Sei sulla pagina 1di 237

Chapman

& Hall, 2-6 Boundary Row, London SE1 8HN, UK Chapman & Hall
2-6 Boundary Row, London SE1 8HN, UK Blackie Academic & Professional,
Wester Cleddens Road, Bishopbriggs, Glasgow G64 2NZ, UK Chapman & Hall
GmbH, Pappelallee 3, 69469 Weinheim, Germany Chapman & Hall USA, One
Penn Plaza, 41st Floor, New York NY 10119, USA Chapman & Hall Japan,
ITP-Japan, Kyowa Building, 3F, 2-2-1 Hirakawacho, Chiyoda-ku, Tokyo 102,
Japan Chapman & Hall Australia, Thomas Nelson Australia, 102 Dodds Street,
South Melbourne, Victoria 3205, Australia Chapman & Hall India, R. Seshadri,
32 Second Main Road, CIT East, Madras 600 035, India First edition 1995 ©
1995 Chapman & Hall Typeset in 10/12 Times by Interprint Limited Malta.
Printed in Great Britain at the University Press, Cambridge. ISBN 0 412 58430 1
0 412 61440 5 (set) 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
concerning 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 Cataloging-in-
Publication data available 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 vii Preface xi List of contents for Volume 2 xiii
List of contents for Volume 3 ix\ Part One Polymorphism in Polypropylene
Homo- and Copolymers 1 1 Molecular structure of polypropylene homo- and
copolymers 3 B. Monasse and J. M. Haudin 2 Crystalline structures of
polypropylene homo- and copolymers 31 S. Z. D. Cheng, J. J. Janimak and J.
Rodriguez 3 Crystallization, melting and supermolecular structure of isotac¬ tic
polypropylene 56 J. Varga 4 Nucleation of polypropylene 116 A. Galeski 5
Epitaxial growth on and with polypropylene 140 J. Petermann Part Two
Processing-induced Structure 165 6 Higher order structure of injection-molded
polypropylene 167 M. Fujiyama 7 Knit-line behaviour of polypropylene and
polypropylene 167 M. Fujiyama 7 Knit-line behaviour of polypropylene and
polypropylene- blends 205 G. Mennig 8 Welding and fracture of polypropylene
interfaces 227 R. P. Wool

VI Contents 9 Self-reinforcement of polypropylene 273 J. Song, M. Prox, >4.


W'eber and G. W. Ehrenstein 10 Processing-induced structure formation 295 H.
Janeschitz-Kriegl, E. Fleischmann and W. Geymayer Index 345

Contributors Prof. S. Z. D. Cheng Institute and Department of Polymer Science


University of Akron Akron Ohio 44325-3909 USA Prof. G. W. Ehrenstein
Lehrstuhl fiir Kunststofftechnik Universitat Erlangen-Niirnberg Am
Weichselgarten 9 D-91058 Erlangen Germany Dr E. Fleischmann Institut fur
Werkstoffkunde und Priifung der Kunststoffe Montan-Universitat Leoben A-
8700 Leoben Austria Dr M. Fujiyama Plastics Research Laboratory Tokuyama
Corp. 1-1 Harumi-cho Tokuyama City, Yamaguchi 745 Japan Prof. A. Galeski
Centre of Molecular and Macromolecular Studies Polish Academy of Sciences
PL-90-363 Lodz Sienkiewicza 112 ^ Poland

Contributors viii Prof. W. Geymayer Zentrum fur Elektromikroskopie


Technische Universitat Graz A-8010 Graz Austria Prof. J. M. Haudin Ecole
Nationale Superieure des Mines de Paris Centre de Mise en Forme des
Materiaux B.P. 207 F-06904 Sophia Antipolis France Prof. Janeschitz-Kriegl
Institut fiir Chemie, Abt. Verfahrenstechnik Johannes Kepler Universitat Linz
Altenberger Str. 69 A-4040 Linz Austria Dr J. J. Janimak J. J. Thomson
Laboratory University of Reading Reading RG6 2AF Great Britain Prof. G.
Mennig Deutsches Kunststoff-Institut SchloBgartenstr. 6 D-64289 Darmstadt
Germany Prof. B. Monasse Ecole Nationale Superieure des Mines de Paris
Centre de Mise en Forme des Materiaux B.P. 207 F-06904 Sophia Antipolis
France Prof. J. Petermann Lehrstuhl fiir Werkstoffkunde Universitat Dortmund
PO Box 50 05 00 D-44205 Dortmund Germany

Contributors Dr M. Prox Lehrstuhl fur Kunststofftechnik Universitat Erlangen-


Niirnberg Am Weichselgarten 9 D-91058 Erlangen Germany Dr J. Rodriguez
Institute of Polymer Science University of Akron Akron Ohio 44325-3909 USA
Dr J. Song Lehrstuhl fur Kunststofftechnik Universitat Erlangen-Niirnberg Am
Weichselgarten 9 D-91058 Erlangen Germany Prof. J. Varga Technical
University Budapest (BME) Department of Plastics and Rubber Stoczek Str. 2
H-l 111 Budapest Hungary Dr A. Weber Lehrstuhl fur Kunststofftechnik
Universitat Erlangen-Niirnberg Am Weichselgarten 9 D-91058 Erlangen
Germany Prof. R. P. Wool Department of Materials Science and Engineering
University of Illinois at Urbana-Champaign 1304 West Green Street Urbana
University of Illinois at Urbana-Champaign 1304 West Green Street Urbana
Illinois 61801 USA

Preface Crystalline polypropylene (PP) was invented in the early 1950s by


indepen¬ dent groups in the United States and Europe. The commercial
production of PP began in 1957 in the USA and in 1958 in Europe [1]. The
reader will find different data about this issue in this book. On the other hand,
there is no doubt that PP became the winner among the commodity of large-
volume thermoplastics into the group of which polyethylene (PE), poly(viny-
chloride) (PVC) and polystyrene (PS) belong, too. The mean consumption rate
of PP was about 10% per year in the last decade, in given application fields this
value was even higher. The forecast for the future trend of PP use is still quite
optimistic. Which are the contributing factors to this success? The key factor is
related to the versatility of PP per se. This means that the structure and
properties (including processability) of PP can be tailored to requirements.
Modifications of PP can be can be performed in different ways: during the
polymerization (e.g. production of syndiotactic homopolymers, copolymers with
different comonomer content or polymers with narrow molecular mass
distribution), in the reactor (reactor-blends), in compounding (e.g. manufac¬
turing of filled and chopped fibre reinforced grades) or in further separate
processing steps (e.g. wetting of glass mat by PP-melt, manufacturing of textile
composite preforms). The concept of this book on PP was to show this variety.
PP being a stereoregular polymer, exists in different tacticities (atactic, isotactic,
syndiotactic) and in different crystalline forms, which turns out in different
rheological, physical and mechanical properties. This is the right place to
mention that research studies carried out on various topics (nucleation,
supermolecular structure, melting, crystallization, crystalline transformation etc.)
by investigating PP, contributed essentially to the present state of knowledge of
polymer science in general. This can be attributed to the aforementioned
stereoregularity and polymorphism of PP. On the other hand, the
(micro)structural variety of PP appears in the products and parts moulded from
unfilled, filled or reinforced PP grades, as

Preface xii well. Thus, it is necessary to encounter the microstructural


characteristics of PP first. Volume 1 of this set is devoted to the structure and
morphology of PP and to their changes due to processing and is split into two
sections: Polymorphism in PP homo- and copolymers and Processing-induced
struc¬ ture. In the first section the molecular and supermolecular structures,
including the development of the latter are surveyed. Outlining this section, it is
intended to show how the morphological features of PP can, or could, be
intended to show how the morphological features of PP can, or could, be
exploited further (e.g. self-reinforcement, epitaxal growth, oriented nu- cleation).
Contributions to the section Processing-induced structure set out the interrelation
between the processing-induced morphology and the mech¬ anical performance.
This is a key issue because designers are often not aware of these effects, which
may be both beneficial or detrimental. Targets of Volume 1 are to demonstrate
the microstructure-property relationships and their changes due to processing in
neat PP and to point out both general and peculiar characteristics of PP as a
representative of semicrystalline polymers. Due to this concept I hope that this
book will be an invaluable source to all people working on different fields of
synthesis, processing and application of semicrystalline polymers. This book is
designed to serve as guide-lines for property improvement, upgrading and
engineering use of PP. It is for this reason that such important fields of PP
application as textile fibres, packaging films and foils etc., are not covered.
Nevertheless, I am sure that the concept of this book meets the expectations of
the reader and thus elder books on PP will remain on the book shelves of the
libraries. Comprehensive referencing included with most of the contributions to
this book, helps to reach this target. The reader will be surprised to see that many
references are dated back to the 1960s. These ‘evergreen’ papers indicate the
high quality level of the works done in the ‘pioneering stage’ of PP
development. It is noteworthy that during editing I did not try to ‘harmonize’ the
conclusions and opinions of the authors (though sometimes I inserted a
footnote). The Romans used to say: Varietas delectat (variety is delightful) and I
share this opinion. On the other hand, I believe that subjects of debate will be
settled by time, since that is the task of the science and the real challenge of the
research. Thanks are due to the contributors, who kindly agreed to cooperate on
this ‘venture’, for their efficient work and for their patience due to the time delay
caused by the transfer of this book from Elsevier Applied Science to Chapman &
Hall, who published this book in a short time. Jozsef Karger-Kocsis
Kaiserslautern and Budapest A.m.D.g. Reference 1. H.R. Sailors and J.P. Hogan
(1981) History of polyolefins, J. Macromol. Sci. Chem., A15, 1377-1402.

Contents for Volume 2 1 Manufacturing and properties of polypropylene


copolymers P. Galli, J. C. Haylock and T. Simonazzi 2 Primary spherulite
nucleation in polypropylene-based blends and copolymers Z. Bartczak, E.
Martuscelli and A. Galeski 3 Polypropylene alloys and blends with
thermoplastics L. A. Utracki and Μ. M. Dumoulin 4 Structure and properties of
polypropylene elastomer blends E. Martuscelli 5 Orientational drawing of
polypropylene and its blends S. V. Vlassov and V. N. Kuleznev 6 Thermoplastic
elastomers by blending and dynamic vulcanization A. Y. Coran and R. P. Patel
elastomers by blending and dynamic vulcanization A. Y. Coran and R. P. Patel

Contents for Volume 3 1 Particulate-filled polypropylene: structure and


properties B. Pukdnszky 2 Processing and properties of reinforced
polypropylene A. G. Gibson 3 Fiber orientation prediction in injection molding
T. Matsuoka 4 Microstructural aspects of fracture in polypropylene and in its
filled, chopped fiber and fiber mat reinforced composites J. Karger-Kocsis 5
Glass mat reinforced polypropylene L. Berglund and M. Ericson 6 Some wetting
and adhesion phenomena in polypropylene composites J. J. Elmendorp and G. E.
Schoolenberg 7 Manufacturing methods for long fiber reinforced polypropylene
sheets and laminates D. M. Bigg 8 Thermoforming of unidirectional continuous
fiber-reinforced polypropylene laminates and their modeling R. Scherer 9
Fracture performance of continuous fiber reinforced polypropylene B. Z. Jang 10
Interfacial crystallization of polypropylene in composites M. J. Folkes

Part One Polymorphism in Polypropylene Homo- and Copolymers

ί Molecular structure of polypropylene homo- and copolymers B. Monasse and


J. M. Haudin 1.1 INTRODUCTION As early as 1869 propylene was
polymerized by Berthelot by reaction with concentrated sulphuric acid. The
resulting viscous oil, at room tempera¬ ture, did not exhibit interesting
properties for industrial applications. Its industrial importance results from the
appearance of crystalline high molecular weight polypropylene (PP), which was
first polymerized in 1955 by Natta et al [1] from organo-metallic catalysts [2]
based on titanium and aluminium. The resulting semi-crystalline polymer has
strong mechanical properties which explain its rapid industrial development
from its introduction into the market by Hoechst AG in 1965. A wide range of
PP homopolymers, and random and block copolymers with various weight
distributions were developed for numerous and versatile applications [3]. PP is
widely injected to form parts, especially components such as bumpers and
dashboards for the automotive industry. Packagings are obtained by blow
moulding. Extrusion leads to profiles, pipes and sheets. These sheets can be
uniaxially or biaxially stretched to obtain film tapes or oriented polypropylene
(OPP), respectively. Textile fibres and non-woven fibres where high molecular
weight is of key importance [4] are obtained by melt spinning. A single PP is not
able to optimize the processability and properties of final products. For each
application, it is necessary to control the polymerization parameters and
subsequently Polypropylene: Structure, blends and composites. Edited by J.
Karger-Kocsis. Published in 1995 by Chapman & Hall, London. ISBN 0 412
58430 1
4 Polymorphism in polypropylene homo- and copolymers to characterize the
polymer chains. This polymer characterization is the aim of this chapter. 1.2
MAIN CHARACTERISTICS OF A POLYMER CHAIN It is necessary to
define the parameters that characterize a PP chain before giving any description
of the technical methods available. Three main levels are necessary to describe a
polymer material: 1. the molecular weight, or degree of polymerization, and its
distribu¬ tion; 2. the chemical composition including chain defects and
copolymerization effect, its content and co-monomer type; 3. the stereoregularity
of a homopolymer sequence. 1.2.1 Molecular size The number and the type of
monomers included inside a polymer chain is the direct result of the
polymerization route. A PP homopolymer consists of chains with only propylene
segments: (CH2—CHCH3)n. For a PP chain the degree of polymerization n and
the molecular weight M(n) are directly linked by the molecular weight of the
propylene monomer Mpr (equation 1): These definitions are sufficient to
characterize the size of the PP chain. For an ethylene-propylene copolymer, the
mole fraction β of propylene in the polymer chain is also needed (equation 2):
The chains inside a material resulting from the polymerization process are not
identical either by composition in a copolymer or by size. The size distribution
can be exhibited by the number N-x of molecules with a weight as a function of
the molecular weight M{. This molecular distribution can be partly reduced to
mean values with various momenta. The first category leads to a number-
average molecular weight Mn defined as: M(ri) = nMpr = n x 42 g/mol (1) Μ(ή)
= (1—β)ηΜ6ΐγι+βηΜρτ (2) ^ weight of molecules number of molecules T (3)
where nx is the number fraction of molecules with a molecular weight Mt.

5 For polymers with a homogeneous molecular size, the Mn, Mw, and Mz values
would be identical. In the more usual situation the molecular size is
heterogeneous and Mn < Mw< Mz. Some experimental methods are sensitive to
the dimension of polymer coil in solution. These methods are distinguished by
whether the contribution of each molecule to the measure¬ ment is proportional
to the number of molecules or to a power of molecular weight. The number-
average molecular weight can be deduced from freezing point depression,
vapour pressure lowering and osmotic pressure of dilute solutions [5]. Light
scattering of such solutions is usually applied to measure the weight-average
molecular weight of polymers. The z-average molecular weight is measured by
ultra-centrifugation. All these average values can be deduced from gel
permeation chromatography (GPC), i.e. size exclusion chromatography (SEC),
which measures the molecular weight distribution. This more versatile method
has largely supplanted previous ones which gave only mean values. 1.2.2
Chemical structure Polymerization of the non-symmetrical propylene molecule
Chemical structure Polymerization of the non-symmetrical propylene molecule
CH2 =CHCH3 leads to three possible sequences (Figure 1.1). Figure 1.1
Isomerism for positions in polypropylene: (a) head-to-tail; (b) head- to-head; (c)
tail-to-tail. Main characteristics of a polymer chain where Wj is the weight
fraction of molecules with a molecular weight The z-average molecular weight
is defined by: The weight-average molecular weight is: (4) (5) (a) 0>) (c)

6 Polymorphism in polypropylene homo- and copolymers The steric effect of the


methyl group highly favours the head-to-tail sequence, which gives a high
chemical regularity of the PP chain. The occurrence of head-to-head and tail-to-
tail sequences induces chemical defects along the PP chain. A random
copolymerization of propylene with a co¬ monomer also induces desired
chemical defects along the chain. Usually ethy¬ lene is used as the co-monomer;
this can be included randomly with various contents inside the polymer chain.
Ethylene content strongly modifies the physical properties of the polymer chain.
The change in the possible conforma¬ tions modifies the crystallization, the
thermal stability [6], the stiffness and the elongation to break of the material.
Ethylene-propylene rubbers (EPR) result from a high ethylene content (at least
10 mol%) randomly distributed, which hinders ethylene and propylene
crystallization. A lower ethylene content (no more than 5 mol%) only reduces
the ability of PP to crystallize [7]. Crudely, a two-step polymerization with two
different monomers produces a block copolymer. Usually ethylene is also used
as co-monomer, resulting in ethylene-propylene block copolymers. This
sequenced copolymerization produces only tiny changes inside each polymer
domain with respect to similar pure polymers. Consequently each polymer
domain can crystallize separately from the other. In elastomeric terpolymers
(ethylene-propylene diene mono¬ mers; EPDM), non-conjugated diene
monomers are incorporated into the polymer chains built up from ethylene and
propylene units. 1.2.3 Stereo-isomerisms in polypropylene chains PP and
ethylene-propylene copolymers (EPM) are exclusively formed from carbon and
hydrogen. Elementary analyses are not able to distinguish these various
polymers. Only the diene links can be specifically characterized by reactive
agents. The chemical composition of all these polymers is more accurately
characterized by physico-chemical methods such as nuclear magnetic resonance
(NMR) or infra-red (IR) spectroscopy. These methods give a quantitative
measurement of chemical groups (CH3, CH2, CH, CH=) and the mean
composition of the polymer. Moreover, they are sensitive to the neigbourhood of
chemical groups. It is also possible to define the chemical sequences along the
chain and to distinguish random and block copolymers. NMR of polymer
solutions measures directly the coup¬ ling of neighbour chemical groups. IR
solutions measures directly the coup¬ ling of neighbour chemical groups. IR
spectroscopy is sensitive to the molecular conformation partly induced by the
vicinity of chemical groups. The methyl branching implies an asymmetrical
carbon C* (tertiary carbon) in the propylene group (CH2—C*HCH3). Two
stereo-isomeric configur¬ ations result from asymmetrical carbons. Ziegler-
Natta catalysts were developed to select one of these configurations during head-
to-tail polymer¬ ization of propylene monomers. It is useful to represent the
polymer in the zigzag planar conformation to exhibit the three typical stereo-
configur¬ ations: isotactic, syndiotactic and atactic (Figure 1.2).

