polymer, 35, 4527.
Spontak, R.J., Smith, S.D. and Ashraf, A. (1993) Macromolecules, 26, 5118.
Spontak, R.I., Fung, Ie., Braunfield, M.B. et al. (1996) Macromolecules, 29, 4494.
RE~ferences 497

Stadler, R., Auschra, c., Beckmann, 1. et al. (1995) Macromolecules, 28, 3080.
Stiihn, B., Vilesov, A. and Zachmann, H.G. (1994) Macromolecules, 27, 3560.
Szwarc, M., Levy, M. and Milkovich, R. (1956) J. Am. Chem. Soc., 78, 2656.
Tanaka, H., Hasegawa, H. and Hashimoto, T. (1991) Macromolecules, 24, 240.
Tang, W., Farris, R.J., MacKnight, W.J. and Eisenbach, C.D. (1994) Macromolecules,
27,2814.
Thomas, E.L. and Lescanec, R.L. (1994) Phil. Trans. R. Soc. London A, 348, 149.
Thomas, E.L., Anderson, D.M., Henkee, C.S. and Hoffman, D. (1988) Nature, 334,
598.
Thomas, E.L., Alward, D.B., Kinning, D.J. et al. (1986) Macromolecules, 19, 2197.
van Dijk, M.A. and van den Berg, R. (1995) Macromolecules, 28, 6773.
Willkomm, W.R., Chen, Z.S. and Macosko, C.W. et al. (1988) Polym. Eng. Sci., 28,
888.
Winey, K.I., Thomas, E.L. and Fetters, L.J. (1991) J. Chem. Phys., 95, 9367.
Winey, K.I., Thomas, E.L. and Fetters, L.J. (1992a) Macromolecules, 25, 422.
Winey, K.I., Thomas, E.L. and Fetters, L.J. (1992b) Macromolecules, 25, 2645.
Winey, K.I., Gobran, D.A., Xu, Z. et al. (1994) Macromolecules, 27, 2392.
Index

ABC triblock copolymers 487-3 gyroid and perforated layers 468-73


use as compatibilisers 487 multi block copolymers 483- 7
core-shell morphologies 487-9 small-angle neutron scattering 457,
cylinders and rods 489-91 462,464,470,474
rods with rings and helices 493 small-angle x-ray scattering 457
ABS solution casting 482 -- 3
fracture toughness 358 surface morphology 458
J value 358-9 techniques of investigation 456-8
load-displacement curves 358 tilt grain boundaries 473-8
Activation volume for yield 189 transmission electron microscopy
A.dhesion, energy balance 436 456-7,472,484
Adiabatic deformation 246 triblock copolymers, 463-7
Affine deformation 70 twist grain boundaries, 473-8
Anisotropic united atom potential 42 Block copolymers 422-34, 451-97
Anomalous diffusion 78-9 architecture 452
Argon's model for yield 198--200,270 concentration at interfaces 431
Arruda and Boyce eight chain network diblock copolymers 458-87
model 272-7 microphase separation theory 453-8
ASTM fracture tests 347 reinforcement of interfaces 422-35
Atomic force microscopy of block synthesis 452
copolymers 465 triblock copolymers 463-7
Atomistic modelling 33-4 Bond orientational motions 51
Bond rotations 131
Blends Bond rupture 364
compatibilizers 487 Bond strength 14
diblock copolymers and Brittle strength 15
homopolymers 479-82 Bulk modulus 161
rubber toughening 366
Block copolymer morphology Cavitation diagrams 387-91
atomic force microscopy 465 Chain configurations, initial 43
blends of diblocks and homopolymers Chain conformations 9
479-82 Chain dynamics 48
complex phases 468-73 Chain ends, entropic segregation 443
diblock copolymers 458-87 Chain flexibility, modelling 41
dislocations 473-9 Chain interaction energies 40
grain boundaries 473-9 Chain pull-out 317
500 Index