7 Main characteristics of a polymer chain Figure 1.2 Stereo-configurations of


propylene sequences: (a) isotactic; (b) syn- diotactic; (c) atactic. Isotactic PP, in
which all the methyl groups are on the same side of the zigzag plane, results
from polymerization of only one isomeric configuration from propylene
monomer. Syndiotactic PP is defined by methyl groups arranged alternatively on
either side of the zigzag chain, and is obtained by alternative addition of the two
stereo-isomeric configurations from propy¬ lene monomer. These two stereo-
configurations are specific of Ziegler- Natta’s catalysis. Atactic PP is formed
with a sterically random sequence of methyl groups. This polymer, which is
incapable of crystallization, can be obtained by a classical radical reaction. These
various stereo-isomeric configurations can be distinguished and quantitatively
analysed by 13C NMR analysis of polymer solutions by the difference of nuclei
coupling with neighbour methyl of the various configurations. These stereo-
isomeric configurations also induce stable, specific overall conformations of the
polymer chain: helix 2 x 3/1 for the isotactic PP [7], helix 4 x 2/1 for the
syndiotactic PP [8] and a random conformation for atactic PP. The stable regular
conformations of isotactic and syndiotactic chains promote crystallization by a
parallel organization of chain segments issued from the same chain or from
chains in the vicinity. Three crystalline forms are known for isotactic PP:
monoclinic a phase [7], hexagonal β phase [9] and triclinic y phase [9]. An
orthorhombic form was observed for syndiotactic PP [10]. For each crystalline
form, thin crystals are formed for kinetic reasons, connected to each other by the
residual amorphous phase. Various technical methods (IR, solid state NMR) are
able to measure the chain conformations, while X-ray diffraction and differential
scanning calorimetry (DSC) establish the crystalline phases and their stability,
re¬ spectively. Furthermore, thermo-mechanical analysis (TMA) measures the
critical conditions (temperature and frequency) for deblocking the move¬ ment
of chain segments or groups. These critical conditions depend on integration of
the segments in the amorphous or crystalline phases.
8 Polymorphism in polypropylene homo- and copolymers 1.3 MOLECULAR
WEIGHT DISTRIBUTION OF POLYPROPYLENES Industrial development of
PP during the 1960s was concomitant with the emergence of GPC. This method
is able to give the molecular weight dis¬ tribution of PP, being heterogeneous by
weight, within a short time. This is why this method is more widely used than
direct methods such as light scattering, intrinsic viscosity or osmometry. All of
these methods need a pure solution of all the molecules of the PP sample. PP,
and especially the higher molecular weights, are difficult to dissolve and are
sensitive to thermal degradation. A special procedure is needed for PP
dissolution. GPC is a part¬ icular form of liquid chromatography where any
specific interaction solution¬ stationary phase must be avoided. This SEC selects
the polymer by molecular weight, eluting the highest molecular weights first.
That stationary phase is characterized by a pore size where only smaller
molecules can access. The smaller molecules diffuse more deeply inside the
pores, which explains the molecular size effect on eluting volume. This PP
fractionation must act for a wide molecular weight range under high
temperature, then a calibration is used to convert eluting volume into molecular
weight. This calibration requires the use of monomolecular fractions. Three main
steps must be con¬ trolled to obtain an accurate determination of molecular
weight distribution: • pure dissolution of all PP chains; • selection of all PP
chains by molecular weight; • conversion of elution volume into molecular
weight. 1.3.1 Dissolution of Polypropylene A pure solution of PP is necessary
for all the molecular weight measure¬ ments. Isotactic PP cannot be dissolved in
any solvent at room temperature [11] , whereas atactic PP solutions are easily
formed with various solvents [12] . This difference seems to be induced by the
resulting high crystallinity of iso tactic PP [12]. As a consequence, dissolutions
are usually conducted at high temperature (130-170 °C) [11-19] with
halogenated aromatic com¬ pounds (mono-, di- or trichlorobenzene) [13-19] or
decaline [11, 12]. High temperature favours dissolution but promotes thermo-
oxidative degrada¬ tion, which certainly acts on tertiary carbons, therefore PP is
more sensitive than polyethylene [12]. Because of this specific sensitivity the
dissolution temperature must be limited to between 140 °C and 145 °C and the
solvent used must contain antioxidation additives [12-19]. Recently a
cyclohexane- decaline solvent system was developed to reduce the temperature
of the solution to 70 °C [11, 12, 20]. This solution is formed by a short-term
dissolution of PP in decaline at high temperature (160 °C), then the long-term
solution is formed by mixing with a larger amount of cyclohexane at 70 °C [11,
12, 20]. This solution seems to be stable at 60 °C for weeks

Molecular weight distribution 9 [11]. Whatever the polymer the higher


Molecular weight distribution 9 [11]. Whatever the polymer the higher
molecular weights are always more difficult to dissolve than lower species. A
dissolution that is not dependent on molecular weight is necessary to form a PP
solution representative of the material. Usually the molecular weight spectra of
PP samples are very wide: 103g/mol^M^ 107 g/mol, i.e. 24^n^2.4xl05 PP
segments [11]. This broad molecular spectrum associated with the low solubility
of PP leads to potential difficulties in forming a representative solution. At least
30 hours are necessary to obtain a stable, aggregate-free solution of a PP with
large molecular weight distribution in trichlorobenzene at 145 °C [19].
According to GPC measurements these stable, aggregate-free solutions exhibit
higher mean values than solutions obtained with shorter dissolution time. This
effect is interpreted as an excessive delay to destroy high molecular weight
aggregates [19]. No molecular degradation under these conditions was detected
until 55 hours had elapsed [19]. Such a long dissolution time is critical to obtain
accurate mean values for Mw and A?z. No specific conditions for dissolution are
required in the case of PP copolymers. 1.3.2 Molecular weight selection The
measuring methods select polymer chains by their molecular weight. Two
methods are used: selective fractionation by molecular weight and by the
dimension of solvated chains. Polymer fractionation Polymer fractionation is
based on the decrease of solubility of a poly¬ molecular sample with increasing
molecular weight. The Flory-Huggins treatment of polymer solutions [21-23]
approximates the partial molar free energy, AGl5 of mixing a solvent with a
polymer by: where the volume fraction of polymer v2 at the absolute
temperature T depends on gas constant R, the number-average degree of
polymeriza¬ tion DPn and the Huggins interaction coefficient, χί9 between
polymer and solvent molecules. The Huggins coefficient χΐ9 deduced from the
activity of the solvent, also depends on the temperature approximately according
to where A and B are constants. The variation of with solvent activity and
temperature leads to several methods of fractionation using a mixture of (6) (7)

10 Polymorphism in polypropylene homo- and copolymers solvents and/or a


thermal gradient. Equation 8 defines the critical condition Xi.crit for phase
separation Upon phase separation a solvent-rich phase a and a polymer-rich
phase b are formed. The distribution between the two phases of a polymer with a
polymerization degree DP is — = exp(—<5DP) (11) vb Parameter δ is a
function of the average degree of polymerization, of the polymer concentration
in both phases, and increases with Huggins coeffi¬ cient χι.Ιη conclusion, any
molecular weight is always present in each phase but the selectivity increases
with molecular weight. This method is applied to fractionation of PP [15, 24,
25]. Selection by the dimension of solvated chains The Huggins interaction
25]. Selection by the dimension of solvated chains The Huggins interaction
parameter also acts on the second virial coefficient in absolute molecular weight
measurements (the polymer dimension is derived from the solvated chain) by
osmotic and light scattering methods: (12) where osmotic pressure Π for a
polymer concentration c is directly linked to the number-average molecular
weight Mn and the second virial coefficient is A2. The light scattering equation
for polydisperse molecules is (13) /red(0, c) is the reduced observed intensity of
scattered light at angle Θ. It depends on weight-average molecular weight Mw
and on the second virial coefficient. Ρ(θ) depends on the angle of measurement
and on the dimension of molecules in solution, while k is a constant calculated
from the refractive From equation 7 and (8) (9) (10)

Molecular weight distribution 11 index of the solvent, its concentration gradient


and the wavelength of light. This method has also been applied to PP [26]. The
second virial coefficient is related to the Huggins interaction parameter χ where
p is the density of the polymer and Vs is the molar volume of the solvent. GPC
measurement of molecular weight distribution Contrary to previously quoted
methods, no specific interaction is needed for GPC. It is a relative method. A
calibration curve is necessary to convert the elution chromatogram to the
molecular weight distribution. Two methods can be used: calibration with a wide
ijange of monomolecular-weight samples of PP or a universal calibration [27].
These PP are not directly available by polymerization. The standard PP samples
defined by a given molecular weight must be obtained by fractionation. This
fractionation can be done by preparative GPC [14], by elution chromatography
[15] or by fractional coacervation [21]. These methods induce narrow molecular
dispersion but not monomolecular samples. The polymolecularity index
(polydispersity) Mw/Mn from GPC is higher (Mw/Mn> 1.25) [14, 15] than that
from elution chromatography (A?w/Mn < 1.02) [15]. These samples are used to
deduce elution volume-molecular weight relationships for PP [18] and to check
universal calibration theory [16, 17]. For a wide range of polymers universal
calibration shows that polymers having the same hydrodynamic volume but
different molecular weights can have the same elution time [27]. The
hydrodynamic volume of a solvated polymer chain is measured by its intrinsic
viscosity [η] using the Mark- Houwink relationship where K and a are empirical
constants and Mv is the viscosity average molecular weight. Equation 15
describes a wide range of molecular weights, though large deviations appear at
high [28] and low [29] molecular weights. A wide range of K is described for PP
homopolymers (10_4^K^ 7.4 x 10"4dl/g) [13, 15-18, 30]. A value of the a
exponent was proposed (0.5^α*ζ0.8) [13,15-18, 30] for the same homopolymers,
with a mean value in the range 0.7^a^0.75 [13, 17, 18]. A high value of a is
with a mean value in the range 0.7^a^0.75 [13, 17, 18]. A high value of a is
usually associated with a good solvent [13]. From the universal calibration
model the product [rf\M must be constant whatever the polymer. This calibration
was verified by GPC on PP fractions compared with narrow distributed fractions
of polystyrene (PS) polymerized by an anionic method: [^ρρ]Μρρ = [>/ρ8]Μρ8
[17]. It should be noted here that the hydrodynamic volume of copolymers in
solution depends not only on the molecular weight, but also on the (14) (15)

12 Polymorphism in polypropylene homo- and copolymers co-monomer content


[30]. The universal calibration is assumed to be independent of copolymer
composition [30]. Only the K and a parameters of the Mark-Houwink equation
depend on composition, with limits fixed by polyethylene (PE) and PP
homopolymers [30] (Figure 1.3): Wpe=4.9 x 10-4M°·74 (dl/g) Mpp = 1 x 10-
4M0·78 (dl/g) (16) The calibration curve was defined by the average mole
fraction β of propylene in the copolymer (EP): KEP = KPE(1 - β) + Kpp/? -
2(KpeKpp)1/2 β(1-β) (17) Propylene content was measured by IR spectroscopy
[30, 31]. The results were also discussed using assumptions about PP-PE, PE-
EP-PP and PE-EP blends. In the case of PP-PE blends, the deviation of the
apparent values from the true ones became small when the two polymers had the
same molecular weight [30]. Mark-Houwink parameters for an ethylene-
propylene copolymer of broad molecular weight distribution were defined by an
independent trial and error method [32]: M-, 5.755-4.650(5^^)· (18) where β is
still the mol% of propylene in the polymer and 0.74<a<0.755. This relationship
implies that in a copolymer K and a are dependent on each other. The molecular
weights determined by means of the universal calibration curve seem to agree
well with those determined by solution viscosity, light scattering and osmometry
[32]. From the universal

Molecular weight distribution 13 calibration a different equation was proposed


for the K parameter of a copolymer [33]: KEp — KPE( 1 — 3/? (19) with
propylene molar fraction β. The effect of block or random copolymeriz¬ ation is
not discussed in these references. An alternative universal calibration based on
unperturbed dimensions has also been suggested [34]. The intrinsic coil
dimension is only defined for a Θ solvent and is necessary to check the universal
calibration model [17]. Direct measurements in Θ condition are not suitable with
a solution representative of the PP material. Unperturbed dimensions are derived
from viscosity measurements in good solvents by eliminating the excluded
volume through extrapolation to zero molecular weight [17]. The limiting
viscosity number of a sample in a good solvent is related to that in a Θ solvent
\_η]θ by the viscosity-radius expansion factor οίη: M = (20) where a3 is given
by an approximate closed function of the excluded volume parameter. The
by an approximate closed function of the excluded volume parameter. The
unperturbed dimensions derived depend on the form assumed. Yamakawa’s
modification [35] of the Stockmayer-Fixman equa¬ tion [36] has empirical
support in the range 0<a3<2.5, and was used to extrapolate the M data for PP
samples [15]. ^5= 1.05 Κβ +0.287 Φο-j^M1'2 (21) Here Ms is the molecular
weight of a segment, β0 is the binary cluster integral for a pair of chain
segments, representing the effective volume excluded to one segment by the
presence of another, Ke is the Mark- Houwink constant for Θ condition (Ke
=0.137±0.009 cm3/g [17] for various solvents in Θ condition) and φ0 is the
universal viscosity constant. GPC results are in some cases compared with
fractionation data [15] and direct measurements from osmometry [16, 17] and
light scattering [16-19]. It must be noticed that these comparisons are applied
with various solvents and dissolution conditions which may lead to different
soluble contents and possible disagreements [17, 18]. GPC results should also be
sensitive to sampling frequency as pointed out by computer simulation [37]. At
least 50 points are necessary to reduce to 10 % the relative error on Mw [37].
GPC has been applied to monitor molecular weight variation during PP ageing
[38], grafting of EPDM [39, 40] and formation of EPDM network by inverse
GPC [41]. Low molecular weight materials such as antioxidants and stabilizers
[42] have also been analysed by GPC.

14 Polymorphism in polypropylene homo- and copolymers 1.4 CHEMICAL


COMPOSITION The term ‘microstructure’ is defined in this chapter as the
structure of an individual polymer chain, mainly PP. The resulting
conformations and the mutual arrangements of segments of polymer chains will
be named ‘arrangements’. The crystallization and stability of PP crystals mainly
result from a high stereoregularity of PP segments in the chain. It is necessary,
firstly, to define the chemical structure and periodicity of the polymer chain:
homo- and copolymer PP chains. This chemical periodicity can be broken by
tail-to-tail, head-to-head or by a co-monomer. Secondly, the configurational
isomerism of PP segments must also be characterized in order to control the final
polymer properties. It is known that a high periodicity, isotactic and syndiotactic
contents are necessary to crystallize PP. The chemical sequences depend on the
catalyst and on polymerization temperature [43]. The terminology used to define
the configuration of the chain is that of Carman, Harrington and Wilkes [44],
where S, T and P refer respectively to secondary (methylene), tertiary (methine)
and primary (methyl) carbons. The Greek subscripts refer to the distance of a
given carbon from the neighbour¬ ing methine carbon bearing a methyl group
(Figure 1.4). In addition, the relative configuration of propylene segments is
defined by meso (m) and racemic (r). Stereo-specific polymerization of
defined by meso (m) and racemic (r). Stereo-specific polymerization of
propylene can produce various sequences [45]. Table 1.1 portrays the above
variants under a Fisher projection. Structures Ba, Ca and Da correspond to the
three isotactic regio-irregular structures, homologues of isotactic PP. Variations
in the tacticity would produce a variety of iso tactic regio-irregular structures,
three of which (Bb, Cb, Db) are shown. The 1-3 insertion structures can
effectively appear during homopolymerization of propylene. This structure may
also develop in copolymerization (head-to-head with two ethylene monomers or
tail- to-tail copolymerization with one ethylene monomer). Randomization of
tacticity would produce structure Gb, while Ga results from an isotactic
configuration of PP sequences. Usually an isotactic PP includes at least 95 mol
% of isotactic configuration. At most, the last 5% correspond to all the other
configurations (Table 1.1) Ppp C1I3 t Py+y+ CH3 — CH— CH2— CH— CH2
— CH2— CH2— CH— CH2— CH2— CH2— CH2— CH2— Tpp S„ I Sfty
Spp Sy« VT* S.J* Sp*+ Sy4* sS4*4+ ch3 P + Figure 1.4 Configuration of
primary (P), secondary (S) and tertiary (T) carbons.

Chemical composition 15 A random copolymerization of propylene with a co-


monomer also induces desired chemical defects along the chain in the same way
as 1-3 insertion. Usually ethylene is used as the co-monomer, which can be
randomly included in the chain with various contents. The ethylene content
strongly controls the physical properties of the polymer chain. Actually, the
change in the polymer configuration modifies the possible conformations and
consequently the crystallization, thermal stability, stiffness and elonga¬ tion to
break of the material. EPRs include a high ethylene content (at least 10mol%) of
random distribution, which hinders the crystallization of ethylene and propylene
segments. A lower ethylene content (no more than 5mol%), on the other hand,
reduces only the crystallization ability of PP [6]. In a block copolymer the
behaviour of each polymer sequence is very similar to that of the pure polymers
(homopolymers). Only the connecting zone seems to be affected. Each polymer
sequence can crystallize separately in the respective homopolymers. This fact
explains the role of the coupling agents applied in non-compatible polymer pairs.
1.4.1 NMR analysis of chain configuration These two levels of PP chains can be
characterized by the same methods. Direct assignment of structure and
quantitative determinations are possible by NMR. Secondary methods such as IR
and Raman spectroscopies, calorimetry, X-ray diffraction, and mechanical or
dielectric relaxation can describe the polymer structure quantitatively by
measuring the final ar¬ rangements. The NMR method must be selective with
the assignment of the configur¬ ation of each segment of the polymer chain. The
method must also be sensitive to analyse segments with a very low content, such
method must also be sensitive to analyse segments with a very low content, such
as chain end segments or transition segments in block copolymers. Despite the
apparent simplicity of high tacticity PP, the *H NMR spectra are complex,
especially at H0 =60 MHz [46], because of the strong coupling of the methyl
protons to the a protons, and because the differences in the chemical shifts
between α, β and methyl protons are rather small [47]. Iso tactic and syndiotactic
PP are distinguished because β protons of isotactic PP appear as multiplets near
8.7τ and under methyl doublet at 9.1τ [47]. The methyl protons are markedly
more shielded in the syndiotactic PP [47]. A higher resolution of the spectra
results from the use of a 220 MHz frequency of *H NMR; the spectra are clearly
separated and nearly first order [48, 49]. Polypropylene exhibits 13C shifts from
tetramethyl silane (TMS) in the range (5 = 45.3-48.1 ppm (Saa), 5 = 28.4 — 28.7
ppm (Τββ) and 5 = 19.5- 20.1 ppm (P^) [44, 50-58]. A detailed assignment
shows a shift difference for isotactic (mmm) and syndiotactic (rrr)
polypropylenes: 5Saa(mmm) = 46.2-47.1 ppm, SSM(rrr) =47.1-47.6 ppm,
<5TW(mm) =28.7 ppm, <5TW(rr) = 28.4ppm, <5P0p(mm) =21.1-22ppm,
<5PWrr) = 19.5-20.1 ppm. These main

Table 1.1 Possible secondary structures observed in polypropylene. Isotactic


Syndiotactic ch3 ch3 ch3 ch3 I I I I —ch2—ch—ch2—ch—ch2—ch—ch2—ch
— ch3 I CH3 I —ch2—ch—ch2—ch—ch2—ch—ch2—ch— I I CH3 CHj
Regio-irregular structures ch3 ch3 ch3 ch3 ch3 ch3 ch3 III I II I —ch2—ch—
ch2—ch—ch—ch2—ch2—ch— —ch2—ch—ch2—ch—ch—ch2—ch2—ch—

1-3 Insertion CH3 ch3 —ch2—ch—ch2—ch2—ch2—ch2—ch—ch2— Ga ch3


—CH2—CH—CH2—CH2—CH2—CH2—CH—CH2— i ch3 Gb Chain end
structure CH3 ch3 i i —CH— CH2— CH=CH2 Ea(m-PP') CH3 ch3 I I —ch—
ch2—ch—ch2—ch2—ch3 ch3 I —CH—CH2— CH=CH2 I ch3 Eb(r-PP') ch3 I
—ch—ch2—ch—ch2—ch2—ch3 I ch3 Fa(m-<503) Fb(r-<5/?3)

18 Polymorphism in polypropylene homo- and copolymers peaks associated


with highly tactic PP are usually accompanied by weaker ones, significant of
regio-irregular regions, copolymer sequences, chain ends and different stereo-
sequences. The various atactic structures have shifts that are displaced 0-2 ppm
from the isotactic structures [44]. Although the intensities of these resonances
are low, they contain much information on polymer structure and mode of
polymerization [45]. Extensive studies were applied to characterize the polymer
chain ends and thus to obtain informa¬ tion on initiation, chain transfer and
termination mechanisms [59]. The very low content of chain ends is a challenge
to the sensitivity of the NMR apparatus. This is the main reason why this chain
end analysis was mainly applied to low molecular weight PP, including
end analysis was mainly applied to low molecular weight PP, including
ethylene-propylene copolymers [59]. The peak assignment is very complex with
many overlap¬ ping resonances [59]. Saturated and unsaturated chain ends are
found depending on the reaction conditions, the nature of the catalytic
complexes and the chain termination mechanisms [43, 45, 59-61]. Two different
mechanisms of isotactic polymerization can occur during homogeneous catalytic
polymerization of propylene with catalyst systems based on group IV
metallocenes [45]. The steric control in the presence of chiral zirco- nocenes or
titanocenes arises from the chiral ligand environment of the atom in the catalytic
complex [45]. However the monomer insertion is less regio-specific. Regio-
irregular insertion may be followed either by a primary (1-2) insertion, which
leads to vicinal methyls, or by isomerization into 1-3 insertion of the monomer
[45]. This 1-3 isomerization is favored by a high polymerization temperature
[43]. The equal amount of vinylidene and n-propyl end groups shows that the
main chain transfer process involves /J-hydrogen abstraction from a regio-
regular monomer unit followed by initiation of a new polymer chain on a metal-
hydrogen bond via primary insertion [43]. 13C NMR is able to characterize
random copolymers as described previously for EPR and for copolymers with
low ethylene content (EPM). The main trouble is associated with the assignment
due to overlapping of resonance peaks [52]. Ethylene-propylene block
copolymers are prepared by Ziegler-Natta catalysts to form EP-, (EP)n- and
EPE-type copolymers [62, 63]. To prepare block copolymers, the catalyst must
have a sufficiently long lifetime to permit sequencing of the monomers. Boor
reported the lifetime for Ziegler-Natta catalysts [63]. An NMR apparatus should
be sensitive to the very low content of transition zone between the sequences and
thus able to distinguish a block copolymer from a mixture of homopolymers.
Block copolymers are characterized by the Sad+ carbon signal representing
propylene in a heterojunction [64]. This result is of great importance from the
point of view of block copolymerization, but very few analyses seem to be
devoted to such copolymers. 13C NMR investigation of PP solutions is able to
define the fine structure of the chain. Other analytical methods measure the
effect of monomer

Chemical composition 19 content and tacticity on conformation and


crystallization. These parameters are the main ones, but not the only ones, that
act on crystallinity. Kinetic aspects [65], i.e. crystallization conditions
(temperature, cooling rate, press¬ ure, shearing and elongation rates) also affect
the polymer crystallinity. For example, crystallinity increases linearly with
temperature in the temperature range 120°C^T^140°C [66]. Solvent extraction is
not a reliable method since the degree of extraction depends on both molecular
not a reliable method since the degree of extraction depends on both molecular
weight and stereoregularity [67]. X-ray diffraction, calorimetry and density
measure¬ ments provide values of crystallinity content. The crystallinity is the
result of a stable periodical conformation of polymer segments and of their
regular packing to form a crystalline lattice. IR and Raman spectroscopies
provide a direct analysis of polymer conformations. Absorption of IR energy
resulting in vibrational transitions requires that a change in the molecular dipole
occurs during the transitions. In Raman spectroscopy the active vibrational
transition comes from the formation of an induced dipole momentum through
interaction with the oscillation of the incident electric field vector. The
molecular symmetry also determines the selection rules for occurrence of IR- or
Raman-active vibrations. These vibrational methods give supple¬ mentary
information on polymer conformations. 1.4.2 IR analysis of EPM and
polypropylene IR spectroscopy was first used for elucidation of PP conformation
by Natta et al [1]. A quantitative analysis of structure and conformation of
polymer chains requires the assignment of absorption bands and a quantitative
measurement of their intensity. The assignment is obtained by correlation with
molecular composition and with other indices of stereoregularity. In ethylene-
propylene copolymers specific IR bands are attributed to ethylene and to
propylene segments. The composition can be separetely analysed from the
monomer sequences. Two regions of the IR spectrum are sensitive to ethylene
sequences [68]: the methylene group rocking frequen¬ cies in the 720-820 cm-1
range reflect differences in the number of contigu¬ ous methylene groups; the
methyl group rocking frequencies between 930 and 980 cm"1 reflect differences
in the number of contiguous propylene units [69]. The band at 752 cm"1 was
ascribed to two contiguous methylene groups apparently coming from head-to-
head configuration of propylene monomer units (Ba, Bb in Table 1.1). On the
contrary, insertion of an ethylene monomer in a propylene sequence induces a
three-methylene contiguous sequence which seems to give a band at 730cm"1
[69]. The band at 720 cm"1 is assigned to a sequence of five or more
methylenes, i.e. two or more ethylene monomers [69]. On this basis the ratio of
the absorbances at 720 cm"1 and 730 cm"1 is used to evaluate the copolymer
structure as a function of overall copolymer composition. This simple

20 Polymorphism in polypropylene homo- and copolymers Figure 1.5


Calibration curves for IR absorbance ratios (a) A 995/A 970, (b) A 840/A 970
versus NMR triad isotacticity from [82]: (O) annealed; (·) hot pressed. method
provided moderately good results [70]. To appreciate more precisely the
distribution of ethylene units, Bucci and Simonazzi used the three bands
characteristic of rocking of methylene groups at 752 cm-1 (two methylene
characteristic of rocking of methylene groups at 752 cm-1 (two methylene
groups), three methylene groups at 733 cm-1 and more than five methylene
groups at 722 cm"1 [71]. The results deduced from their tabulated data generally
agreed with Natta’s theoretical calculations [72]. This method applied to
methylene segments gives only indirect information on PP sequences. The peak
at 972 cm"1 is attributed to two or more contiguous head-to-tail propylene units
[73, 74], while the peak at 1155 cm"1 seems to result from isolated propylene
units [73, 74]. The ratio of intensities A 972 cm" VA 937 cm"1 was used to
study the distribution of propylene units in EPM [73]. Some disagreements with
theoretical predictions appear, which may arise from head-to-head structures as
an effect on propylene inversion [70]. For the propylene part, the band at 960
cm"1 was interpreted as a structure with two contiguous propylene units [71,
72]. The length effect on absorption frequency of ethylene and propylene
sequences is attributed to a modification of the chain conformation [75]. For
example, the 972cm"1 band in PP becomes asymmetric toward lower
frequencies when atactic content increases [75]. The distortion of the chain
caused by syndiotactic placement in the isotactic chain could reduce the helical
content [75]. For PP homopolymer, Natta et al postulated an isotacticity index
equivalent to the percentage of the insoluble part in boiling heptane, which is
still used for evaluation of PP stereoregularity [76]. From the first attempt for a
quantitative determination of isotactic content [77] various absorption bands at
809, 841, 900, 998, 1168 and 1220 cm"1 have been associated with isotactic
helices [67]. The absorbance at 995 cm"1 was attributed to a crystalline (helix)
band which disappears on heating the sample above the melting temperature
(180 °C) [77]. This band and those at 808 and 840 cm"1 were linearly related to
polymer density as a measure of

Chemical composition 21 its crystallinity [78]. More recently, extensive studies


have been carried out by Kissin and colleagues [79-81]. They observe that
specific absorption bands are related to the critical length of isotactic sequences:
975 cm -1 for at least 5 units, 998 cm-1 (11-12 units) and 841 cm-1 (13-15
units). The bands at 841 and 998 cm-1 have been principally used as standards
for the calibration curves [82, 83]. These calibration curves suppose a direct
relationship between absorption intensity and isotactic content. From a kinetic
point of view this assumption is unrealistic because highly tactic parts of the
chain participate in the amorphous phase. A review has pointed out that the IR
isotacticity index would be dependent on the thermal history [84]. To
standardize the measurement it was proposed to anneal samples in the
temperature range 160 °C^ 165 °C for at least 15 minutes [82, 85, 86]. Figure
1.5 gives an example of variation of absorption intensity of bands at 840 and 995
cm-1 with the isotacticity index deduced from NMR and with annealing effect
cm-1 with the isotacticity index deduced from NMR and with annealing effect
[82]. This attribution has been used to analyse the kinetics of polymer
crystallization by the change in absorption of ‘crystalline’ IR bands [87, 88]. The
analysis of orientation of crystalline and amorphous phases was also applied
during and after plastic deformation experiments [89]. 1.4.3 Raman spectroscopy
of polypropylene and EPM Raman and IR spectroscopies give similar
information on polymer structure and on its conformation. In contrast with IR
analyses, Raman data are mainly devoted to PP homopolymer (isotactic [90-102]
and syndio- tactic [90, 103, 104]) and rarely to copolymers [105, 106]. As an
effect of selectivity of the method, tacticity analysis is the exception [90]. Most
of the analyses concern the chain conformation in the crystalline and the molten
phases [90-107]. This conformational analysis was largely favoured by the early
development of theoretical calculation of the spectra from expected
conformations [108, 109] and of dichroism experiments [90,111]. The normal
vibrations of the helical conformation (30 of isotactic PP, in the crystalline
phase, has a symmetry of class C3 [92, 95]. As a consequence the vibrations fall
into two classes, A and E, both of which are Raman and infra-red active and
which have infra-red dichroic characteristics and Raman depolarizations: A-
Raman (pol.), infra-red (parallel) and E-Raman (depol.), infra-red
(perpendicular) [92, 95]. The 104 non-zero-spectroscopi- cally active phonons
are distributed as follows: 25 of type A and 26 of type E [96]. The vibrational
spectra of iso tactic [108] and syndio tactic [109] forms have been calculated by
a coordinate analysis assuming that the vibrational characteristics are those of
the isolated single chain. The Raman bands observed by various authors are in
very good agreement [90, 93, 96, 97].

22 Polymorphism in polypropylene homo- and copolymers The observed


frequencies correspond to those calculated from chain conformation [90]. No
significant difference resulting from the non-isotactic parts of the chain were
revealed [90, 93, 96, 97]. The main Raman bands can be separated into three
groups which depend on symmetry Α/Λ A± and E [96]: Ah bands at 808, 973
and 1167cm-1; A± bands at 841, 997 and 1151cm-1; and E bands at 1219, 1330
and 1458 cm-1. Ah bands are characteristic of extended chains in the crystalline
phase, while A_l and E bands depend both on amorphous and crystalline phases
[93, 96]. The band at 808 cm-1 is attributed to the vibration of the valence bond
along the chain axis vsq(CC), the bands at 973 cm-1 and 1167cm-1 should result
from a rocking of CH3 group p//(CH3) and angular deforma¬ tion of CH group
(5//(CH), respectively [93]. Other bands are interpreted in the same way [93].
Raman spectra of the α-, yS-, y- and smectic polymorphs of isotactic PP were
recorded at various temperatures [93]. The spectra of the polymorphs were very
recorded at various temperatures [93]. The spectra of the polymorphs were very
similar except in the antisymmetric C-H stretch¬ ing region. The sensitivity of
Raman bands to chain direction has been applied to analyse polymer orientation
during stretching [95] and after processing [94, 96, 99, 110, 111]. This method is
pertinent for analysis of polymer orientation and to obtain real-time information
on the change in the chain conformation by longitudinal acoustic mode (LAM)
analysis [101, 102]. Raman spectroscopy is rarely applied to copolymers [105,
106] and to heterotacticity analyses [90]. Up to now, the higher sensitivity of IR
spectroscopy is more suitable for the structural analysis of PP chains. Raman
spectroscopy is an invaluable method for conformation analysis of isotactic [91,
104, 105] and syndiotactic [90] PP chains in the melt. All of these papers
confirm that a large amount of helical sections of the polymer chain still persists
in the polymer melt. It is not possible to define the length of these oriented
segments [90]. This unexpected but widely recognized result must be borne in
mind for the understanding of the mechanisms of crystallization. 1.4.4
Crystalline phases and their stability deduced from X-ray diffraction and
calorimetric analyses The presence of crystalline phases in PP was evidenced
first by X-ray diffraction on isotactic and syndiotactic PP [7-10]. The three
phases of isotactic PP correspond to various packings of 3t helices. In
monoclinic a phase each chain is in contact with three adjacent enantiomorphous
and two adjacent isomorphous chains [7]. The hexagonal β phase corresponds to
groups of three isomorphous chains [9]. These various packings of chain
segments induce a change in their interaction. Consequently, a higher
thermodynamic stability of the phases results from an increase of chain
interactions. The equilibrium melting temperature and the enthalpy of

Chemical composition 23 fusion mainly characterize the thermodynamic


properties of each phase. The entropy change during equilibrium transition is
then deduced. These thermodynamic properties of each crystalline phase can be
deduced from calorimetric measurements. These values are extrapolated from
experiments performed on semi-crystalline PP. The equilibrium melting
temperature is depressed to the experimental melting temperature Tm by the
very small thickness T of crystalline lamellae. The equilibrium melting
temperature is extrapolated to hypothetical large dimension crystals with the
same crystal¬ line structure. The equilibrium melting temperature is
extrapolated, either from plots of Tm=f(l/T), where T is deduced from small
angle X-ray scattering (SAXS) experiments, or from plots of Tm =/(Tc), where
Tc is the crystallization temperature, which controls the lamellae thickness. Long
range extrapolates causes discrepancies in the equilibrium melting tempera¬
ture: from 180 °C to 220 °C for a phase [66, 112-114], and 176 °C for β phase of
ture: from 180 °C to 220 °C for a phase [66, 112-114], and 176 °C for β phase of
isotactic PP [114]. The most probable value for a phase is Tm° = 210°C±2°C
[66, 113]. Block copolymerization does not seem to modify the crystalline
organization of polymer chains either in the crystal¬ line or in the amorphous
phase. On the contrary, a random copolymeriz¬ ation with ethylene depresses
the equilibrium melting temperature of the PP a phase [6]. This effect is
explained by the rejection of ethylene segments into the amorphous phase,
outside the PP crystalline phase [6]. The structural defects caused by random
segments are largely incompatible with a regular crystalline arrangement.
Enthalpy of fusion can be deduced from the heat of fusion of semi-crystalline PP
as measured by calorimetry. The enthalpy of fusion ΔHm of PP a phase is also
widely dispersed as an effect of the extrapolation method: 63 J/g<Aifm ^260 J/g
[115]. The most probable value seems to be ΔHm = 148 J/g [66]. These
thermodynamic characteristics are a consequence of the structure, the
conformation and the packing of chain segments. 1.4.5 Dynamic changes of
conformation from dynamic mechanical analysis and solid-state NMR In its
solid state the semi-crystalline PP is a material out of thermo¬ dynamic
equilibrium since two phases coexist in the solid state (crystal¬ line and
amorphous) and only one in the melt. The organization of the polymer can
change under a thermal (annealing) or a mechanical (plastic deformation)
treatment. It is necessary to apply only weak mechanical perturbations to avoid a
change in the organization of the material during its analysis. Various methods
have been applied to analyse the chain organization and the resulting dynamic
change under small thermal or mechanical perturbations: TMA, dynamic
mechanical analysis (DMA), thermo-stimulated currents (TSC), thermo-
stimulated relaxation, solid- state NMR. Of all these methods, DMA and solid-
state NMR are most

24 Polymorphism in polypropylene homo- and copolymers widespread and


useful in PP analysis. A DMA spectrum displays three marked mechanical
relaxations (α, /?, y) under periodic oscillation during heating [116]. Solid-state
NMR provides the relaxation decrement time of a specific chemical group and a
more precise analysis of the con¬ formation change in the crystalline and the
amorphous phases. The β relaxation is the glass-rubber relaxation of the
amorphous portions of the solid (T=0°C at 1 Hz) [116, 117]. The y relaxation is
also an amorphous relaxation (Tss —100 °C at 1 Hz) [116] involving a few
chain segments [117]. The a relaxation seems to be a crystal relaxation that
depends on crystal morphology (100°C<T^130°C as a function of thermal
treatment) [118]. The assignment of segments involved in these relaxations has
not yet been unambigously determined. Solid-state NMR is able to distinguish
not yet been unambigously determined. Solid-state NMR is able to distinguish
the crystalline and amorphous phases, to reveal chain dynamics during thermal
treatment of PP with various tacticities. The structural change during
deformation and its time constant can be analysed. A specific attribution of
solid-state NMR peaks was first applied to isotactic, syndiotactic and atactic PP
[119, 120]. Solid-state NMR can be sensitive not only to chain conformational
arrangements but also to different packing environments. Cross-polarization
with rapid magic angle sample spinning (CPMAS) allows study of the packing
of α, β and smectic crystalline polymorphs of isotactic PP, thus checking the X-
ray diffrac¬ tion results [121, 122]. The main X-ray results were confirmed by
this way except for the smectic form of isotactic PP [121]. The 13C chemical
shifts and the spin-lattice relaxation times of smectic form carbons indicate that
their packing in smectic form is very similar to that in β form (hexagonal)
crystals [121]. Slow chain dynamics leading to ex¬ change between different
conformations can be detected by high resolu¬ tion ID or 2D NMR. These
conformational transitions have been studied near the glass transition of atactic
[120], isotactic [123] and syndio¬ tactic [124] PP. In the amorphous atactic PP a
residual conforma¬ tional order is present [120]. The conformational exchange
near the glass transition is due to rotational motions but not with well-defined
rotational angles [120]. This diffusive motion of atactic chains has a temperature
variation in the mean correlation time which seems to follow the Williams-
Landel-Ferry (WLF) equation over 11 orders of magnitude between 10“10 and
101 s [123]. On the contrary, the change of chain conformation proceeds by 120°
rotation of the 3X helix about the helix axis of isotactic PP, which produces
discrete jump motion in the crystals [123]. The symmetry of the helix allows a
rotational jump motion of the polymer chain around the helix axis in connection
with a transla¬ tion so that a given monomer unit, after the jump, occupies the
position of the neighbouring monomer unit before the jump [125]. This analysis
is now applied to characterize stretched films and during stretching of the PP
films [126, 127].

Appendix 25 1.5 CONCLUSIONS Many studies have been devoted to the


analysis of the structure, conforma¬ tion and dynamics of PP and copolymer
chains. These physico-chemical characterizations are necessary foundations to
link molecular structure, crystalline organization and material properties for
polymers. Nevertheless, a complete analysis of the chemical composition
(copolymer distribution, chemical defects, tacticity, chain end) of a molecule in a
polymeric sample as a function of its molecular weight has not yet been
published. Usually mean structures and characteristics (copolymer composition,
tacticity) are con¬ sidered for polymolecular samples. Methods resulting in
mean parameters deliver information about how the chemical structure affects
mean parameters deliver information about how the chemical structure affects
the conforma¬ tion, the development of crystalline phases and morphologies, the
phase transitions and dynamic properties. Meanwhile, these data are sufficient
for the understanding of the main features of a polymer chain with regard to its
organization, processability and properties. 1.6 APPENDIX: SYMBOLS AND
ABBREVIATIONS A2 Second virial coefficient c Polymer concentration DPn
Number-average degree of polymerization DSC Differential scanning
calorimetry DMA Dynamic mechanical analysis EPM Ethylene-propylene
copolymer EPDM Ethylene-propylene diene monomer EPR Ethylene-propylene
rubber GPC Gel permeation chromatography /red(0, c) Intensity scattered at
angle Θ by a solution with a concentration c IR Infra-red k experimental
constant in light scattering experiments K Constant in the Mark-Houwink
relationship Κθ Constant in the Mark-Houwink relationship in Θ condition T
Mean thickness of crystalline lamellae MeΛ Molecular weight of ethylene
monomer Mpr Molecular weight of propylene monomer Mn Number-average of
molecular weight of the polymer Mw Weight-average of molecular weight of the
polymer Mz z-average of molecular weight of the polymer n Degree of
polymerization Nj Number of molecules with a weight M-x NMR Nuclear
magnetic resonance Ρ{θ) Coefficient in light scattering experiment for the
observation angle Θ

26 Polymorphism in polypropylene homo- and copolymers R Gas constant


SAXS Small angle X-ray scattering T Absolute temperature Tm Experimental
melting temperature Tm° Equilibrium melting temperature of a crystalline phase
TMA Thermo-mechanical analysis TMS Tetramethyl silane TSC Thermo-
stimulated currents v2 Volume fraction of polymer in the solution Vs molar
volume of the solvent a Exponent in the Mark-Houwink relationship β Molar
fraction of propylene in a copolymer β0 Binary cluster integral for a pair of
chain segments δ Partage coefficient of polymer molecules between two media
during a fractionation ΔHm Enthalpy of fusion of a crystalline phase AGi Partial
molar free energy of mixing in a solution M Intrinsic viscosity of polymer (also
limiting viscosity) [η]θ Intrinsic viscosity of polymer in Θ condition Π Osmotic
pressure p Polymer density φ0 Universal viscosity constant Xi Huggins
interaction coefficient between solvent and polymer REFERENCES 1. Natta, G.,
Pino, P., Corradini, P. et al. (1955) Journal of the American Chemical Society,
77, 1708. 2. Ziegler, K. (1952) Angewandte Chemie, 64, 323. 3. Galanti, A. V.
and Mantell, C. L. (1965) Polypropylene Fibers and Films, Plenum Press, New
York. 4. Mahajan, S. J., Bhaumik, K. and Deopura, B. L. (1991) Journal of
Applied Polymer Science, 43, 49. 5. Bonnar, R. V., Dimbat, M. and Stross, F. H.
(1964) Number Average Molecular Weights, Wiley-Interscience, New York. 6.
(1964) Number Average Molecular Weights, Wiley-Interscience, New York. 6.
Monasse, B. and Haudin, J. M. (1988) Colloid and Polymer Science, 266, 679. 7.
Natta, G., Corradini, P. and Cesari, M. (1956) Atti dell ’Accademia Nazionale
dei Lincei. Rendiconti, 21, 365. 8. Natta, G., Corradini, P. and Ganis, P. (1962)
Journal of Polymer Science, 58, 1191. 9. Turner-Jones, A., Aizlewood, J. M. and
Beckett, D. R. (1964) Makromolekulare Chemie, 75, 134. 10. Lotz, B.,
Lovinger, A. J. and Cais, R. S. (1988) Macromolecules, 21, 2375. 11. Ibhadon,
A. O. (1991) Journal Applied Polymer Science, 42, 1887. 12. Ying, Q., Xie, P.
and Ye, M. (1985) Makromolekulare Chemie, Rapid Com¬ munications, 6, 105.