Characteristic ratio 67 molecular mechanisms 310-19


Coarse grain modelling 34, 80 molecular scission 315
Compact tension 348 molecular weight 323-5, 338
Compatibilizers 487 morphology, 297-301
Compression nucleation 302
effect on crazing 165 optical microscopy 295-6
plane strain 160, 165-6, 226, 229, orientation 327-9
262-3 particle size effects 389
volume change 181 physical ageing 319
yield in polystyrene 156 scission 320
Configurational relaxation 48 shear deformation zones 313,
Conformational changes, methacrylates 319-21
133 small-angle x-ray diffraction 297-8
Conformational motions 49 solubility parameter 330
Conformational transitions freezing out, solvent diffusion 330
56 strain rate 321-7
Considere criterion for necking 221-3 stress concentration at craze tip,
Gaussian and Langevin models 222 308-9
Contour length stress conditions 295
between entanglements 311 surface energy 316, 318, 327
end-to-end distance 311 temperature 321-7
Core-shell morphologies in ABC termination 302
triblock copolymers 487-9 toughness 295
Crack tip plastic zone 344 Creep 12
Crankshaft mechanism 128 compliance 140
Crazing 25, 295-341 poly(vinyl chloride) 139
chain pull-out 317 toughened polymers 396
compression 165 Criteria for yielding 172
craze breakdown 334-7, 430 Crosslinking 331
craze initiation 301-2 crazing and shear 332
craze midrib 309 disentanglement 333
craze widening 305 suppression of crazing 322
craze-tip structure 300 toughened thermosets 406
cross-linking 322 thermoplastics 259
cross-tie fibrils 299
dilatational component 297 Defects, craze initiation 301
dilatometry 386 Deformation
Dugdale model 307-8 adiabatic and isothermal 246
entanglement density 311 affine 70
extension ratio, 299, 312-13 calorimetry 183-4
fibrillar structure, 297-8 entropy 257-8
fracture toughness 295 thermal effects 242-54
initiation and growth 301-7 Densified glasses, pressure dependence
initiation by defects 301 125
at interfaces 428-31 Density fluctuations 106
mechanism of reinforcement 429-31 Diblock copolymer morphology 458-
meniscus instability 303-6, 385 62
micromechanics 307-10 Dielectric relaxation studies 136
Index 501

Differential scanning calorimeter (DSC) Enthalpy


traces changes after ageing 235-6
dependence on pressure 127 dependence on pressure 126
glass transition temperature 88 Enthalpy relaxation 95-9
polystyrene 127, 184, 236 differential scanning calorimetry 95-7
Diffusion Krauzmann temperature 97
anomalous 78 mathematical expression 96
coefficient 74 relaxation times 98-9
configuration partition function 78 Entropy 18
free volume theory 77 deformation 257-8
jump events 76 Epoxy resins
mechanism 75 volume response on quenching 146
penetrant 73 yield 174, 178
solvent crazing 330 Equation of motion 38
transition state models 77 ESIS protocol 351
Dilatometric measurements 117 Essential work tests 356-7
Direct modelling of a glass 57 Excluded volume 43-4
Dislocations Extensibility, maximum 27-8
block copolymer morphology 473- Extension ratio in a craze 299
9 Eyring rate theory 188-90
model for yield 200-2 rubber toughening 377
Divinyl benzene polymers, modulus and
density 11 Failure modes of polymer interfaces 428
Ductile fracture 219 Fibre modulus 10, 11, 14
Dugdale model for craze stress profile Fibrillar structure of crazes 297-8
307-8 First-order transition 6
Dynamic mechanical thermal analysis Flexible polymers 23
(DMTA),366 Flory-Rehner model 271
Force fields 39
Ehrenfest relations, 116-17, 119-23 Forward recoil spectrometry 423
Elastic fracture mechanics 346 Fracture
Elastic moduli 9 definition of 343
relationship between 161 ductile 219
variation with temperatun: 10 energy 17
Energy minimization 35 high temperatures during 214, 253-4
Energy storage for yield and moderate large deformations 215
strains 236-7, 243-5 melting 214
Entanglement density 311 tests 349
calculation 22, 23 toughness 344
measurement 22 Fracture mechanics 343-62
Entanglement length 35, 55 essential work tests 354-6
interfaces 422, 427 linear elastic 349
Entanglements 71, 80 test geometries 348
density 19, 21, 23 Fracture strength
limited extensibility 18 effect of backbone bonds 15
molecular weight 215-16, 311 crack length 15-17
strain recovery 17, 18 defects 15, 16
variation with temperature 273-4 theoretical 14
502 Index