References 27 13. Kinsinger,' J. B. and Hughes, R. E. (1959) Journal of Physical


Chemistry, 63, 2002. 14. Vaughan, M. F. and Francis, M. A. (1977) Journal of
Applied Polymer Science, 21, 2409. 15. Crouzet, P., Fine, F. and Mangin, P.
(1969) Journal of Applied Polymer Science, 13, 205. 16. Ouano, A. C. and
Mercier, P. L. (1968) Journal of Polymer Science, Part C, 21, 309. 17. Atkinson,
C. M. L. and Dietz, R. (1976) Makromolekulare Chemie, 177, 213. 18. Coll, H.
and Gilding, D. K. (1970) Journal of Polymer Science, Part A2, 8, 89. 19.
Grinshpun, V. and Rudin, A. (1985) Journal of Applied Polymer Science, 30,
2413. 20. Qicong, Y., Ping, X., Yong, L. and Renyuan, Q. (1986) Journal of
Liquid Chromatography, 9(6), 1233. 21. Huggins, M. L. (1942) Journal of
Physical Chemistry, 46, 151. 22. Flory, P. J. (1942) Journal of Chemical Physics,
10, 51. 23. Huggins, M. L. (1964) Journal of the American Chemical Society,
86(17), 3535. 24. Ogawa, T., Tanaka, S. and Oshino, S. (1972) Journal of
Applied Polymer Science, 16, 2257. 25. Ogawa, T., Suzuki, Y. and Inaba, T.
(1972) Journal of Polymer Science, Part Al, 10, 737. 26. Chiang, R. (1958)
Journal of Polymer Science, 28(116), 135. 27. Grubisic, Z., Rempp, P. and
Benoit, H. (1967) Journal of Polymer Science, Part B, 5(9), 753. 28. McIntyre,
D., Fetters, L. J. and Slagowski, E. (1972) Science, 176, 1041. 29. Bianchi, U.
and Peterlin, A. (1968) Journal of Polymer Science, Part A2, 6, 1759. 30.
Ogawa, T. and Inaba, T. (1977) Journal of Applied Polymer Science, 21, 2979.
31. Ivan, B., Lazlo-Hedvig, Z., Kelen, T. and Tudos, F. (1982) Polymer Bulletin,
8, 311. 32. Wang, K. Q., Zhang, S. Y., Xu, J. and Li, Y. (1982) Journal of Liquid
Chromatography, 5(10), 1899. 33. Scholte, T. G., Meijerink, N. L. J.,
Schoffeleers, Η. M. and Brands, A. M. (1984) Journal of Applied Polymer
Science, 29, 3763. 34. Dawkins, J. V. (1968) Journal of Macromolecular
Science-Physics, B2(4), 623. 35. Yamakawa, H. (1971) Modern Theory of
Polymer Solutions, Harper and Row, New York, p. 364. 36. Stockmayer, W. H.
and Fixman, M. (1963) Journal of Polymer Science, Part C, Polymer Symposia,
1, 137. 37. Tchir, W. J., Rudin, A. and Fyfe, C. A. (1982) Journal of Polymer
Science, Polymer Physics Edition, 20, 1443. 38. Sadiku, E. R. and Peacock, N.
Science, Polymer Physics Edition, 20, 1443. 38. Sadiku, E. R. and Peacock, N.
(1990) Angewandte Makromolekulare Chemie, 181, 67. 39. De Chirico, A.,
Arrighetti S. and Bruzzone, M. (1981) Polymer, 22, 529. 40. Smith, W. V.
(1974) Journal of Applied Polymer Science, 18, 3685. 41. Haidar, B., Vidal, A.,
Balard, H. and Donnet, J. B. (1984) Journal of Applied Polymer Science, 29,
4309. 42. Coupek, J., Pokorny, S., Protivova, J. et al (1972) Journal of
Chromatography, 65, 279. 43. Grassi, A., Zambelli, A., Resconi, L. et al. (1988)
Macromolecules, 21, 617. 44. Carman, C. J., Harrington, R. A. and Wilkes, C. E.
(1977) Macromolecules, 10, 536.

28 Polymorphism in polypropylene homo- and copolymers 45. Cheng, Η. N. and


Ewen, J. A. (1989) Makromolekulare Chemie, 190, 1931. 46. Woodbrey, J. C.
(1964) Journal of Polymer Science Part B, Polymer Letters, 2(3), 315. 47.
Bovey, F. A. (1972) High Resolution NMR of Macromolecules, Academic
Press, New York, Chapter VII. 48. Heatley, F. and Zambelli, A. (1969)
Macromolecules, 2(6), 618. 49. Ferguson, R. C. (1971) Macromolecules, 4(3),
324. 50. Randall, J. C. (1978) Macromolecules, 11(1), 33. 51. Smith, W. V.
(1980) Journal of Polymer Science, Polymer Physics Edition, 18(7), 1573. 52.
Cheng, Η. N. (1984) Macromolecules, 17, 1950. 53. Cheng, Η. N. and Lee, G.
H. (1987) Macromolecules, 20, 436. 54. Miyatake, T., Kawai, Y., Seki, Y. et al.
(1989) Polymer Journal 21(10), 809. 55. Asakura, T., Nishiyama Y. and Doi, Y.
(1987) Macromolecules, 20, 616. 56. Schilling, F. C. and Tonelli, A. E. (1980)
Macromolecules, 13, 270. 57. Zambelli, A., Locatelli, P., Provasoli, A. and
Ferro, D. R. (1980) Macro¬ molecules, 13(2), 267. 58. Cheng, Η. N. and Lee, G.
H. (1985) Polymer Bulletin, 13, 549. 59. Cheng, Η. N. and Smith, D. A. (1986)
Macromolecules, 19, 2065. 60. Zambelli, A., Ammendola, P., Grassi, A. et al
(1986) Macromolecules, 19, 2703. 61. Busfield, W. K. and Hanna, J. V. (1991)
Polymer Journal, 23(10), 1253. 62. Allport, D. C. and Janes, W. H. (eds) (1973)
Block Copolymers, Wiley, New York. 63. Boor Jr, J. (1979) Ziegler-Natta
Catalysts and Polymerization, Academic Press, New York. 64. Prabhu, P.,
Schindler, A., Teil, Μ. H. and Gilbert, R. D. (1980) Journal of Polymer Science,
Polymer Physics Edition, 18, 389. 65. Hoffman, J. D., Davis, G. T. and
Lauritzen Jr, J. I. (1976) The rate of crystallization of linear polymers with chain
folding, in Treatise on Solid State Chemistry, Vol. 3 (ed N. B. Hannay), Plenum
Press, New York. 66. Monasse, B. and Haudin, J. M. (1985) Colloid and
Polymer Science, 263, 822. 67. Burfield, D. R. and Loi, P. S. T. (1988) Journal
of Applied Polymer Science, 36, 279. 68. Drushel, Η. V., Ellerbe, J. S., Cox, R.
C. and Lane, L. H. (1968) Analytical Chemistry, 40, 370. 69. Veerkamp, T. A.
and Veermans, A. (1961) Makromolekulare Chemie, 50, 147. 70. Drushel, Η. V.
(1970) CRC Critical Reviews in Analytical Chemistry, 161. 71. Bucci, G. and
(1970) CRC Critical Reviews in Analytical Chemistry, 161. 71. Bucci, G. and
Simonazzi, T. (1964) Journal of Polymer Science, Part C, 7, 203. 72. Natta, G.,
Mazzanti, G., Valvassori, A. et al. (1960) Chimica e Industria (Milan), 42(2),
125. 73. Ciampelli, F. and Valvassori, A. (1967) Journal of Polymer Science,
Part C, 16(1), 377. 74. Natta, G., Valvassori, A., Ciampelli, F. and Mazzanti, G.
(1965) Journal of Polymer Science, Part A, 3, 1. 75. Koenig, J. L., Wolfram, L.
E. and Grasselli, J. G. (1966) Spectrochimica Acta, 22, 1233. 76. Natta, G.,
Mazzanti, G., Crespi, G. and Moraglio, G. (1957) Chimica e Industria, (Milan),
39, 275. 77. Luongo, J. P. (1960) Journal of Applied Polymer Science, 3(9), 302.
78. Quynn, R. G., Riley, J. L., Young, D. A. and Noether, H. D. (1959) Journal
of Applied Polymer Science, 2, 166.

References 29 79. Kissin, Yu. V., Tsvetkova, V. I. and Chirkov, N. M. (1972)


European Polymer Journal, 8, 529. 80. Kissin, Yu. V., Lekae, I. A., Chermova,
E. A. et al. (1974) Vysokomolkulyarnie Soedineniya, Series A, 16(3), 677. 81.
Kissin, Yu. V. and Rishina, L. A. (1976) European Polymer Journal, 12, 757. 82.
Burfield, D. R. and Loi, P. S. T. (1986) Approaches of the problem of
isotacticity determination in polypropylene, in Catalytic Polymerization of
Olefins (eds T. Keii and K. Soga), Kodansha, Tokyo, p. 387. 83. Mirabella Jr, F.
M. (1990) Journal of the American Chemical Society, 21, 357-750. 84. Krimm,
S. (1960) Advances in Polymer Science, 2, 51. 85. Brader, J. (1960) Journal of
Applied Polymer Science, 3(9), 370. 86. Tadokoro, H., Kobayashi, M., Ukita,
M. et al. (1965) Journal of Chemical Physics, 42(4), 1432. 87. Mirabella Jr, F.
M. (1987) Journal or Polymer Science, Part B, Polymer Physics, 25, 591. 88.
Eder, M. and Wlochowicz, A. (1984) Acta Polymerica, 35(8), 548. 89.
Chalmers, J. M., Mackenzie, M. W., Willis, H. A. et al. (1991) Spectrochimica
Acta, 47A(12), 1677. 90. Masetti, G., Cabassi, F. and Zerbi, G. (1980) Polymer,
21, 143. 91. Hallmark, V. M., Bohan, S. P., Strauss, H. L. and Snyder, R. G.
(1991) Macromolecules, 24, 4025. 92. Fraser, G. V., Hendra, P. J., Watson, D. S.
et al. (1973) Spectrochimica Acta, 29 A, 1525. 93. Chalmers, J. M., Edwards, H.
G., Lees, J. S. et al. (1991) Journal of Raman Spectroscopy, 22, 613. 94. Bailey,
R. T., Hyde, A. J. and Kim, J. J. (1974) Spectrochimica Acta, 30A, 91. 95.
Vasco, P. and Koenig, J. L. (1970) Macromolecules, 3(5), 597. 96. Cambon, L.
D. and Luu, D. V. (1983) Journal of Raman Spectroscopy, 14(4), 291. 97.
Cambon, L. D., Ramonja, J. L. and Luu, D. V. (1987) Journal of Raman
Spectroscopy, 18, 129. 98. Ize-Iyamu, Μ. I. (1983) Spectroscopy Letters, 16(1),
29. 99. Hendra, P. J. and Willis, H. A. (1967) Chemistry and Industry, 23, 2146.
100. Schaufele, R. F. (1967) Journal of the Optical Society of America, 57, 105.
101. Rabolt, J. F. and Franconi, B. (1977) Journal of Polymer Science, Polymer
Letters Edition, 15, 121. 102. Hsu, S. L., Krimm, S., Krause, S. and Yeh, G. S.
Y. (1976) Journal of Polymer Science, Polymer Letters Edition, 14, 195. 103.
Y. (1976) Journal of Polymer Science, Polymer Letters Edition, 14, 195. 103.
Chalmers, J. M. (1977) Polymer, 18, 681. 104. Zerbi, G. and Hendra, P. (1969)
Journal of Molecular Spectroscopy, 30, 159. 105. Fraser, G. V., Hendra, P. J.,
Walker, J. H. et al. (1973) Makromolekulare Chemie, 173, 205. 106. Schreier, G.
and Peitscher, G. (1972) Zeitschrift fur analytische Chemie, 258, 199. 107.
Killough, P. M., De Vito, V. L. and Asher, S. A. (1991) Applied Spectroscopy,
45(7), 1067. 108. Snyder, R. G. and Schachtschneider, J. H. (1964)
Spectrochimica Acta, 20, 853. 109. Schachtschneider, J. H. and Snyder, R. G.
(1965) Spectrochimica Acta, 21, 1527. 110. Satija, S. K. and Wang, C. H. (1978)
Journal of Chemical Physics, 69(6), 2739. 111. Wang, C. H. and Cavanaugh, D.
B. (1981) Journal of Applied Physics, 52(10), 6003. 112. Martuscelli, E.,
Pracella, M. and Zambelli, A. (1980) Journal of Polymer Science, Polymer
Physics Edition, 18, 619.

30 Polymorphism in polypropylene homo- and copolymers 113. Mucha, M.


(1981) Journal of Polymer Science, Polymer Symposia, 69, 79. 114. Samuels, R.
J. (1975) Journal of Polymer Science, Polymer Physics Edition, 13(7), 1417.
115. Wunderlich, B. (1980) Macromolecular Physics, Vol. 3, Academic Press,
New York. 116. Flocke, H. A. (1960) Kolloid Zeitschrift und Zeitschrift fur
Polymere, 180(2), 118. 117. Passaglia, E. and Martin, G. M. (1964) Journal of
Research of the National Bureau of Standards, 68A(5), 519. 118. McCrum, N.
G. (1964) Journal of Polymer Science, Polymer Letters, 2, 495. 119. Bunn, A.,
Cudby, Μ. E. A., Harris, R. K. et al (1981) Journal of the Chemical Society,
Chemical Communications, 1, 15. 120. Zemke, K., Chmelka, B. F., Schmidt-
Rohr, K. and Spiess, H. W. (1991) Macromolecules, 24, 6874. 121. Tonelli, A.
L., Gomez, M. A., Tanaka, H. et al. (1990) Journal of the American Chemical
Society, 24, 409. 122. Briickner, S., Meille, S. V., Sozzani, P. and Torri, G.
(1990) Makromolekulare Chemie, Rapid Communications, 11, 55. 123.
Schaefer, D., Spiess, H. W., Suter, U. W. and Fleming, W. W. (1990) Macro¬
molecules, 23, 3431. 124. Tanaka, H. (1991) Makromolekulare Chemie, Rapid
Communications, 12, 307. 125. Hagenmeyer, A., Schaefer, D., Schmidt-Rohr,
K. et al. (1990) Molecular Crystals, Liquid Crystals, 187, 223. 126. Tanaka, H.
and Takagi, K. (1989) British Polymer Journal, 21, 519. 127. Tanaka, H. and
Harada, H. (1991) European Polymer Journal, 27(1), 1.

2 Crystalline structures of polypropylene homo- and copolymers S. Z. D. Cheng,


J. J. Janimak and J. Rodriguez 2.1 INTRODUCTION Since 1958, when isotactic
polypropylene (i-PP) became commercially available, the crystalline architecture
and crystallization of i-PP from the melt and solution have been topics of
interest. Industrial applications have also been developed for copolymers
interest. Industrial applications have also been developed for copolymers
associated with polypropylene (PP). Recently, a renewed interest has also been
directed towards syndio- tactic polypropylene (s-PP). PP is a member of the first
class of vinyl polymers. It is naturally expected that crystal structure and
crystalli¬ zation behavior are not only dependent upon molecular weight and
molecular weight distribution, but also upon tacticities. Very often, it is difficult
to separate the tacticity effect from other factors in the study of PP crystal
structure, unit cell geometry and crystallization behaviour. Each i-PP chain
molecule possesses a chain conformation with a 2 x 3/1 helix (tgtg). Since the
presence of the asymmetrically substituted methyl groups causes rotation around
the backbone bonds to be direction-dependent, both right-handed and left-
handed helices with stereo-isomer configurations of d and / result [1-5].
Combination of these possibilities leads to four distinguishable chain
conformations (see Chapter 1). The intramolecular interaction energies of all
four types of helices are indentical. Their intermolecular interactions with each
other Polypropylene: Structure, blends and composites. Edited by J. Karger-
Kocsis. Published in 1995 by Chapman & Hall, London. ISBN 0 412 58430 1

32 Crystalline structures of polypropylene homo- and copolymers in the crystal


packing, however, depend upon packing geometry. Best packing is frequently
achieved when the nearest neighbors of right- handed helices are the
enantiomorphic left-handed helices and vice versa. Different packing geometries
lead to four well-known crystal structures (polymorphs): the monoclinic (a) form
[6], the hexagonal (β) form [6-9], the triclinic (y) form [6, 7, 10-12] and the
quenched form [6]. Some reports have also proposed the existence of a δ form
[13]. Among these crystal structures, the monoclinic (a) form is by far the most
common, being formed in normal melt-crystallized and solution-crystallized i-
PP samples. s-PP possesses three different chain conformations. One of them is a
4x2/1 helix (ttgg)2, and its close packing leads to a high temperature
orthorhombic lattice in the crystals with two different unit cell sizes [14-20]. A
planar zigzag chain conformation (tt) can also be observed [3,21], and a low
temperature, quenched orthorhombic lattice is found [22]. Finally, a packing
with (t6g2t2g2) chain conformation forms a triclinic lattice in the crystals, which
serves as an intermediate state between the two orthorhom¬ bic forms [23].
Detailed descriptions of the chain conformations can be found in Chapter 1.
Copolymers associated with PP have a very broad composition spectrum. At
least two different categories can be identified based on structure parameters.
The first is the copolymers with different chemical composi¬ tions associated
with PP comonomers. One example is a series of ethylene-propylene copolymers
in which polyethylene (PE) is the host and the propylene comonomer unit is the
in which polyethylene (PE) is the host and the propylene comonomer unit is the
guest [24-29]. More generally, one may classify other polyolefins as the hosts
and propylene as the guest. This category, ethylene-propylene copolymers in
particular, has attracted the most attention in both academic and industrial
research because of possible applications. Although the propylene comonomer
units may be included into the PE crystals as defects, the mole per¬ centage of
the comonomer units critically affects the crystal structure, unit cell geometry
and crystallization of the copolymers. Another example is PP hosted copolymers
with ethylene and other short branch co¬ monomer units [30-33]. These
copolymers are studied much less than the previous category. Nevertheless,
research activities regarding these copolymers are gaining strength. The second
category was less recognized for a number of years compared with the previous
category since these copolymers are basically the result of a stereo-effect,
namely, i-PP with certain isotacticity can be viewed as a stereo-copolymer with
comonomer units having methyl groups positioned in different directions [34-
37]. When one studies the crystal structure and unit cell geometry of the
copolymers, another important issue is the stereo-irregular sequence dis¬
tribution of the comonomer units in the chain molecules. In some cases, the
distribution may not be random, and stereoblocks may form

33 In isotactic polypropylene along the chain molecules. As a result, the crystal


structure and crystalliza¬ tion behaviour may be changed. In this chapter, we
focus on the crystal structures and unit cell geometries of different PPs and their
copolymers. It is particularly important to understand the relationships of the
crystallographic data to molecular parameters such as chemical and stereo-
isomer compositions, thermo¬ dynamic properties of the different PP crystals
and various crystallization conditions. 2.2 CRYSTAL STRUCTURES IN
ISOTACTIC POLYPROPYLENE The monoclinic lattice of the a form in i-PP
was identified by Natta and Corradini [6]. Shortly afterwards, a polymorph with
a hexagonal lattice was recognized, and designated as the β form [6-9]. An even
rarer third poly¬ morph was found, based on a triclinic lattice; this was called
the y form [6, 7, 10-12]. There is mention of another structure, the δ form, in i-
PP with a high percentage of amorphous material, which may be associated with
the nonisotactic portion of PP [13]. In addition to these crystal structures, a
quenched crystal form, called the smectic form by Natta and Corradini [6], was
also observed in i-PP. In all of these structures, the chain conformation of each
individual chain molecule is believed to be identical with a 2 x 3/1 helix (tgtgtg
conformation) observed from infrared (IR) spectroscopy results [1-5]. The
different polymorphs are distinct due to the different chain packing geometries
of the helices. The unoriented powder wide angle X-ray diffraction (WAXD)
patterns for different forms in i-PP crystals are shown in Figure 2.1. Formation
patterns for different forms in i-PP crystals are shown in Figure 2.1. Formation
of these different forms is critically dependent upon crystallization conditions
and molecular characteristics. 2.2.1 Monoclinic lattice of a form The crystal unit
cell of the a form was reported to be a = 0.666nm, b = 2.078 nm, c = 0.6495 nm,
β = 99.62° and a = y = 90°, with a crystallographic symmetry (space group) of
P2i/c [7]. The number of repeating units involved in each unit cell is twelve, and
the crystallographic density is 0.946 g/cm3. Based on the unit cell reported by
Natta and Corradini [6], 0 = 0.665 nm, b = 2.096 nm, c = 0.650 nm, β = 99.33 °
and a = y = 90°, the crystallographic density of the a form should be 0.936
g/cm3. It is particu¬ larly interesting that the three-fold screw axis of the i-PP
molecule is not part of the crystal symmetry, as shown in Figure 2.2. For
simplicity, only isoclined helices have been drawn. From this figure, the three
contacts with enantiomorphic helices (opposite handedness contacts) are evident,
but the a form of i-PP also has two isomorphic (equal handedness contacts)
helices. This leads to an increase in the coordination number without loss of
packing by contacting helices of equal handedness. Another important

34 Crystalline structures of polypropylene homo- and copolymers 2Θ Figure 2.1


Unoriented powder WAXD patterns for different forms in i-PP crystals. Figure
2.2 Projection of crystal unit cells of the a form in i-PP crystals on the ab plane.
The chain direction (c axis of the crystals) is perpendicular to this plane. The
methyl groups are all in the inclined positions for simplicity. feature lies in the
positioning of the methyl groups within the polymer, namely, isoclined (the
methyl group is ‘up’ related to the left glide plane) and anticlined (the methyl
group is ‘down’ related to the right glide plane) or vice versa. This introduces a
considerable degree of disorder in i-PP crystals. Mencik [38] proposed that 45%
disorder with respect to the isoclined and the anticlined helices must exist in
order to explain the experimentally observed WAXD results.

35 In isotactic polypropylene Recently, it was also reported that based on


WAXD and differential scanning calorimetry (DSC) experimental observations,
the a form in i-PP can be recrystallized and/or annealed from a less ordered a*
form with a random distribution of ‘up’ and ‘down’ chain packing of methyl
groups to a more ordered a2 form with a well-defined deposition of ‘up’ and
‘down’ helices in the crystal unit cell [39-44]. It was also recognized that the
crystallographic symmetry of the ax form is C2/c. In fact, the packing energy
difference between these two forms is only a few tenths of one kJ/mol. The
space groups of these two forms are shown in Figure 2.3. Figure 2.3 Two
monoclinic space groups, corresponding to ax and a2 forms in i-PP.
36 Crystalline structures of polypropylene homo- and copolymers The a form is
the most conventional and most extensively studied crystal structure in i-PP.
Several research groups have reported their morphological observations of
spherulitic i-PP through polarized light microscopy (PLM) in wide
crystallization temperature (Tc) range [45-48]. They classified the i-PP
spherulitic textures into four distinct types: I to IV. Types I and II spherulites are
found in different temperature ranges with different respect¬ ive birefringence
values. They all have the monoclinic lattice of the a form. On the other hand,
types III and IV less commonly appear within certain constraints of isothermal
Tc, and are characterized by a strong negative birefringence. They crystallize
with the hexagonal lattice of the β form (see below). From transmission electron
microscopy (TEM) observations, the crystal¬ line morphology of i-PP is
dominated by a highly characteristic lamellar branching (cross-hatching), which
has no counterpart in other crystalline polymers [10, 45-55]. The branching is
observed only for the a form, and not for the β and y forms. It has been
recognized that this lamellar branching is characterized by a constant angle
between daughter (tangential direction in the spherulites) and mother (radial
direction in the spherulites) lamellae (80 ° or 100 °). The degree of branching is
Tc-dependent. This leads to the birefringence change from a positive to a
negative value in spherulites with increasing Tc. Recently, it has also been found
that the degree of branching decreases with isotacticity in i-PP crystals. In
principle, the branching is an epitaxial growth phenomenon. Explanations have
been proposed in three different aspects: molecular, crystallographic and mor¬
phological [10,46,48, 51-56]. The most thorough study was made by Lotz and
Wittmann [56]. They suggested that structurally the daughter lamellae in i-PP
grow epitaxially on the lateral (010) crystallographic plane of the mother
lamellae by a satisfactory interdigitation of the methyl groups of facing planes.
This condition arises from the near identity of the size of the a and c
crystallographic axes in the unit cell of the monoclinic a form. From a molecular
point of view, chains that deposit onto the (010) plane for the initiation of this
epitaxy have the same helical handedness as chains in the (010) plane, but with a
substantial angle of 80° or 100° in order to have favorable interactions of the
methyl groups in the helices [56]. A detailed description of this crystal
morphology will be reviewed in Chapter 3. The thermodynamic properties of the
a form crystals have been actively investigated. The equilibrium melting
temperature of the crystals (Tm°) was reported as ranging between 185 °C and
209 °C based on DSC and small angle X-ray scattering (SAXS) experimental
observations. Two groups of data were presented: one at the lower end of this
temperature range in the vicinity of 185-187.5 °C, which were found by
Krigbaum and Uematsu (186±2°C) [57], Miller and Seeley (186°C) [58],
Krigbaum and Uematsu (186±2°C) [57], Miller and Seeley (186°C) [58],
Kamide and Toyama (187.5 °C) [59], Martuscelli, Pracella and Crispino (185
°C) [36], Bu, Cheng and Wunderlich (187.5 °C) [60] and Cheng et al. (185 °C)
[37]; one at the

37 In isotactic polypropylene higher end of this range, where a Tm° of around


208 °C was extrapolated by Fatou [61]. Samuels found that in i-PP fibers
multiple melting peaks arise when a certain minimum orientation of the non-
crystallizable chains is achieved [62]. The high temperature melting peak can be
extrapolated versus chain orientation along the fiber direction to an ultimate
melting point for fully oriented chains of 220 °C [62]. The low temperature
melting peak yields an extrapolated value of 185 °C. The establishment of the
AHf° of the a form in i-PP crystals also seems to be difficult. Literature data
range from 2.65kJ/mol to 10.9kJ/mol. The data obtained from calorimetry are
averaged from five data sources and yield a value of 6.9±0.8 kJ/mol [63], while
data from melting temperature depression due to the effect of a diluent generate
a higher value for the AH{° (9.1 ± 1.6 kJ/mol) [59,64]. A value of 8.7 ± 1.6
kJ/mol was proposed recently [60], which was indirectly supported by analyses
of microhardness [65] and calculations of the residual entropy of amorphous i-
PP at absolute zero. 2.2.2 Hexagonal lattice of β form Determination of the
crystal unit cell of the β form in i-PP was earlier based on WAXD experiments,
which showed two reflections at d-spacings of 0.553 nm and 0.417 nm [8,13].
These two peaks were believed to fit into a hexagonal lattice with a=1.274nm, c
= 0.635 nm, y = 120° and a = /? = 90°, similar to the case of the a form. They
thus correspond to (2 0 0) and (2 01) reflections in this hexagonal lattice [8]. It
was also suggested that the hexa¬ gonal lattice should possess a unit cell of a =
0.636 nm, c = 0.635 nm, y = 120 ° and οί = β = 90°. Therefore, these two
reflections should be (100) and (1 0 1) reflections [13]. Another report, however,
proposed that the hexagonal lattice should be a = 1.908 nm, c = 0.649 nm,
y=120° and a = /? = 90°, thus assigning (3 00) and (3 0 1) reflections to these
two d-spacings [7]. Samuels and Yee confirmed this hexagonal lattice based on
sixteen WAXD reflec¬ tions in the β form. The crystallographic density was
calculated to be 0.921 g/cm3 [66]. The chain packing in the β form was expected
to be parallel chains in groups with equal handedness, as shown in Figure 2.4.
This has recently been confirmed by 13C solid-state nuclear magnetic reso¬
nance (NMR) [67]. Some reports suggested that this metastable crystal form may
be recognized as a trigonal lattice, which could also be grouped as a subdivision
of the hexagonal lattice system. One important reason for the difficulty of
determining the crystal unit cell geometry is because the β form in i-PP is
thermodynamically less stable than the a form under normal crystallization
thermodynamically less stable than the a form under normal crystallization
conditions. In most cases, the β form can only be partially formed in samples
mixed with other crystal forms. The preparation of a rich or even pure β form in
i-PP has been reported through rapid quenching [7], zone solidification [68-70],
crystal¬ lization in a temperature gradient [71-76], or use of selective nucleating

38 Crystalline structures of polypropylene homo- and copolymers 0.636 nm


Figure 2.4 Projection of crystal unit cell of the β form in i-PP crystals, reported
b; Samuels and Yee [66] on the ab plane. agents [77-82]. It was also found that
during drawing, the β form in i-PI can be transformed into the a form at high
temperatures, while only i ‘smectic’ structure (see below) can be seen at low
temperatures [83]. Th< transformation of the β form into the a form can also be
carried out by mel recrystallization at elevated temperature close to the melting
temperature o the bulk samples [68, 83-85]. The structural similarity of the β
form anc the quenched (‘smectic’) form may also obscure this determination.
The thermodynamic properties of the β form have not been as extensivelj studied
as those of the a form. The Tm° was extrapolated by several researcl groups, and
values varied from 170 °C [62] to 200 °C [68]. Both 176 °C [81] and 184 °C
[86] were also suggested. The value of AH{° was even lesi certain, ranging
between 4.76kJ/mol [86] and 7.45kJ/mol [81]. 2.2.3 Triclinic lattice of y form
The crystal structure and unit cell geometry of the y form in i-PP are known only
approximately, but bear strong resemblance to those of the a form [7] The
proposed triclinic unit cell parameters are [10] a = 0.654nm, b = 2.14nm, c =
0.650nm, α = 89°, β = 99.6° and y = 99°. The crystallographic density was 0.954
g/cm3. This unit cell appears to result from the monoclinic a form by a simple
shear along the a axis [10]. The chain arrangement in the two forms is probably
analogous (Figure 2.5). However, recent confor¬ mational analysis on the crystal
structure of the y form has failed to reach a local minimum in the energy [87].

39 In isotactic polypropylene The y form can be observed in low molecular


weight i-PP [10] or crystal¬ lization of i-PP at elevated pressure above 200 MPa
[11, 88-90]. However, in most cases one may find coexistence of the a and y
forms in the crystals [91]. The y form is known to nucleate on crystals of the a
form, but the crystallographic relationship between these two forms is still not
fully understood. The first attempt by Padden and Keith [52, 53] was to establish
the mechanism of lamellar branching via homoepitaxy on the lateral ac faces of
the a form crystal laths. They suggested later that the y form may deposit on the
a form with a chain direction rotating by about 80°. This might initiate the
branching process based on experimental observation of thin film crystallization
of i-PP mixed with a paraffin (n-C32H66). Morrow and Newman performed
of i-PP mixed with a paraffin (n-C32H66). Morrow and Newman performed
crystallization experiments on low molecular weight i-PP fractions [10]. They
proposed a different growth pattern of the y form onto the a form crystals with
parallel chain axes of the two forms and mixing of forms within the individual
crystals. The crystallographic relation¬ ship between these two forms was
recently investigated by Lotz, Graff and Wittmann [92]. They found that the y
form crystals are elongated in the b* axis direction and have their chain axes
inclined at 40° to the lamellar surface normal in contrast to the a form crystals.
The y form crystals display screw dislocations. When the y form crystals are
nucleated on the a form crystals, the contact planes of both the a and y forms are
(010). They have their b* and c axes parallel. The y form crystal overgrowth
reflects the mirror symmetry that exists at the molecular level in the a form
crystals. Since the y form crystals have their lamellar surface inclined to the
chain axis, while in the a form crystals it is perpendicular to it, the y form
crystals Figure 2.5 Projection of crystal unit cell of the y form in i-PP crystals,
reported by Morrow and Newman [10] on the ab plane.

40 Crystalline structures of polypropylene homo- and copolymers grown on the


a form crystals appear to branch. However, the branching, or more precisely the
inclination, reveals only a difference in crystal morphol¬ ogy, and it does not
result from a different orientation of the a and y form chains, which are actually
parallel in the composite structure [92]. Finally, the thermodynamic properties of
this form, Tm° and AHf°, are not well determined. 2.2.4 Quenched form of i-PP
crystals The crystals grown during quenching in i-PP are the least understood. It
is interesting to note that rapid quenching of i-PP fails to produce totally
amorphous samples even when the cooling rate is as high as 10000 °C/min [93].
Natta and Corradini described this structure as ‘smectic’ and suggested that it is
composed of parallel 2 x 3/1 helices but that disorder exists in the packing of the
chains perpendicular to their axes [6]. Miller proposed that the quenched form is
‘paracrystalline’ [94], while Wunderlich and Grebowicz classified it as a
‘conformationally disordered (condis) crystal’ [95]. Gailey and Ralston [96] as
well as Gezovich and Geil [97] identified the quenched form as consisting of
very small hexagonal crystallites. Later, McAllister, Carter and Hinde [98]
suggested from their WAXD results that the quenched form represented
approximately 60% of the amorphous fraction with the remainder of the sample
having the helices arranged in a ‘square array with a cubic or tetragonal
symmetry’. Recently, Gomez, Tanaka and Tonelli reported that from their high
resolution 13C solid-state NMR study the local packing of the helices are very
similar in the quenched (‘smectic’) form and the β form in i-PP [67]. On the
other hand, Corradini et al [99] indicated that in quenched i-PP a fairly high
correlation of distances is present within each chain and between neighboring
correlation of distances is present within each chain and between neighboring
chains to form small bundles. The local correlations between chains are probably
nearer to those in the a form than to those in the β form [99]. Because of the
metastability of this form, no equilibrium thermodynamic properties have been
reported. 2.3 CRYSTAL STRUCTURE OF SYNDIOTACTIC
POLYPROPYLENE s-PP was first synthesized during the 1960s, and its
configuration and conformation were characterized by Natta and colleagues [14-
17]. It was the first α-olefin polymer to be obtained in a highly regular
syndiotactic form. The syndiotacticity can be studied by IR and NMR techniques
[16,18,100-110]. 2.3.1 High temperature orthorhombic form The crystal
structure and unit cell geometry of s-PP was determined by Natta et al. [14] from
WAXD studies on oriented fibers. The unit cell

Of syndiotactic polypropylene 41 was proposed to be orthorhombic with


a=1.45nm, h = 0.580nm (or b = 0.560 nm [17]), c = 0.740 nm and a = /? = y =
90°, and space group of C2221. The crystallographic density was calculated to
be 0.90 g/cm3, slightly lower than the density of isotactic polypropylene (0.936
g/cm3). Along the ident¬ ity period c axis, four monomer units are involved, and
each unit cell contains eight units (two parallel chains). The chain predominantly
adopts a two-fold helical form, 4x2/1. Figure 2.6 illustrates the chain conforma¬
tion (ttgg)2 in the s-PP unit cell with a projection of the chain on the ab plane.
Right-handed and left-handed helices in s-PP chain molecules may also affect
the crystal packing. In the lattice, the chain symmetry is fully maintained and
packing depends on the bulkiness of the methyl groups. For a C-centered unit
cell, chains of the same handedness are required. Therefore, the lattice is fully
isoclined with all molecules of the same handedness of a be plane. During the
late 1980s, Lotz, Lovinger and Cais [18] proposed an ortho¬ rhombic unit cell
with a = 1.45 nm, b = 1.12 nm, c = 0.740 nm and <χ = β = γ = 90°, and a space
group of Ibca based on their electron diffraction (ED) results obtained from
TEM. Additional reflections and streaks from the single crystals produced by
thin-film single crystal growth revealed features not seen in the fiber patterns
studied previously by Natta’s group. A very strong subcell was recognized with
b = 0.560 nm, which can be identified through powder WAXD patterns (20=
18.8°) for s-PP crystallized at differ¬ ent temperatures [111, 112]. The
‘doubling’ of the unit cell is a consequence of a different packing scheme. In the
new unit cell symmetry, packing of both left-handed and right-handed helices is
allowed. Figure 2.7 illustrates this new packing in an anticlined form as
compared to the original isoclined packing. Two packing schemes are possible,
which depend on the handed¬ ness of the depositing chain and those already on
the crystal growth face. All these results were confirmed for polycrystalline thick
the crystal growth face. All these results were confirmed for polycrystalline thick
films and bulk samples obtained directly from polymerization, as well as for
uniaxially oriented s-PP [19]. Figure 2.6 Projection of the high temperature
orthorhombic s-PP crystal form reported by Natta et al. [14] on the ab plane.

42 Crystalline structures of polypropylene homo- and copolymers Figure 2.7


Two projections of the high temperature orthorhombic s-PP crystal form
reported by Lovinger, Davis and Lotz [20] on the ab planes: (a) helices in
adjacent sheet are of opposite handedness; (b) helices in each sheet alternate in
handedness. Most recently, Lovinger, Davis and Lotz [20] investigated the
tempera¬ ture dependence of the s-PP crystal structure and morphology and its
epitaxial relationship with i-PP. At the highest crystallization temperature region
(Tc> 105 °C) large, rectangular and highly regular single crystals were observed
by phase-contrast light microscopy. These crystals had the same unit cell and
interchain packing as previously observed. As Tc was decreased the morphology
became axialitic and eventually spherulitic (in the vicinity of Tc = 70 °C). It was
suggested that in this lower crystallization temperature region, intermolecular
packing disorder with both handedness is introduced due to kinetic
considerations in the crystallization process. Overgrowths observed by TEM on
the s-PP crystals were also identified as individual i-PP crystals. This
phenomenon was believed to be initiated by a specific epitaxial relationship with
the lateral surfaces of s-PP lamellae.

Of syndiotactic polypropylene 43 The thermodynamic properties of s-PP have


been reported by several researchers. Boor and Youngmann [107] estimated Tm°
to be in the range 140-170 °C. Values of 151 °C and 155 °C for s-PP with
relatively short sequences were published by Haftka and Konnecke [106]. Miller
and Seeley determined a Tm° of 158.8 °C using the Hoffman-Weeks method,
which utilizes observed Tm and Tc extrapolation [58], while Cheng and
Rodriguez obtained a Tm° of 160 °C through both SAXS and DSC experimental
results [113]. A calculated ΔHf° of about 1.88kJ/mol was obtained assuming
amorphous and unit cell densities of 0.854 and 0.90 g/cm3, respectively [105,
107]. The higher end value of the ΔH{° was reported to be 8.26 kJ/mol using
WAXD and DSC data [106]. Very recently, a AH{° of 7.9 kJ/mol was achieved
through extrapolation methods [113]. As compared with i-PP the syndiotactic
form has a lower unit cell density, a lower melting temperature and a lower heat
of fusion. This may be important when the ultimate properties of s-PP are
discussed. 2.3.2 Low temperature, quenched orthorhombic form The helical
conformation of s-PP chains is the most stable one. However, an unstable form
was also observed by Natta, Peraldo and Allegra [21], consisting of planar
was also observed by Natta, Peraldo and Allegra [21], consisting of planar
zigzag polymer chains. The samples were obtained by cold-drawing polymer
quenched from the melt. The helical form was completely absent. Upon
annealing at about 100 °C the samples were con¬ verted from the zigzag planar
to the helical form. In more recent work [22] a c axis (fiber direction) of 0.505
nm in a planar zigzag form was obtained. From WAXD fiber patterns an
orthorhombic unit cell of a = 0.522nm, h = 1.117nm, c — 0.506 nm and <χ = β =
γ = 90° was determined. Assuming that four monomer units are contained in
each unit cell, the calculated crystallographic density was 0.945 g/cm3 (the
observed density is in the vicinity of 0.885 g/cm3). The crystallite size in this
quenched form was extremely small, approximately a few nanometers. Figure
2.8 shows the Figure 2.8 Projection of the low temperature orthorhombic s-PP
crystal form reported by Chatani et al [22] on the ab plane.

44 Crystalline structures of polypropylene homo- and copolymers b Figure 2.9


Projection of the triclinic s-PP crystal form reported by Chatani et al [23] on the
ab plane. crystal structure and unit cell geometry with the planar zigzag chain
conformation. No equilibrium thermodynamic property data are available for
this crystal form due to its thermodynamic metastability. 2.3.3 Triclinic form In
some of the WAXD fiber patterns a reflection indicating a periodic d-spacing of
about 0.28 nm or a multiple of this value suggested the existence of a third
crystalline phase. Chatani et al. [23] reported a new crystalline form, a triclinic
lattice of a = 0.572nm, b = 0.764nm, c=1.16nm, a = 73.1°, /? = 88.8° and y =
112.0°, with a space group of PI. The crystal structure is shown in Figure 2.9.
The crystallographic density was calculated to be 0.939 g/cm3 where six
monomer units are included, while the observed density was in the vicinity of
0.89 g/cm3. The chain does not have centers of symmetry. The two
conformations suggested are formed first by deforming the (ttgg)2 helix,
resulting in a chain repeat of three (ttgg) in two turns of the helix, and second
through a (t6g2t2g2) conformation. The latter was con¬ sidered as an
intermediate state between planar zigzag and helical forms. As with the planar
zigzag form, this third crystalline phase may be transformed into the helical form
by annealing above 50 °C and/or solvent vapor absorption [23]. Thermodynamic
properties have not been reported. 2.4 CRYSTAL STRUCTURE OF
POLYPROPYLENE COPOLYMERS Copolymers containing PP can be divided
into two different types based on chemical sequences: block copolymers and
random copolymers. In the block copolymers (including grafting copolymers),
crystallizable repeating unit sequences such as PP can easily build up
crystallinity with its own crystal structure and unit cell geometry, as in the case
of the corresponding
of the corresponding

Of polypropylene copolymers 45 homopolymer. If both comonomer units are


crystallizable, two different crystal lattices can be found. In contrast, the random
copolymers may form mixed crystals sharing a common crystal lattice, called
isomorphism, or else only one of the comonomer units may crystallize and
another one is rejected into the amorphous region. In the latter case, a sharp
reduction of crystal¬ linity with increasing concentration of repeating units
which cannot undergo isomorphous substitution is evident. In this section, only a
few extensively studied random copolymers containing PP, which are important
in indus¬ trial applications, are reviewed. 2.4.1 Copolymers with different
compositions The most common copolymers in this category are ethylene-
propylene copolymers. It is well known that the propylene units can be
incorporated into the crystal unit cell of the copolymers as defects [24-28].
However, since the different repeating units are nonequivalent, large changes in
lattice parameters are found. Early studies showed variation of the a axis length
of the PE unit cell with copolymer concentration when the propylene units were
less than 15% (Figure 2.10) [27, 28]. The b and c axes remain practi¬ cally
constant, while the crystal structure remains that of PE for all concentrations of
PP units studied. There was also some evidence for the formation of triclinic PE
crystals [28]. These results have been confirmed in recent years by several
research groups [115-117]. In general, the shorter side chains (such as
propylene) were incorporated in somewhat larger amounts in the crystals. The
overall small effect of longer side chains on the Figure 2.10 Crystal dimension of
the a axis of orthorhombic PE as a function of propylene comonomer
concentration [24-28].