Fracture toughness 346 related to other transitions 6


ABS 358 as a second order transition 4
amorphous nylon 359 thermodynamics 4-6
compact tension 348 thin films 444
definition 344 transformation 52
determination of 350-1 Glass transition temperature 2, 86, 451
entanglement density 405 effect of cooling rate 3, 4, 87
ESIS protocol 351 crystallization temperature, 87
essential work tests 356-7 volume changes 3, 4, 87
initiation 347, 349 Glassy polymers
J testing 351-4 chemical formulae 2
loading system 347 nodular structure 9
measurement of interfacial strength small-angle x-ray scattering 8, 9
416,417,426-7,433 wide-angle x-ray scattering 6-8
phenolic resins 356-7 Grain boundaries in block copolymers
polycarbonate 356 473-9
poly(methyl methacrylate) 356 Griffith fracture 344
polystyrene 356 Gyroid and perforated layers 468-73
poly(vinyl chloride) 360
resistance curve 346 Harmonic valence angle potential 41
strain hardening 364 Heat
temperature effects 351 conversion of mechanical work into
test geometries 348 245
three-point-bend 348 of deformation 219
toughened poly(phenylene oxide) 359 during high-strain deformation 246
Free volume 105-16 Henry's law 74
changes in compression 235 High-impact polystyrene (HIPS) 379-
change during deformation 182 85
changes at high strains 237 High-strain deformation, heating during
distribution of 110 246
photochromic probes 107-13 Hourglass test-pieces 225
positron annihilation lifetime Hydrostatic pressure 45
spectroscopy 113-16, 182
positron annihilation mechanism 114 Impact 239-41, 353-4
yielding 181-2 Inherent flaw size 16
Inherent flaws 16
Gas solubility 74 Interfaces 413-50
Gaussian equation areal density of crossing chains 428
application to experiments 262-5 chain conformations 442
strain hardening 260- 7 crazing 428-31
Gaussian and Langevin models 222 critical micelle concentration 424
Gaussian modulus 19,27 entanglement length 422, 427
Glass transition 2-6 equilibrium 414
cooling rate 52 failure modes 428
density relaxation 53 forward recoil spectrometry 423
experimental and calculated 59 free surface 416
Gibbs-DiMarzio model 97 incompatible polymers 415-22
pressure dependence 118-24 length scales of interactions 414
Index 503

polymer-inorganic 415 Meniscus instability in crazing 303-6


polystyrene-poly(methyl Mesoscopic simulation 80
methacrylate) 416-20 Microphase separation 453-6
random copolymers 434 entropy penalty 453
reactive grafting 432 Landau mean-field theory 453
reinforcement with block copolymers phase diagram 454
422-35 self-consistent field theory 455
toughness 416-17 Microstructure of glassy polymers 9
Internal energy changes 244 Mode coupling theories 81
Internal stress in densified glasses 127 Molecular distortion 185
Intramolecular potentials 40 Molecular dynamics 36--7, 207-8
Isothermal deformation 246 codes 81
density changes 207
J testing 351-4 problem of time-scale 80
ABS 358-9 simulations 34
impact 353-4 temperature effects 207
measurement 353 Molecular mechanics
crazing 310-19
Kohlrausch--Williams- Watts equation local shear strains 205
52, 191 polycarbonate 205
Kraton morphology 463-5 structure of amorphous
Kuhn length 21 polypropylene 202
yield 202
Langevin model for strain hardening polypropylene 203-4
271-3 Molecular models of yield 196-200
Langevin treatment 19 argon model 198-200
Large deformations dislocation model 200-2
fracture 215 Robertson model 196-8
in processing 215 Molecular network 18
reversibility above ~ 217 Molecular weight 21
universitality 217-18 post-yield deformation 215-17
Lennard-Jones potential 41 toughened polymers 405
Linear elastic fracture mechanics 349 Monte Carlo sampling 35-6
Linear viscoelastic models 190 Multiblock copolymers 483-7
Local energy approximation 47
Localized deformation 214 Natural draw ratio 26, 266-7, 365
Necking
Mean square end-to-end distance 35- condition for 265-6
49 formation 162,220-3
Mechanical deformations 61 temperature distribution 249
Mechanical properties Neutron reflectivity 419, 436
effect of loading time 12 Nominal stress 161
physical ageing 138-46 minimum 265-6
Mechanical tests, types 226 Non-bonded interactions 39
Mechanical work, conversion into heat Notches in toughened polymers 393-4
245-6 Nucleation of crazing 302
Melt viscosity 20 Nylon, amorphous
Melting during fracture 214 J testing 359
504 Index