46 Crystalline structures of polypropylene homo- and copolymers expansion of


the crystal unit cell (often along a axis) thus finds an explanation in a reduced
incorporation of defects within the crystals. The decrease in crystallinity upon
increasing the propylene concentration in the copolymers can be observed [27,
63]. The methyl groups reduce the crystallinity after the initial, more rapid
reduction by about four to five additional CH2 units. About 18-20% of the
methyl groups are necessary to reduce the crystallinity towards zero [63]. A
depression of greater than 60 °C in the melting temperature was reported for the
introduction of 10% propylene groups into the copolymers [63,118]. Isotactic
copolymers of propylene with ethylene, 1-butene and other short branchings
were also studied [30]. One example is propylene-1-butene copolymers with
different propylene concentrations. Poly-1-butene has two crystal forms. Form 1
(trigonal) possesses a 2x3/1 helix and form II (tetragonal) a 2 x 4/1 helix. Form I
is thermodynamically more stable than form II, and form II can be transferred to
is thermodynamically more stable than form II, and form II can be transferred to
form I during heating or drawing. Propylene comonomer units decrease the poly-
1-butene form I lattice dimension, as shown in Figure 2.11. At 44% propylene
concentration the cross-sectional area of the helix approaches that of the
expanded PP with 56% butene-1. In the middle range of concentration both
crystal structures exist side by side (the crystal structure and unit cell geometry
of poly- 1-butene is trigonal with a = b= 1.77 nm and c = 0.65 nm, with a space
group of R3c [1]). At lower concentrations only crystal structures of the major
components were observed. On increasing the concentration, the unit cell
parameters of the host crystal changed continuously, with a match of unit Mole
% comonomer Figure 2.11 Crystal dimension of the a axis of trigonal poly-1-
butene as a function of propylene comonomer concentration [30].

Of polypropylene copolymers 47 cell parameters of both components in the


middle concentration range. The crystallinity drops from about 65% for the
homopolymers to about 40% for the middle range of composition, but they
remain crystalline over the whole concentration range with a minimum AHf at
around 45% 1-butene comonomer units [31]. The dissolution and melting
temperatures of propylene-1-butene copolymers as functions of concentration
are shown in Figure 2.12. Finally, it is worth mentioning that chlorinated i-PP
shows increased unit cell parameters similar to those of PE [32, 33]. 2.4.2
Copolymers with different tacticities A special case of helix interruption is
presented by the occurrence of different stereo-isomers of propylene units along
the chain. Isotactic-PP crystals con¬ tain all four possible combinations of
sequences of d- and /-type repeating unit structure with left- and right-handed
helices. Any particular change of a once established sequence within the crystal
is forbidden. The condition of crystall¬ ization is thus the presence of an
uninterrupted sequence of isotactic configu¬ rations with sufficient length. The
distribution of this interruption is also particulary important since it affects the
crystal structure and crystallization process. Special efforts were made to
separate the tacticity effect from other factors such as molecular weight and
molecular weight distribution. The cry¬ stallization and melting behavior of a
series of five i-PP fractions with similar molecular weight and molecular weight
distribution but different isotacticities was studied recently [37, 119, 120]. It was
found that the crystal structure 1-butene concentration (mole %) Figure 2.12
Dissolution and melting temperatures of propylene-1-butene copolymers as a
function of 1-butene concentration [31].

48 Crystalline structures of polypropylene homo- and copolymers parameters of


the i-PP change with isotacticities, as illustrated in Figure 2.13, which shows the
the i-PP change with isotacticities, as illustrated in Figure 2.13, which shows the
relationship between the WAXD intensity of the powder patterns and the
diffraction angle 20. Not only the crystallinity and apparent crystallite size
decrease with isotacticity, but the crystal structure and unit cell geometry are
changed as well. For all the fractions the crystal lattice is monoclinic.
Nevertheless, the unit cell size and angle (β) change with supercooling (AT) for
individual fractions as well as isotacticities. Figure 2.14 shows two examples
with isotacticities of 98.8% and 78.7% [37]. In the supercooling region of 20-90
°C, the changes in unit cell sizes and angles exhibit a similar tendency. For
instance, the a axis of the unit cell of the i-PP fraction with higher isotacticity
changes from 0.678 nm at AT= 83 °C to 0.670 nm at AT=28 °C, while its
counterpart fraction with lower isotacticity decreases from 0.683 nm to 0.679
nm. The b axis changes from 2.122 nm to 2.116 nm for the i-PP fraction with
lower isotacticity, and from 2.118 nm to 2.109 nm for the i-PP fraction with
higher isotacticity in the same supercooling region. The c axis, however, is
essentially unchanged, Figure 2.13 WAXD powder patterns for the i-PP
fractions with different isotactici¬ ties crystallized at different Tc [37].

Of polypropylene copolymers 49 Figure 2.14 The crystal unit cell parameters (a,
b, c and β) change with AT for two i-PP fractions with 98.8% and 78.7%
isotacticities [37]. as 0.655 + 0.001 nm. The angle β, on the other hand, changes
from 98.5° to 98.8° for the i-PP fraction with lower isotacticity, and from 98.8°
to 99.2° for the i-PP faction with higher isotacticity. The crystallinity and
apparent crystallite size also changes with different supercooling and isotacticity
[37]. The thermodynamic properties of the i-PP which possess different isotac¬
ticities with a random distribution sequence of stereo-defects were also reported
[37]. The equilibrium melting temperatures of the i-PP fractions were obtained
through two extrapolation methods. The first method is based on a Hoffman-
Weeks extrapolation (Tm versus Tc plots). The second method is extrapolated
from the relationship between lamellar thickness obtained from small angle X-
ray scattering (SAXS) and observed Tm based on the Thomson-Gibbs equation.
Similarly, these two methods can be adopted in obtaining the equilibrium heat of
fusion. The data generated by these two methods show reasonable agreement.
Discussion can be further extended to the molecular aspects of the crystallization
behavior of these i-PP stereo-copolymers. Theories of random copolymer
crystallization have been developed for the melting

50 Crystalline structures of polypropylene homo- and copolymers temperature


[121-125], the kinetics of crystallization [122,124] and the amount of
noncrystallizable comonomer that is incorporated into the crystalline phase
[123]. Two extreme cases are usually discussed: uniform inclusion and complete
[123]. Two extreme cases are usually discussed: uniform inclusion and complete
exclusion of comonomers. The former can be described as the crystalline phase
of a solid solution for both comonomers (A and B); the B comonomers produce
defects in the crystalline A lattice and both crystalline and amorphous phases
have the same com¬ position. The latter is concerned with the crystalline phase
being composed entirely of A comonomers and in metastable equilibrium with a
mixed amorphous phase of A comonomers and noncrystallizable B
comonomers. The detailed calculation based on the experimental data indicates
that the crystallization behavior of this series of i-PP fractions with different
isotacticities is close to the inclusion of the comonomers [37]. The equilibrium
melting temperature decreases with isotacticty. The energy of the stereo-defects
introduced due to the decrease of isotacticity also increases. This leads to a total
decrease of the equilibrium heats of fusion in this series of i-PP fractions [37].
Stereoblock PP, on the other hand, shows isotacticity sequence dis¬ tributions,
first identified by Natta and colleagues [126,127]. It was found that the
alternating isotactic and stereo-irregular sequences can be approxi¬ mately
described by second-order Markov polymerization statistics [128, 129]. The
isotactic blocks in the PP may still crystallize, depending upon the block length
[129], with a lattice structure of i-PP crystals. The morphology consists of
continuous hard and soft phases of crystalline i-PP and amor¬ phous, atactic
polypropylene (a-PP), respectively [129]. This leads to an indication of
elastomeric properties. Blends can be obtained from the stereoblock PP mixed
with i-PP in different molecular weights, molecular weight distributions and
concentrations to achieve a wide range of physical properties for specific
applications [130]. 2.5 ACKNOWLEDGEMENTS This work was supported by
the National Science Foundation’s Presidential Young Investigator Award
(DMR-9157738) to S.Z.D.C. and by Industrial Matching Funding from Exxon
Education Foundation. 2.6 APPENDIX: SYMBOLS AND ABBREVIATIONS
a-PP Atactic polypropylene DSC Differential scanning calorimetry ED Electron
diffraction i-PP Isotactic polypropylene IR Infrared

NMR References Nuclear magnetic resonance PE Polyethylene PLM Polarized


light microscopy PP Polypropylene SAXS Small angle X-ray scattering s-PP
Syndiotactic polypropylene Tc Crystallization temperature TEM Transmission
electron microscopy Tm Crystal melting temperature rp O ■* m Equilibrium
melting temperature WAXD Wide angle X-ray diffraction AHf Heat of fusion
AH{° Equilibrium heat of fusion AT Supercooling, AT = Tm° — Tc 2.7
REFERENCES 1. Natta, G. and Corradini, P. (1959) Journal of Polymer
Science, 39, 29. 2. Natta, G. and Corradini, P. (I960) Nuovo Cimento, Supply
15, 9. 3. Natta, G., Corradini, P. and Ganis, P. (1960) Makromolekulare Chemie
15, 9. 3. Natta, G., Corradini, P. and Ganis, P. (1960) Makromolekulare Chemie
39, 238. 4. Natta, G., Corradini, P. and Ganis, P. 1962 Journal of Polymer
Science, 58, 1191. 5. Natta, G., Dall’Asta, G., Mazzanti, G., et al. (1962)
Makromolekulare Chemie, 54, 95. 6. Natta, G. and Corradini, P. (1960) Nuovo
Cimento, Supply 15, 40. 7. Turner-Jones, A., Aizlewood, J. M. and Beckett, D.
R. (1964) Makromolekulare Chemie, 75, 134. 8. Keith, H. D, Padden, Jr, F. J,
Walker, N. M. and Wyckoff, H. W. (1959) Journal of Applied Physics, 30, 1485.
9. Turner-Jones, A. and Cobbold, A. J. (1968) Journal of Polymer Science, Part
B Polymer Letters, 6, 539. 10. Morrow, D. R. and Newman, B. A. (1968)
Journal of Applied Physics, 39, 4944. 11. Kardos, J. L., Christiansen, A. W. and
Baer, E. J. (1966) Journal of Polymer Sciences, (Part A2\ 4, 111. 12. Sauer, J. A.
and Pae, K. D. (1968) Journal of Applied Physics, 30, 4950. 13. Addink, E. J.
and Bientema, J. (1961) Polymer, 2, 185. 14. Natta, G., Pasquon, I., Corradini, P.
et al (1960) Atti dellAccademia Nazionale dei Lincei. Rendiconti, 28, 539. 15.
Natta, G., Pasquon, I. and Zambelli, A. (1962) Journal of the American
Chemical Society, 84, 1488. 16. Zambelli, A., Natta, G. and Pasquon, I. (1963)
Journal of Polymer Science, Part C, 4,411. 17. Corradini, P., Natta, G., Ganis, P.
and Temussi, P. A. (1967) Journal of Polymer Science, Part C, 16, 2477. 18.
Lotz, B., Lovinger, A. J. and Cais, R. E. (1988) Macromolecules, 21, 2375. 19.
Lovinger, A. J., Lotz, B. and Davis, D. (1990) Polymer, 31, 2253. 20. Lovinger,
A. J., Davis, D. and Lotz, B. (1991) Macromolecules, 24, 552.

52 Crystalline structures of polypropylene homo- and copolymers 21. Natta, G.,


Peraldo, M. and Allegra, G. (1964) Makromolekulare Chemie, 75, 215. 22.
Chatani, Y. Maruyama, H., Noguchi, K. et al (1990) Journal of Polymer Science,
Part C, Polymer Letters, 28, 393. 23. Chatani, Y., Maruyama, H., Asanuma, T.
and Shiomura, T. (1991) Journal of Polymer Science, Polymer Physics Edition,
29, 1649. 24. Walter, E. R. and Reding, F. P. (1956) Journal of Polymer Science,
20, 561. 25. Cole, E. A. and Holmes, D. R. (1960) Journal of Polymer Science,
46, 245. 26. Swan, P. R. (1962) Journal of Polymer Science, 56, 23. 27.
Wunderlich, B. and Poland, D. (1963) Journal of Polymer Science, Part A, 1,
357. 28. Baker, C. H. and Mandelkern, L. (1966) Polymer, 7, 71. 29. Patel, G. N.
and Keller, A. (1975) Journal of Polymer Science, Polymer Physics Edition, 13,
2281. 30. Turner-Jones, A. (1966) Polymer, 7, 23. 31. Cavallo, P., Martuscelli,
E. and Pracella, M. (1977) Polymer, 18, 42, 891. 32. Plate, N. A., Khen, T. and
Shibayev, V. P. (1967) Journal of Polymer Science, Polymer Symposia, 16,
1133. 33. Plate, N. A., Shibayev, V. P., Petrukhin, B. S. and Kargin, V. A.
(1968) Journal of Polymer Science, Polymer Symposia, 23, 37. 34. Newman, S.
(1960) Journal of Polymer Science, 47, 111. 35. Pavan, A., Provasoli, A.,
Moraglio, G. and Zambelli, A. (1977) Mak¬ romolekulare Chemie, 178, 1099.
Moraglio, G. and Zambelli, A. (1977) Mak¬ romolekulare Chemie, 178, 1099.
36. Martuscelli, E., Pracella, M. and Crispino, L. (1983) Polymer, 24, 693. 37.
Cheng, S. Z. D., Janimak, J. J., Zhang, A.-Q. and Hsieh, E. T. (1991) Polymer,
32, 648. 38. Mencik, Z. (1972) Journal of Macromolecular Science, Phys. B(6),
101. 39. Hikosaka, M. and Seto, T. (1973) Polymer Journal 5, 111. 40.
Petraccone, V., De Rosa, C., Guerra, G. and Tuzi, A. (1984) Makromolekulare
Chemie, Rapid Communications, 5, 631. 41. Guerra, G., Petraccone, V.,
Corradini, P. et al. (1984) Journal of Polymer Science, Polymer Physics Edition,
22, 1029. 42. De Rosa, C., Guerra, G., Napolitano, R. et al. (1984) European
Polymer Journal, 20, 937. 43. Petraccone, V., Guerra, G., De Rosa, C. and Tuzi,
A. (1985) Macromolecules, 18, 813. 44. Napolitano, R., Pirozzi, B. and Varriale,
V. (1990) Journal of Polymer Science, Polymer Physics Edition, 28, 139. 45.
Padden, Jr, F. J. and Keith, H. D. (1959) Journal of Applied Physics, 30, 1479.
46. Norton, D. R. and Keller, A. (1985) Polymer, 26, 704. 47. Awaya, H. (1988)
Polymer, 29, 591. 48. Olley, R. H. and Bassett, D. C. (1989) Polymer, 30, 399.
49. Binsbergen, F. L. and DeLange, B. G. M. (1968) Polymer, 9, 23; (1970)
Polymer, 11, 309. 50. Khoury, F. (1966) Journal of Research of the National
Bureau of Standards, Part A, 70, 29. 51. Geil, P. H. (1963) Polymer Single
Crystals, Wiley-Interscience, New York. 52. Padden, F. J. and Keith, H. D.
(1966) Journal of Applied Physics, 37, 4013. 53. Padden, F. J. and Keith, H. D.
(1973) Journal of Applied Physics, 44, 1217. 54. Lovinger, A. J. (1983) Journal
of Polymer Science, Polymer Physics Edition, 21, 97. 55. Bassett, D. C. and
Olley, R. H. (1984) Polymer, 25, 935.

References 53 56. Lotz, B. and Wittmann, J. C. (1986) Journal of Polymer


Science, Polymer Physics Edition, 24, 1541. 57. Krigbaum, W. R. and Uematsu,
I. (1965) Journal of Polymer Science, Part A, 3, 767. 58. Miller, R. L. and
Seeley, E. G. (1982) Journal of Polymer Science, Polymer Physics Edition, 20,
2297. 59. Kamide, K. and Toyama, K. (1969) Kobunshi Kaguku, 25, 49. 60. Bu,
H.-S., Cheng, S. Z. D. and Wunderlich, B. (1988) Makromolekulare Chemie,
Rapid Communications, 9, 76. 61. J. G. Fatou, (1971) European Polymer
Journal, 7, 1057. 62. Samuels, R. J. (1975) Journal of Polymer Science, Polymer
Physics Edition, 13, 1417. 63. Wunderlich, B. (1980) Macromolecular Physics,
Crystal Melting, Vol. 3, Aca¬ demic Press, New York. 64. Danusso, F. and
Gianotti, G. (1968) European Polymer Journal, 4, 165. 65. Balta-Calleja, F. J.,
Martinez-Salazar, J. and Zsano, T. (1988) Journal of Mate¬ rials Science Letters,
7, 165. 66. Samuels, R. J. and Yee, R. Y. (1972) Journal of Polymer Science,
Part A2, 10, 385. 67. Gomez, M. A., Tanaka, H. and Tonelli, A. E. (1987)
Polymer, 28, 2227. 68. Lovinger, A. J., Chua, J. O. and Gryte, C. C. (1977)
Journal of Polymer Science, Polymer Physics Edition, 15, 641. 69. Fujiwara, Y.
Journal of Polymer Science, Polymer Physics Edition, 15, 641. 69. Fujiwara, Y.
(1968) Kolloid Zeitschrift und Zeitschrift fur Polymere, 226, 135. 70. Crissman,
J. M. (1969) Journal of Polymer Science, Polymer Physics Edition, 7, 389. 71.
Fujiwara, Y. (1975) Colloid and Polymer Science, 253, 273. 72. Asano, T. and
Fujiwara, Y. (1978) Polymer, 19, 99. 73. Asano, T., Fujiwara, Y. and Yoshida,
T. (1979) Polymer Journal, 11, 383. 74. Yoshida, T., Fujiwara, Y. and Asano, T.
(1983) Polymer, 24, 925. 75. Fujiwara, Y. (1987) Colloid Polymer Science, 265,
1027. 76. Morrow, D. R. (1969) Journal of Macromolecular Science, Physics,
133, 53. 77. Leugering, Η. I. (1967) Makromolekulare Chemie, 109, 204. 78.
Duswalt, A. A. and Cox, W. W. (1970) ACS Division of Organic Coatings, 30,
93. 79. Ulmann, W. and Wendorff, J. H. (1979) Progress in Colloid and Polymer
Science, 66, 25. 80. Moos, K. H. and Tilger, B. (1981) Angewandte
Makromolekulare Chemie, 94, 213. 81. Shi, G.-Y., Huang, B. and Zheng, J.-Y.
(1984) Makromolekulare Chemie Rapid Communications, 5, 573. 82. Jacoby, P.,
Bersted, B. H., Kissel, W. J. and Smith, C. E. (1986) Journal of Polymer
Science, Polymer Physics Edition, 24, 461. 83. Shi, G.-Y., Chu, F., Zhou, G.-E.
and Han, Z.-W. (1989) Makromolekulare Chemie, 190, 907. 84. Varga, J. (1989)
Journal of Thermal Analysis, 35, 1891. 85. Varga, J., Fujiwara, Y. and Ille, A.
(1990) Periodica Polytechnica, Chemical Engineering, 34, 256. 86. Varga, J. and
Garzo, G. (1991) Acta Chimica Hungarica, 128, 303. 87. Corradini, P.,
Petraccone, V. and Pirozzi, B. (1983) European Polymer Journal, 19, 299. 88.
Baer, E. and Kardos, J. L. (1965) Journal of Polymer Science, Part A, 3, 2827.

54 Crystalline structures of polypropylene homo- and copolymers 89. Leute, U.,


Dollhopf, W. and Liska, E. (1978) Colloid and Polymer Science, 256, 914. 90.
Nakafuku, C. (1981) Polymer, 22, 1673. 91. Kojima, M. (1968) Journal of
Polymer Science, Part A2, 6, 1255. 92. Lotz, B., Graff, S. and Wittmann, J. C.
(1986) Journal of Polymer Science, Polymer Physics Edition, 24, 2017. 93. Wu,
Z.-Q., Veronica, D. L., Cheng, S. Z. D. and Wunderlich, B. (1988) Journal of
Thermal Analysis, 34, 105. 94. Miller, R. L. (1960) Polymer, 1, 135. 95.
Wunderlich, B. and Grebowicz, J. (1984) Advances in Polymer Science, 60/61,
1. 96. Gailey, J. A. and Ralston, R. H. (1964) SPE Transactions, 4, 29. 97.
Gezovich, D. M. and Geil, P. H. (1968) Polymer Engineering Science, 8, 202.
98. McAllister, P. B., Carter, T. J. and Hinde, R. M. (1978) Journal of Polymer
Science, Polymer Physics Edition, 16, 49. 99. Corradini, P., Petraccone, V., De
Rosa, C. and Guerra, G. (1986) Macromole¬ cules, 19, 2699. 100. Youngmann,
E. A. and Boor, J. (1967) Macromolecules Review, 2, 33. 101. Grant, I. J. and
Ward, I. M. (1965) Polymer, 6, 223. 102. Inagaki, H., Miyamoto, T. and Ohta, S.
(1966) Journal of Physical Chemistry, 70, 3420. 103. Gee, D. R. and Melia, T. P.
(1969) Polymer, 10, 239. 104. Tadokoro, H., Kobayashi, M. and Kobayashi, S.,
et al. (1966) Reprints of Progress in Polymer Physics Japan, 9, 181. 105. Boor, J.
et al. (1966) Reprints of Progress in Polymer Physics Japan, 9, 181. 105. Boor, J.
and Youngmann, E. A. (1966) Journal of Polymer Science, Part Al, 4, 861. 106.
Haftka, S. and Konnecke, K. (1991) Journal of Macromolecular Science,
Physics, B30, 319. 107. Boor, J. and Youngmann, E. A. (1965) Journal of
Polymer Science, Part B, 3, 577. 108. Ferguson, R. C. (1967) Transactions of
New York Academy of Sciences, 29, 495. 109. Heatley, F. and Zambelli, A.
(1969) Macromolecules, 2, 618. 110. Tonelli, A. E. and Schilling, F. C. (1981)
Accounts of Chemical Research, 14, 233. 111. Galambos, A., Wolkowicz, M.,
Zeigler, R. and Galimberti, M. (1991) ACS Polymer Materials Science and
Engineering, 64, 45. 112. Lovinger, A. J., Lotz, B. and Davis, D. D. (1992) ACS
Polymer Preprints, 33(1), 270. 113. Rodriguez, J., Cheng, S. Z. D., Lovinger, A.,
Hsieh, E. T., Chu, P., Johnson, T. W., Honnell, K. G., Geerts, R. G., Palackal, S.
J., Hawley, G. R., Welch, Μ. B. (1964) Polymer, 35, 1884. 114. Natta, G.,
Peraldo, M. and Allegra, G. (1964) Makromolekulare Chemie, 75, 215. 115.
Balta-Calleja, F. J. and Hidalgo, A. (1969) Kolloid Zeitschrift und Zeitschrift fur
Polymere, 229, 21. 116. Balta-Calleja, F. J. and Rueda, D. R. (1974) Polymer
Journal, 3, 216. 117. Martinez de Salazar, J. and Balta-Calleja, F. J. (1979)
Journal of a Crystal Growth, 48, 283. 118. Griskey, R. G. and Foster, G. N.
(1970) Journal of Polymer Science, Polymer Chemistry Edition, 8, 1623. 119.
Janimak, J. J., Cheng, S. Z. D., Giusti, P. A. and Hsieh, E. T. (1991) Macro¬
molecules, 24, 2253. 120. Janimak, J. J., Cheng, S. Z. D., Zhang, A.-Q. and
Hsieh, E. T. (1992) Polymer, 33, 728.

References 55 121. Flory, P. J. Transactions of the Faraday Society, 51, 848.


122. Sanchez, I. C. and Eby, R. K. (1973) Journal of Research of the National
Bureau of Standards, Part A, 77, 353. 123. Helfand, E., and Lauritzen Jr, J. I.
(1973) Macromolecules, 6, 631. 124. Sanchez, I. C. and Eby, R. K. (1975)
Macromolecules, 8, 639. 125. Sanchez, I. C. (1977) Journal of Polymer Science,
Polymer Symposia, 59, 109. 126. Natta, G. (1959) Journal of Polymer Science,
34, 531. 127. Natta, G., Mazzanti, G., Crespi, G. and Moraglio, G. (1957) Chim.
Ind. Milano, 39, 27. 128. Collette, J. W., Tullock, C. W. and MacDonald, R. N.
et al. (1989) Macro¬ molecules, 22, 3851. 129. Collette, J. W., Ovenall, D. W.,
Buck, W. H. and Ferguson, R. C. (1989) Macro¬ molecules, 22, 3858. 130. Su,
A. C. L. and Shih, C. K. (1987) High molecular weight stereoblock poly¬
propylene, in Thermoplastic Elastomers (eds N. R. Legge, G. Holden and Η. E.
Schroeder), Hanser, New York.

3 Crystallization, melting and supermolecular structure of isotactic


polypropylene J. Varga 3.1 INTRODUCTION Isotactic polypropylene (IPP) was
the first and is still the most important type of stereoregular polymer [1-3]. IPP is
the first and is still the most important type of stereoregular polymer [1-3]. IPP is
a semicrystalline polymer; its high tendency to crystallization is due to its
regular chain structure. A similar tendency to crystallization is implied in
propylene copolymers with relatively low comonomer content (ethylene, α-
olefins), of which there are two types: block copolymers (BC-PP) and random
copolymers (RC-PP). IPP and propylene copolymers with low comonomer
content are polymorphic materials having several crystal modifications [4], such
as monoclinic (a), hexagonal (β) and triclinic (y) forms. The crystallographic
problems of polymorphism of IPP were recently reviewed by Bruckner et al [4].
By melt quenching, a product with intermediate orderedness is formed, which
has two diffuse bands in the X-ray diffractogram and is often called smectic.
There have been many different explanations of its structure and even its
designation in the literature [4]. Because of its high technical and economic
significance and scientific interest, the literature for IPP is very rich. Next to
polyethylene, IPP is the most important subject of research in polymer physics
[5-7]. Investigations of IPP have contributed considerably to the understanding
of the crystal structure [5], the kinetics of crystallization [6], the melting
characteristics [7] and the different Polypropylene: Structure, blends and
composites. Edited by J. Karger-Kocsis. Published in 1995 by Chapman & Hall,
London. ISBN 0 412 58430 1

57 oi-Modification of IPP supermolecular structures (spherulite, cylindrite,


hedrite, quadrite) of poly¬ mers. In this chapter this topic is covered in
association with the author’s field of interest, which makes the treatment rather
subjective, emphasizing more recent results concerning the ^-modification of
IPP. Scientific research is focused on a-IPP because the conventional IPP grades
crystallize into this form under the thermal conditions that occur in practice.
/MPP and y-IPP were formerly regarded as curiosities in the laboratory, but
interest in them increased markedly during the 1980s: /MPP is discussed in [8-
79], while y-IPP is mentioned in [11, 12, 51, 80-97]. The technical and economic
significance of both modifications is expected to increase in the future, as
suggested by some recent investiga¬ tions [67, 72-75] into the feasibility of
manufacturing /J-IPP based prod¬ ucts, including multicomponent systems [73-
75], under normal processing conditions. In some fields, such as film forming,
the application of /MPP is definitely favourable [72]. The practical importance
of y-IPP stems from the observation that certain RC-PP grades crystallize
preferentially in y form. It is likely, therefore, that both /MPP and y-IPP will be
used in industry. Moreover, they might be come complementary grades for a-IPP
in the future. 3.2 a-MODIFICATION OF IPP 3.2.1 Crystallization of a-IPP
Crystallization is a first-order thermodynamic phase transition. The general rules
Crystallization is a first-order thermodynamic phase transition. The general rules
of crystallization for low-molecular weight substances are valid for polymers.
However, because of the large molecular size of polymers, their anisometric
geometry (thread-like molecules) and their polymolecularity in terms of size and
chemical structure, several specific features should also be considered [6]. A
kinetic theory of polymer crystallization [6, 8, 98] was developed, and has since
been updated and modified, and even diversified by alternative theories (see
[99]). For polymers, primary and secondary crystallization are usually distin¬
guished. Primary crystallization is the transition of a amorphous molten
substance into a crystalline phase. However, the structure formed during a
primary crystallization is in a non-equilibrium state, and thus the crystalli¬
zation is never completed. This structure may be improved by post-crystalli¬
zation and recrystallization processes, which are collectively designated
secondary crystallization. Since the rate of primary crystallization is mark¬ edly
higher than that of secondary, these two processes are generally consecutive and
well separated. Primary crystallization consists of two elementary processes:
nucleation and growth [6]. The overall rate of crystallization is determined by
these two partial processes.

58 Crystallization, melting and supermolecular structure In the nucleation


process, centres of the new phase are formed. Under isothermal conditions, their
number is either constant (athermal nucleation) or increasing linearly in time
(thermal nucleation). Three types of nuclei can be identified: homogeneous,
heterogeneous and self-nuclei. It is worth noting that during industrial-scale
crystallization of IPP, the addition of hetero¬ geneous (artificial) nucleating
agents and the formation of a characteristic type of self-nuclei, known as row
nuclei (see Section 3.4.2), may play an important role. During crystal growth,
new crystal layers are deposited on the primary nuclei, extending the dimensions
of the crystalline phase. An important quantitative characteristic of growth is the
linear growth rate. Other essential features of growth are the secondary and
tertiary nucleation processes. Secondary nucleation is the rate-controlling step of
growth, and may follow one of three different mechanisms depending on the
extent of supercooling, as described by the kinetic theory of crystallization [98].
These growth regimes are designated I, II and III in order of increasing
supercooling (AT) [98]. A detailed discussion of primary and secondary
nucleation is given in Chapters 4 (Volume 1) and 2 (Volume 2). The
crystallization characteristics of a-IPP under isothermal and non-isothermal
conditions have been inves¬ tigated by many researchers [100-146]. The general
rules of nucleation are covered in [111, 112, 114, 116-119, 126-128, 131, 135-
137, 140-143, 146- 152] and growth kinetics in [9, 15, 28, 59, 68, 102, 103, 117,
125, 152-166]. Studies of crystallization kinetics have involved various IPP
125, 152-166]. Studies of crystallization kinetics have involved various IPP
grades [100-146], their particular fractions [105, 109, 129, 144, 166], low-
molecular weight products [138], polymers of different isotacticities [128, 129,
144, 166], and propylene copolymers [123, 145, 146]. The majority of these
studies refer to the primary crystallization of IPP although some also deal with
secondary crystallization [107, 109, 133]. The crystallization of partially molten
and self-nuclei containing samples is discussed in [70, 104, 115, 124, 132, 134].
The effects of various additives on the crystalliza¬ tion of IPP have also been
investigated. These additives include different nucleating agents [111, 112, 114,
116-119, 142], mineral fillers (e.g. talc [126, 127], CaC03 [135, 140], sepiolite
[136], mica [148]), various pigments [131, 137], carbon black [143], certain
flame retardants [154] and other ingredients [139, 141]. Many recent reports
[139, 150, 156, 158, 162, 167-174] have dealt with crystallization in IPP-based
blends, considering their increasing industrial significance (see Sections 2.3. 2.4
(Volume 2)). These investigations of the crystallization kinetics of a-IPP have
made a significant contribution to polymer physics [6]. Indeed, the theory of
heterogeneous nucleation [151] was substantiated by experiments using a-IPP as
a model. Most of the systematic investigations on artificial nucleating agents
were also associated with IPP [111, 112, 114, 116-119, 126, 127, 135, 137-143,
147-149, 154, 155]. A great number (several

a-Modification of IPP 59 thousands) of materials were also tested for nucleating


activity in the crystallization of IPP, including /?-nucleation [36-71]. Results
from kinetic studies on the growth of a-IPP crystals [9, 28, 59, 68, 100, 103, 117,
125, 138, 153, 155, 156, 159-161, 163, 165, 166] have contributed towards
checking the applicability and the validity of the growth regime theory proposed
by Hoffman, Davies and Lauritzen [98]. IPP is a favourable model substance,
because its linear growth rate (G) can be determined with high precision in a
wide temperature range (Tc) in term of the changes in the radii of spherulites
with time. Under isothermal conditions, G is constant up to high degrees of
conversion [165]. Clark and Hoffman [159] evaluated many data from the
literature [28, 102, 103, 117, 125, 153, 155, 156] relating to the growth kinetics
of a-IPP on the basis of the regime theory. According to this theory, the
experimental G=f(Tc) function is linear within a given regime if lg (G) corrected
by a transport factor is plotted against [TCAT]_1. Break-points between the
linear sections refer to the transition temperatures T(I-II) and T(II-III) from one
regime to the other. Theoretically, the quotient of slopes between two adjacent
regimes 2 (or 0.5). From the slopes, the fold surface free energy (σβ) and the
work of chain folding (q) can be calculated if the necessary material constants
are known. It follows from the literature [59, 68,159, 160] that the experimental
are known. It follows from the literature [59, 68,159, 160] that the experimental
values of G for a-IPP fall in the range of regimes II and III. The transition
temperature T(II-III) is 410 K [59, 159], 411 K [160] or 408 ± 1 K [68]. The
transition temperature T(I-II) as estimated by Monasse and Haudin [160] is 428
± 3 K. The calculated values of σ6 show a large scattering: between 62.5 x 10"3
and 72.2 x 10'3 J/m2 [159], 122 x 10"3 J/m2 [160] and even 230 x 10"3 J/m2
[156]. Overall, reported values values range from 40 x 10"3 [120] to 230 x 10"3
J/m2 [156]. The differences are mainly attributed to the highly diverse material
constants used. In particular, the chosen equilibrium melting point (Tm°) has a
significant influence on σβ [127, 132, 161]. Despite the large numerical
differences reported, it is unequivocal that <re is a multiple of the free energy on
the surfaces parallel to the chains: σ» 11 x 10"3 J/m2 [159, 160]. The value of G
may be decreased by orders of magnitude by the reduc¬ tion of chain regularity,
either through the incorporation of comonomer units [161] or decreasing
isotacticity [166] along with changes in the transition temperatures between the
regimes. In the case of low mole¬ cular mass [144] and less isotactic [166] IPP,
the existence of all three growth regimes has been demonstrated experimentally.
Moreoever, in the latter case, a fourth (intermediate) range was also found. In
addition, it was revealed that the regime transitions I-HI and II->111 took place
at the same supercooling (ΔΤ = 48Κ and ΔΓ = 60Κ), despite the different
transition temperatures (T(I-II) and T(II-III) and different degrees of isotacticity
[166], It thus seems reasonable to suppose that some

60 Crystallization, melting and supermolecular structure morphological changes


may be associated with the growth regime transi¬ tions [160, 161, 163, 166].
Quantitative characteristics of crystallization, as well as the relation between the
molecular structure and the susceptibility to crystallization, are determined by
kinetic studies under isothermal crystallization conditions. For this purpose,
changes in the crystallinity with time, i.e. x=f{t) functions (crystallization
isotherms), are recorded at different Tc. The time course of crystallization was
formerly traced by dilatometry [100-105, 109, 110, 117], optical microscopy
[106, 117, 121] and differential thermal analysis (DTA) [108, 111-116].
Calorimetric measurements have now become predominant [113, 118-124, 126-
130, 132-146, 167-173], though other methods, such as infrared (IR)
spectroscopy [125, 131], are also used. By means of calorimetry, the actual rate
of crystallization can be recorded continuously in time or, under non-isothermal
conditions, in temperature. In some cases, crystallization isotherms are plotted
from the measured melting heats of samples previously crystallized for different
periods of time [144, 166]. Crystallization isotherms are generally processed
using Avrami’s equation [175, 176] or its modified variants [177-179] linearized
using Avrami’s equation [175, 176] or its modified variants [177-179] linearized
as follows: In In ~=n ln(i) + In{K) (1) where x is the relative proportion of
crystallized material, K is the rate constant of crystallization, and n is Avrami’s
exponent (an integer between 1 and 4). Since in polymers x<l, the relative
proportion of crystallized mass corresponds to the ratio of the crystallinity at
time i(xt) to the maximum crystallinity attained during the primary
crystallization (x*,). The interpretation of the kinetic data revealed some
difficulties: e.g. the starting time of crystallization (t0) cannot always be
precisely determined, the initial section of the crystallization isotherm is
recorded inaccurately at low Tc, and the value of x*, cannot be determined
exactly (secondary crystallization). The uncertainty of the initial time of
crystallization is eliminated by introducing an apparent induction period. Its
actual value is calculated upon the minimum of standard deviation [73, 117, 127,
132]. Grenier and Prudhomme [176] analysed the sources of error in the
evaluation of kinetic data due to the uncertainty of tQ, xt and x^. They
established that K and n values are most sensitive to the inaccuracy of xt, while
the effect of x^ is negligible. A positive error of t0 results in a reduction in the
above values. These errors may lead to n values different from integers despite a
satisfactory linear fitting of the crystallization isotherms. K and n can be
determined in other ways, based either on the conversion and time values at the
maximum rate of crystallization, xmax and £max, respectively [120], or on the
maximum rate of crystallization itself and the correspond¬ ing time, imax [65].

61 a-Modification of IPP Crystallization isotherms of a-IPP can be linearized


properly by their processing according to Avrami’s equation [100, 103-105,
113,114,117, 120, 121, 123-125, 131, 132, 135-137, 139, 143, 144, 169, 170,
172], except when residual self-nuclei are present in the molten phase [124, 132,
134] or when pronounced secondary crystallization occurs [109, 129, 144]. From
the dilatometric data, the Avrami’s exponent was n&3 [101-105, 110], while the
calorimetric, optical and spectroscopic results gave n values between 2.4 and 3.6
[119, 121, 123, 124, 127, 129, 131, 132, 135-137, 139, 141, 169, 172]; even n&2
was obtained in some cases [113, 120, 128, 144]. The non-integer values may be
due to the error sources discussed above [176]. The rate constant of
crystallization (K) is in a complex relation to the temperature because of the
temperature-dependence of the elementary processes involved. It was
demonstrated for a-IPP that lg(K) can be linearized as a function of (TCAT)_1 in
a wide temperature range [103, 120, 123, 125, 127-129, 137-139, 172]. This
might be explained upon the athermal nucleation, according to which the value
of n is 3. Therefore, it is very likely that σβ values [113, 120, 123, 125, 127, 132,
138, 139, 163, 172], and even the temperature ranges of Hoffman’s regimes
138, 139, 163, 172], and even the temperature ranges of Hoffman’s regimes
[138, 163], are determined from the temperature dependence of K in some cases.
Attempts [118, 130, 146] have also been made to determine the kinetic
characteristics from non-isothermal crystallization studies [180-183]. Due to the
complex temperature dependence of the rate constant and the peculiar feature of
crystallization, these efforts may result only in qualitative information. In order
to compare the crystallization tendencies of various IPP grades and to
characterize the effects of additives, non-isothermal DTA and calorimetry
studies have been conducted [108, 111-114, 116-119, 127, 145, 146]. Cooling at
a constant rate (vc) leads to crystallization in a range of temperature (ΔTc),
given by the initial (Tc0) and final (Tcf) crystallization temperature. In between,
the temperature to the maximum rate of crystallization can be identified as the
peak temperature of the exothermic crystallization curve (Tcp). By increasing
the rate of cool¬ ing, all these characteristic temperatures are shifted toward
lower values while the range Δ Tc becomes wider [132, 145, 146]. The non-
isothermal investigations provide useful information about the expected
behaviour of a given IPP grade (supercooling, temperature range of
crystallization) under cooling conditions in practice. Additives with nucleating
activity reduce the supercooling, consequently the temperature range of
crystalliza¬ tion is also reduced and the characteristic temperatures are shifted to
higher values (see [111, 127]), while inactive ingredients have no appreciable
influence. Similarly, qualitative information can be obtained about the role of
self-nuclei and row nuclei by means of non-isothermal crystallization

62 Crystallization, melting and supermolecular structure It is worth noting that


the phenomenon of double crystallization was first observed during non-
isothermal crystallization, when two crystallization peaks appeared [114, 116,
124, 132, 136, 140, 148, 167]. This phenomenon may also emerge under
isothermal conditions [124, 132]. Double crystalliza¬ tion was found when
samples containing nucleating additives [114, 116, 136, 140, 148] or previously
melted under ‘mild’ conditions, (thus retaining molecular orientation [124, 132])
were crystallized. It is obvious that double crystallization is the result of two
processes having different mechanisms with considerably diverse temperature
(or time) ranges. In the presence of an additive, the peak at higher temperature
can be attributed to the heterogeneous nucleation while that at lower temperature
referes to the crystallization induced by homogeneous nuclei [114, 116]. On the
other hand, Acosta et al [136] interpreted the double peak as a crystallization
into both spherulitic and mesomorphic structures. The double crystalliza¬ tion of
oriented samples melted under mild thermal conditions (TF^Tm +10 K) revealed
that the peak at higher temperature is caused by self- or row nuclei that
‘survived’ the melting procedure [124, 132]. By elevating the temperature of
‘survived’ the melting procedure [124, 132]. By elevating the temperature of
fusion (TF) or by increasing the number of melting cycles, these self-nuclei are
destroyed. This lead, to a successive reduction in the intensity of the higher
temperature crystallization peak, up to its complete disappearance. Although the
kinetic characteristics of the crystallization of a-IPP have been studied
extensively, many questions about the quantitative description of this process
remain. 3.2.2 Melting characteristics of a-IPP Polymers generally crystallize
under non-equilibrium conditions (Tc^Tm0) with considerable supercooling,
consequently non-equilibrium structures are formed. The further the
crystallization conditions are from the equilibrium (low TC9 high vc)9 the less
perfect is the structure produced (lower degree of crystallinity, smaller size
crystallites, lower melting point) implying a lower structural stability, which is
reflected in the characteristics of the melting process. Generally speaking,
melting of polymers is a complex process; its nature is highly dependent on the
thermal and mechanical conditions during the formation of the crystal structure,
on the thermal history of the crystalline samples, and even on the testing
parameters (e.g. heating rate (uh)) used for melting. The course of melting can
be followed suitably by any methods based on the measurement of properties
related to the decreasing amount of the crystalline phase (e.g. specific volume).
However, the most widespread technique is calorimetry. Calorimetric melting
curves provide valuable and exact information about the thermal (Tm) and
structural (degree of crystallinity) parameters as well as about the complex
character of melting [7].