Optical microscopy, crazing 295-6 Plane strain compression 160, 165-6,


Orientation 226,229,262-3
effect on crazing 327-9 Plastic deformation 13
effect on extension ratio 328-9 Plastic instability 25, 27, 218-23, 238
Oriented polymers shear bands 238
large strains 229 in tension 238
polycarbonate compression tests 231 Plastic zone at crack tip 344
poly(ethylene terephthalate) filaments Plateau modulus 19
230 Poisson's ratio 160
Poly(ethylene terephthalate) (PET)
Particle-matrix adhesion 396 deformation of spun yarns 231
Penetrant diffusion 73-9 isothermal and adiabatic extension
Periodic boundaries 47 246
Persistence length 28, 67 J testing 359-60
Phase diagram for microphase thermal fracture 253
separation 454 Poly(methyl methacrylate) (PMMA)
Phenolic resins, fracture toughness chemical repeat unit 2
356-7 dependence of the enthalpy on
Photochromic probes 107-13 pressure 126
Physical ageing 136-46 fracture toughness 356
p-relaxation for different polymers 137 free volume 110
p-relaxation in polystyrene 136 interface with polystyrene 416-20
p-relaxation and thermal history 137 reversible heating effect 243
chain conformations, 241-2 rubber toughening 377
crazing 319 secondary relaxations 131-3
creep compliance 140-1 shear zones 178
creep in poly(vinyl chloride) 139 true stress-strain curves 228, 271, 276
double logarithmic shift rate 143 uniaxial compression 195
dynamic mechanical thermal analysis Poly(phenylene oxide)
142 crazing 309
epoxy resins 146 J testing of toughened 359
impact strength 240 Poly(vinyl chloride) (PVC)
mechanical properties 138-46 chemical repeat unit 2
molecular mobility 138 creep 139
poly(ethylene terephthalate) 234 fracture toughness 360-1
effect on properties 144 Gaussian plots 262
rejuvenation 145, 234 thermal fracture 253
rejuvenation by deformation 176 yield criteria 173
shift parameters 141 yield strength 223
strain softening 232-7 Polycarbonate (PC)
effect on strain softening 175 chemical repeat unit 2
temperatures close to ~ 143-4 deformation zones 314
time scales 143 fracture toughness 356
viscoelasticity 139 free volume 111
volume changes 241 impact strength 240
yield stress 232, 234 natural draw ratio 26
effect on yield stress 175 neck formation 220
Pivot algorithm 43 preorientation in compression 230
Index 505

rubber toughening 399 interface with poly(methyl


secondary relaxations 129-31 methacrylate) 416-20
simple shear stress-strain curve 167 melting during fracture 214
strain cycling 176 effect of physical ageing 156
temperature distribution in a neck secondary relaxations 133-5
249-50 shear bands 179, 218
tensile stress-strain curves 169-71 storage modulus 20
true stress-strain curves 275 transmission electron microscopy of
uniaxial compression 171 crazes 298, 300, 320, 335, 380
Polyethersulphone (PES) viscosity 21
chemical repeat unit 2 volume relaxation 90--4
deformation 324-5, 333 Positron annihilation lifetime
Polyethylene spectroscopy (PALS) 113-16, 182
amorphous, 55, 62-6 Post-yield behaviour 65. 213-93
chain configuration 71 Preorientation
Considere construction 222 effect on crazing, 327 --9
cooling curves 56-8 effect on post-yield behaviour, 229-
density 72 32
effect of strain and trans conformers Pressure dependence
70, 72 densified glasses 125
heating curve 58 DSC traces 127
models of 56 enthalpy 126
percent trans conformers 57 glass transition 118-24
persistence length and trans internal stress 127
conformers 67-8 yielding 171
strain hardening 66-9
strain recovery 72 Radius of gyration 44- 5
stress-strain curves 62-3, 222 alkane chains 47
tube model 70 Random copolymers for reinforcement
yield stress 64 of interfaces 434
Polymer-solid interfaces Random statistical chains 258
chain conformations 436 Reactive grafting 432
grafted chains, 434-9 Rejuvenation, physical ageing 145
Polystyrene Relaxation of trans states 50
autoadhesion 448 Relaxation times
block copolymer morphologies activation energies 102
458-63 distribution of 105, 191
compressive yielding 155-6 'fragile' melts 102-3
crazing 298, 300, 309, 320, 322, 325, non-Arrhenius melts 102-3
327, 331 Resistance curve 346
crosslinking 331 Reversible thermoelastic effect 242
deformation calorimetry 183 Rigid polymers 23
deuterated 421 Robertson's model of yield 196-8
DSC traces 127, 184,236 Rotational states
elongation to break 216 increased trans conformers after yield
fracture toughness 356 184-5
free volume 110 isomeric 34
high-impact 379-85 Rouse-like models 54
506 Index

Rubber elastic 18, 22, 25 Sample size effects 46-7


Rubber particles Scission crazing 320
a-transition 406-7 Second-order transition 4, 6
cavitation of 371-5, 382, 386 Secondary relaxations 88, 128-38
critical volume strain 374 fJ and y peaks 128-36
cross-linking of rubber 403 b-relaxations 134
dilatation bands 378-9 crankshaft mechanism 128
dilational stresses 371-2 effect of ageing upon 135-8
effective volume strain 376 localize main-chain motion 128
extension ratio 380, 383 merging of a- and fJ-relaxations 135
fracture initiation 383 methacrylates 131-3
morphology 367-8, 402-3 phenyl group rotation in polystyrene
poly(vinyl chloride) 402 134
polycarbonate 399 phenyl group stacking in polystyrene
polystyrene 402 134
shape 369 phenyl rings in chain, 129-31
size 368, 399-402 side groups 128-9
stress concentration 370 Segment mobility near free surface
structure 368 446-7
transmission electron microscopy Self-consistent field theory 424
367-70 microphase separation 455
volume fraction 367-8 Self-oscillation during drawing of films
Rubber relaxation behaviour 406-8 and fibres 251-3
Rubber toughening, 363-412 Shear deformation bands (zones) 313
blends 366 competition with crazing 319-21
cavitation 381-2, 388, 394-5 criteria for formation 180
crazing 379-85, 386, 389 diffuse shear zones 176-7, 179
creep 396 epoxy resins 178
deformation behaviour 391 extension ratio 314
DMTA 366 intense shear bands 179
effect on modulus 370 nucleation 162
Eyring equation 377 poly(methyl methacrylate) 176-7
high-impact polystyrene 379-85 polycarbonate 178
matrix cross-linking 406 polystyrene 179, 218
matrix molecular weight 405 effect of strain rate 178
mechanisms 369-86 Shear modulus 161
miscible blends 366 Shear yielding 322
notches 393-4 Eyring equation 377
particle-matrix adhesion 396-402 modulus ratio 187
poly(methyl methacrylate) 377 molecular weight 323-5
polycarbonate 399 effect of pressure 375
rubber content 394-6 rubber toughening mechanisms
rubber relaxation 406-8 375-8
shear yielding 375-8 shear modulus 186-7
small-angle x-ray scattering 384 effect of temperature 187
strain rate effects 392 von Mises criterion 375
temperature effects 391-4 yield envelope 377, 387
thermosets 406 Simple shear 160, 166
Index 507

measurement with polycarbonate 167 Thermal fracture 253


Small-angle neutron scattering 185, 419 Thermoelastic effect 2423
block copolymer morphology 457, Three-dimensional deformation
462,464,470,474 equations 278-80
Small-angle x-ray scattering load-displacement curve 289
block copolymer morphology 457, prediction of necking in tension
460,463,466 280-9
crazing 297-8 variation of strain in neck 286-8
glassy polymers 8 Three-dimensional modelling 268-77
Solubility parameter, crazing 330 elastic components 269
Specific heat capacities 96-7, 105 model constituents 269
Specific volume, combined effect of Three-point-bend 348
pressure and temperature 119 Time-displacement relaxation 79
Stepwise cooling 53 Torsional potential 41
Strain Transition state models of diffusion 77
recovery 72 Transmission electron nllcroscopy
types 160 block copolymer morphology 456- 7
Strain hardening 25-7, 29, 219, 271 complex phases 472
definition 164 diblock copolymers 459, 462, 472
Gaussian equation 260- 7 multiblock copolymers 482- 7
Langevin model 271- 3 staining agents for block copolymers
Strain rate effects 456
crazing 321-7 triblock copolymers 463, 465
rubber toughening 391-4 Triblock copolymer morphology 463- 7
yielding 169-70 kraton 463-5
Strain softening transmission electron microscopy 463
dislocation loops 201 True strain 223-9
local elastic strain energy 194-5 softening 218,227-8,232
model 270-1 True stress 161, 223-9
viscoelastic interpretation 192 decline after yield 164
Strand density 333 True stress-strain curves 227, 228, 275
Stress concentration at a craze tip 308- Truncated potential 44
9 Tube theory 24
Stress intensity factor 17
determination of 350 Ultimate shear strength 186-8
Stress relaxation 12 Uniaxial compression 160, 165-6
Surface energy for crazing 316. 318, avoidance of crazing 165
327 conditions for uniform deformation
165
Temperature poly(methyl methacrylate) 195
distribution in a neck 249 Uniaxial tension 60-5
effect on crazing 321-7 constant local strain rate 164
effect on rubber tougheneing 391-4 crazing and fracture 162
effect on yielding 168 definition of terms 159
Theoretical strength 14, 365 neck formation 162
Theoretical yield stress 187 shear bands 162
Thermal activation of yield 188 stress-strain curve 163
Thermal effects in deformation 242-54 true and nominal strains 159
508 Index

Uniform deformation 28, 29 molecular models 196-200


in tension 223 molecular simulation of 202-6
United atom model 40 network polymers 174
shear displacement 157
Viscoelasticity 13 strain-softening rate 180
physical ageing 139 test specimen geometries 160
yielding 190 thermal activation 188
Viscoplastic behaviour 270 effect of thermal and mechanical
Volume relaxation 89-95 history 175
coupling model 95 volume contraction 180-1
dilatometric measurements 90-1, 113, Yield criteria
118-19 poly(vinyl chloride) 173
effective relaxation time 93-4 in shear 174
mathematical expressions for 92-3 uniaxial stress 172
relative excess volume 93-4 von Mises criteria 172, 345
Von Mises criterion 345 Yield stress 26, 64, 208-10
shear yielding 375 amorphous polyethylene 64
effect of hydrostatic pressure 171-2
Wide-angle x-ray scattering 6-8 effect of temperature 164, 168-9
Williams-Landel-Ferry equation 54 theoretical 187
Yielding 185-97
Yield 155-212 different specimen geometries 167
activation volume 189 free volume 180-2
copolymers 174 near ~ 168, 170
cross-linked polymers 174-5 stored energy 183-4
crystals and glasses 156-7 strain uniformity 176-80
definition 155 effect of strain rate 169
dislocation model for 200-2 effect of temperature 168
elementary deformation modes 157 viscoelasticity 190
energy storage 236-7, 243-5 Young's modulus 160
enthalpy of activation 188 effect of pressure 172

You might also like