63 a-Modification of IPP Melting of polymers extends over a wide temperature


range since the size and structure of crystallites vary considerably. The
temperature of fusion of crystallites having the highest melting point is treated
[7] as the experimen¬ tal melting point (Tm). Tm is not a material constant; for
this reason the thermodynamical equilibrium melting point (Tm°) has been
introduced. This corresponds to the melting point of a perfect single crystal of
infinite size. The following relation can be established between Tm and Tm°
[184]: τ =r ° 1 — -*■ m 1 m I i 2σε IAH(' rj-i O rji O C m ~ m IAH{° (2) where
σ6 is the fold surface free energy, l is the lamellae thickness, AH(° is the
equilibrium heat of fusion per unit volume. Tm° is determined by indirect
extrapolation methods [185-187]. The most commonly used is that of Hoffman
and Weeks [185], where Tm is plotted against Tc and the intersection of this
straight line with the diagonal Tc = Tm (representing the equilibrium conditions)
gives Tm°. The majority of Tm° values of a-IPP have been determined by this
method. Tm° can also be determined from / values obtained by X-ray
diffractometry through the Equation 2 from the function Tm=/(/_1) (see [35, 39,
diffractometry through the Equation 2 from the function Tm=/(/_1) (see [35, 39,
188]). Another important material constant is the heat of fusion of a completely
crystalline sample (ΔHt°). If it is known, the degree of crystallinity of different
samples can easily be determined by calorimetry. Literature data [35, 39, 120,
128, 142, 160, 189-214] for Tm° and AH(° of a-IPP are very contradictory.
Reported values of Tm° are: 447 K [192], 449 K [120], 449.6 K [128], 451.4 K
[191], 458 K [200, 205], 459 K [39, 193, 210, 213], 460.5 K [197], 475 K [35],
481 K [160, 199, 205], 482 K [142], 485 K [201], 493 K [194, 197, 200]. Tm°
values of very low molecular mass (2-3000) IPP are much lower: 425.2 K or
427.2 K [138]. Bu, Cheng and Wunderlich [203] analysed the literature data and
found a value of 460.7 K to be most reliable. Low Tm° values are probably
unrealistic since a-IPP samples having Tm = 456K [200] or Tm = 460-461 K
[194], which are far from equilibrium structure, could be prepared by isothermal
crystallization or by long-term annealing. By disturbing the regularity of the
polymer chain (reducing isotacticity or random incorporated comonomer units),
Tm and Tm° are decreased. Con¬ sidering the contradictions in Tm° values,
only the data obtained by identical methods in the same laboratories are worth
comparing. Martuscelli, Pracella and Crispino [129] prepared products of
different isotacticities with various catalysts and obtained Tm° values of 458,
456 and 436 K by the Hoffman-Weeks method. Same researchers [128]
produced ‘high-yield’ IPP with highly active catalysts (isotacticity of 0.975 to
0.90) having Tm° of 464.5, 457.7 and 453.7 K, while they determined a Tm°
value of 449.6 K for a conventional a-IPP sample. Bu, Cheng and Wunderlich
[203] obtained

64 Crystallization, melting and supermolecular structure values between 457.4


and 436.6 K for Tm° for samples of different isotac- ticities (0.998 to 0.787) by
the Hoffman-Weeks method using the lower temperature peak of the melting
curve for extrapolation. Tm° values de¬ rived from the X-ray diffractometric
lamella thickness data by Equation 2 agreed well with the above results. Random
propylene copolymers have lower melting points, lower crystallinity and greater
‘supercoolability’ than IPP [123, 145, 161]. It was demonstrated by Crispino,
Martuscelli and Pracella [123] that Tm° of random propylene/butene copolymers
decreased linearly with increasing concentration of butene-1 from 457.6 K to
392 K at butene-1 content of about 50%. Surprisingly, in spite of the random
structure and high comonomer concentration, the products were capable of
crystallization. Monasse and Haudin [161] found Tm° = 457±5K for
ethylene/propylene copolymer containing 2.8% ethylene and Tm° = 481K for
the homopolymer [160]. They used the higher temperature melting peak for the
extrapolation. The effect of chain irregularities on the depress¬ ion of the
extrapolation. The effect of chain irregularities on the depress¬ ion of the
melting point was interpreted by two extreme models. The defective chain
segments are either included uniformly into or rejected completely from the
lattice. Cheng and colleagues [138, 188] favour a partial inclusion of the defects,
while Monasse and Haudin [161] support complete rejection on the basis of their
experimental data. The chain regularity influences ae values in Equation 2 but
the relevant literature is inconsistent. Cheng and colleagues [138, 188] calculated
that ae was increased by decreasing isotacticity and molecular mass, while
Martuscelli and colleagues [128, 129] and Monasse and Haudin [161] found the
opposite tendency. AHf° values for a-IPP are also widely variable in the
literature and depend on the methods of determination. AH{° calcualted from
solution parameters is generally high: 188 ± 30 J/g [195], 202 ±30 J/g [193], 221
±16 J/g [54], as is that derived from specific heat data: 181.4 J/g [192] or 207 ±5
J/g [198]. By combined calorimetric methods, markedly lower values were
obtained: 137.7 J/g [199] or 138 J/g [196] from melting heat/specific volume
data, 138 J/g [195] or 146.5 J/g [190] from melting heat/crystallinity relations.
Cheng, Janimak and Zhang [188] reported considerably different values within
the range 197.4 to 206.7 J/g. Monasse and Haudin [160] extrapolated the
measured melting heats of samples crystallized at different temperatures to Tm°
and obtained a value of 148 ±10 J/g for AH{°. Bu, Cheng and Wunderlich [203]
found AH{° = 207.1 J/g to be most reliable. The shape of the calorimetric
melting curve, the melting profile, provides valuable information about the
complexity of the melting process as well as about the thermal history, structural
characteristics and stability of the sample. Melting profiles of IPP may be very
diverse. Often two or more peaks appear in the melting curve. Duplication or
multiplication of melting peaks can be attributed to structural reasons and
processes.

65 a-Modification of IPP Peak multiplication due to separated melting of


different crystallites or crystallite fractions Melting processes of crystallites
formed or perfected under different thermal conditions are generally clearly
separated. Melting curves of IPP samples crystallized [70] or annealed [64, 205,
209, 214] according to a stepwise program show individual peaks referring to
separate melting of crystallite fractions. They differ in size (lamella thickness)
and in perfectness but they are identical in their crystalline form. It is remarkable
that this stepwise melting of crystallites formed at different Tc levels can also be
visualised directly by optical microscopy [15, 17, 68, 70]. In samples
crystallized under high mechanical load, extended-chain crystallites of high
melting point may also be formed. For this reason, beside the main melting peak
of folded- chain crystallites, a higher temperature peak or shoulder appears on
of folded- chain crystallites, a higher temperature peak or shoulder appears on
the melting curve. During the crystallization of conventional IPP, more or less ^-
modification may also form. Consequently, its lower temperature small peak is
superimposed on the main melting peak of the a-modification [132, 200].
Melting curves of samples containing both a- and ^-modi¬ fications are,
however, more complex [27, 62]. This problem is detailed in Section 3.3.3. Peak
duplication due to recrystallization superimposed on melting Due to the non-
equilibrium feature of the crystal structure of poly¬ mers, under favourable
thermal conditions (such as during the heating for recording melting curves),
perfection of the structure may proceed within the given crystalline
modification. It is a general feature of poly¬ mers, therefore, that an exothermic
recrystallization is superimposed on the endothermic melting processes. This
leads to a duplication of melting peaks. The higher the deviation of the structure
from the equilibrium (crystallized at low Tc or by rapid cooling), the higher is its
tendency to recrystallization [132, 160, 161, 194]. Recrystallization is a time-
and temperature-dependent process, thus decreasing the heating rate increases
the proportion of a more perfectly formed structure with a higher melting
temperature. Consequently, a reduced heating rate leads to greater ex¬ pressed
peak duplication and to a higher peak at the higher temperature (Figure 3.1). It is
worth mentioning that various nucleating agents reduce the tendency to
recrystallization since they shift the range of crystal¬ lization toward higher
temperatures. Figure 3.2 clearly illustrates that an increasing amount of talc (a
nucleating agent) successively decreases the relative size of the higher
temperature melting peak, referring to the proportion of the recrystallized
material. Consequently, nucleat¬ ing additives, besides other beneficial effects,
increase the structural stability.

66 Crystallization, melting and supermolecular structure Figure 3.1 The effect of


the heating rate (vh) on the melting profile of an a-IPP sample (Pro-Fax H6331)
crystallized at a cooling rate i;c = 20K/min. * It also follows from the above that,
when recrystallization takes place, the melting curve refers to the thermal
characteristics (e.g. Tm) of a partially perfected structure instead of those of the
original sample [194]. It is worth mentioning that these processes take place on
the level of primary fibrils (lamella thickening) without involving the ordered
structure of the elementary cell in general (but see ax to a2 recrystallization). For
the sake of unanimity, the former is designated as αα-recrystallization. This
recrystallization can be eliminated by rapid heating, partial cross-linking or
destruction of chain folds [7]. In order to determine whether a peak duplication
is attributed to a recrystallization or two separated and super¬ imposed melting
processes, the effect of the heating rate on the melting profile (on the relative
intensities of the peaks) could be utilized [7]. For clarifying this question,
intensities of the peaks) could be utilized [7]. For clarifying this question,
annealing [209] or partial heating [61, 64] may also be suitable tests.

a-Modification of IPP 67 Figure 3.2 Melting curves of talc-filled a-IPP samples


(Pro-Fax H6331) crystallized at a cooling rate yc = 20K/min (uh = 2.5 K/min).
Under non-isothermal conditions, at considerable supercooling, the
crystallization of a-IPP leads to samples with a strongly expressed tendency to
recrystallization. Interpretation of melting curves of iso- thermally crystallized a-
IPP is much more complicated and controversial [200]. According to
experimental observations [160, 194, 200, 204-213], the feature of melting
curves of a-IPP as functions of Tc is characterized as follows: 1. Melting curves
of samples crystallized at very low Tc (<385 K) show doubled α-peaks due to
the tendency to recrystallization caused by the high structural instability. In this
range, Tm is independent of Tc since these Tm values refer to the recrystallized
products instead of to the original ones [160, 194]. 2. Melting curves after
crystallization at medium Tc levels (~385 to 405 K) comprise a single relatively
sharp peak. Tm increases with increasing Tc and this α-peak becomes even
sharper. In accord with the above, all these observations prove that elevating Tc
leads to more and more stable

68 Crystallization, melting and supermolecular structure structures that are not


susceptible to recrystallization at a given vh value (5 to 20 K/min). 3. Melting
curves of samples crystallized at high Tc (405 to 425 K) show peak duplication
[160, 200, 204-213] but Tm still increases with Tc. Samuels [200] could not find
a satisfactory interpretation for this. Kamide and Yamaguchi [204] attributed this
peak duplication to the formation of crystallites of higher melting point, which
had been perfec¬ ted by the annealing effect during the isothermal crystallization
(lamella thickening). Yadav and Jain [205], on the other hand, derived the peak
duplication from the fractional crystallization of molecules with different
stereoregularities in this range of Tc. Both interpretations are based on the
separated melting of two different crystallite fractions. The Italian research
group [206-213] explained this peak duplication as a new form of
recrystallization and excluded the possibility of a superposition of two separated
melting processes. They considered that the recrystallization proceeds between
two α-modifications having differ¬ ent crystal order (ο^ and a2). Their structures
were determined by Hikosaka and Seto [215]. The recrystallization was proved
by structural studies and by the effect of heating rate on the relative intensity of
melting peaks [206, 210]. It was also established [211] that the rate of a1 to a2
recrystallization was markedly higher than that of the rearrange¬ ment processes
in samples crystallized at lower Tc leading to lamella thickening. 4. Melting
in samples crystallized at lower Tc leading to lamella thickening. 4. Melting
curves of samples crystallized at very high Tc levels (> 425 K) comprise a single
melting peak again, even at low vh 2.5 K/min [210] due to the increasingly
stable structure and to the decreasing rate of recrys¬ tallization. Melting
characteristics of a-IPP, therefore, have several specific features. The
controversial literature data for Tm° of a-IPP can be explained partly by this.
With the Hoffman-Weeks extrapolation [185] Tm° values of the lower and the
higher temperature peaks are around 459 K and 499 K, respectively [160, 194,
200, 204-213]. The latter values seem to be unreliable after the results of Italian
researchers [210, 213], since they are based on the melting parameters of the
recrystallized proportion of the material. 3.3 jS-MODIFICATION OF IPP 3.3.1
Preparation of /f-IPP Commercial grades of IPP, when processed in laboratory or
industrial scale, crystallize essentially into the α-modification. At first Padden
and colleagues [9, 10] demonstrated by X-ray diffractometric [10] and optical
microscopic [9] studies that, at high supercooling, the formation of the a-
modification

β-Modification of IPP 69 was accompanied by the appearance of the /^-


modification. Soon after, this observation was supported by other researchers
[11, 12]. Crystal structure of the /5-phase was determined in conventional IPP
samples with low /^-content [10-14]. Structure [9, 13-21], melting [9, 15, 17]
and growth [9, 15-18] of /?-spherulites were studied and some of their
characteristics were described. The β-modification was also investigated by IR
[22, 23] and by nuclear magnetic resonance (NMR) [24] spectroscopy. Turner
Jones, Aizlewood and Beckett [12] were the first to report the formation of
samples highly rich in /^-modification. They crystallized a particular
(unspecified) IPP sample in the range of 393-403 K into nearly pure /5-IPP. The
authors did not clearly understand what this high suscepti¬ bility of their sample
to /5-crystallization was derived from. They intro¬ duced a k value for
characterizing the proportion of ^-modification by X-ray diffractometric data, as
follows: Ηβί Ηβ1+{Ηαι+Ηα2 + Ηα3) (3) where Hal, Ha2i Ha3 are intensities of
α-diffraction peaks corresponding to angles 0 = 7,1°, 8.5° and 9.4°, respectively,
and Ηβ1 is the intensity of /5-diffraction peak corresponding to angle 0 = 8.1°.
The k value is a relative measure for characterizing the polymorphic
composition but it does not express the absolute value of /?-content numerically,
although its value is 0 for a-IPP and 1 for /5-IPP, as is apparent. A k value of
^0.85 was found for the samples prepared by Turner Jones, Aizlewood and
Beckett [12]. It is notable that polarized light microscopy (PLM) is also capable
of relative quantification of the ^-modification since the peculiar high
birefringence and type of /5-spherulites are different from those of the a-
birefringence and type of /5-spherulites are different from those of the a-
modification [9]. Calorimetric melting curves also provide information about the
polymor¬ phic composition. However, melting of the ^-modification is a very
complex process including pronounced melting memory effect (see Section
3.3.3). In fact, calorimetric curves give no exact information on the polymorphic
composition unless the possibility of /?a-recrystallization is eliminated. This
elimination can be maintained if samples containing /5-IPP are never cooled
below the critical temperature of T%~ 383 K. It should be noted that all the
calorimetric curves and thermo-optical micrographs in this chapter were
recorded under these conditions. Several attempts were subsequently made to
prepare the ^-modification in pure form. Two methods have been used
successfully to prepare samples rich in the ^-modification: the temperature
gradient method [25-35] and the introduction of /?-nucleating agents [36-75].
Furthermore, it is known that shear stress also promotes the formation of the /5-
modification [76-79]. However, as is reflected in the literature, pure /5-IPP failed
to form in most cases.

70 Crystallization, melting and supermolecular structure Under laboratory and


technological conditions, pure /MPP was prepared by Varga and colleagues [61-
67] from specially pigmented and stabilized IPP samples using particular
thermal conditions for the crystallization. Formation of pure jS-phase was
proved unequivocally [61-65, 67] by X-ray diffractometry (Figure 3.3), and
morphological studies were conducted by optical microscopy (Figure 3.4) and
calorimetry (Figure 3.5). More compre¬ hensive investigations revealed that the
formation of /MPP had an upper and a lower (threshold) temperature, Τ(βa) and
T(a/?), respectively [67-69], which is discussed below, with reference to the
experimental data. Methods for the preparation of /?-IPP rich samples or even
pure /MPP are described in the following sections. Temperature gradient method
Crissman [25], Kamida and Nakamura [26], Fujiwara [27] and Lovinger, Chua
and Gryte [28] prepared small oriented /MPP films by tempera¬ ture gradient
crystallization. The method was reviewed in detail by Fujiwara in his monograph
[29]. The procedure is based on the appear¬ ance of jS-nuclei on the surface of
growing α-crystal fronts propagating into jS-spherulite segments (α/Mransition).
Due to the higher growth rate of the /?-phase [68], it spreads over the total bulk
of melt. Although this method principally permits the formation of pure /?-
phase, a-nuclei may accidentally emerge in the melt region, growing into a-
spherulites situated in the continuous jS-phase as teardrop-shaped inclusions.
Film samples obtained in this way were used for determination of several
peculiarities and characteristics of the structure and behaviour of the jS-phase
[25-35].
β-Modification of IPP 71 Figure 3.4 Optical micrograph of pure β-ΙΡΡ
crystallized at Tc = 398 K [73]. Addition of selective β-nucleating agents
Crystallization in the presence of /^-nucleating agents promises to be the most
reliable method for industrial applications. The most well known and widespread
β-nucleating agent is a y-quinacridone red pigment (Permanent Rot E3B), which
was first suggested by Leugering [36] as an efficient agent. It was later used by
several research groups [37-49]. IPP samples crystallized in the presence of y-
quinacridone contain large amounts of the ^-modification. Their k values are
around 0.8 as calculated from X-ray diffractometric data [40,44] but, when
crystallized under given thermal conditions, k may reach about 0.9 [41,42]. Even
though y- quinacridone is one of the most effective /^-nucleating agent, pure /?-
IPP could not be prepared in its presence. In other words, its α-nucleating effect
is not negligible. It was also observed [42] that its ^-nucleating activity is
reduced during the processing. This indicates a modification in the physical or
chemical structure of this red pigment due to the high pressure and temperature
of processing [42]. A drawback of y-quinacridone is its intense red colour, which
may restrict its practical application. Garbarczik and Paukszta [45] found that
heterocyclic stabilizers, such as triphenol ditriazine, were effective ^-nucleating
agents (fc^0.3). According to the literature, aluminium salt of quinizarin
sulphonic acid [50], disodium [51] and calcium salts of phthalic acid [52] have
jS-nucleating activity. The activity of disodium phthalate has not been confirmed
by the present author’s experiments.

72 Crystallization, melting and supermolecular structure Figure 3.5 Melting


curves of ^-nucleated IPP samples crystallized at different temperatures plotted
on a common baseline (uh= lOK/min) [67]. Shi, Zhang and Jins [53] revealed a
very high β-nucleating activity for some two-component compounds obtained by
the reactions of certain organic acids with CaC03, but not defined unequivocally
in their patent disclosure. This research group devoted a number of reports [54-
59] to investigations of their products of predominantly β-modification
crystallized by means of their patented nucleating agents. They report very high
k values, even as high as 0.93. It has recently been observed that the mineral
wollastonite (calcium silicate) is also active as a ^-nucleating additive [60]. We
studied [71] the /J-nucleating activities of Ca, Ba, Zn, Cd and A1 salts of
aliphatic and aromatic dicarboxylic and polycarboxylic acids in our laboratories.
It was established that several Ca and Zn derivatives were very active while
some Ba salts were moderately active ^-nucleating agents. No Cd and A1
derivatives were found to be active. Figures 3.6 and 3.7 show calorimetric
melting curves for samples crystallized in the presence of some ^-nucleating
additives. It can be seen that Ca salts of pimelic, phthalic and
additives. It can be seen that Ca salts of pimelic, phthalic and

Figure 3.6 Melting curves of IPP samples (Tipplen H523) nucleated with ,Ca
salts (1 w/w %) of different organic acids (vc = 10 K/min, uh=10K/min, TR =
383 K: (1) original IPP; (2) benzoic; (3) phthalic; (4) terephthalic; (5)
pyromellitic acid. pyromellitic acids have the highest nucleating activities, as
proved by the presence of intensive /?-peaks in the melting curves. It can be
supposed upon these results that there may be a large number of ^-nucleating
agents (including selective ones), and that the discovery and introduction of
further ones can be predicted. This supposition was supported by a
comprehensive calorimetric essay on commercial IPP grades, some of which
showed a high tendency to β- crystallization.

74 Crystallization, melting and supermolecular structure 120 130 m 150 <60 1T0
180T[aCj \ ^ \ β ot i \ ' I 1 I I 1 1 1 390 MOO M10 M20 M30 MMO M50 T[k]
Figure 3.7 Melting curves of IPP samples (Tipplen H523) nucleated with
different salts of pimelic acid (Tc = 397 K, vh = 10 K/min): (1) Ca; (2) Zn; (3)
Al. Crystallization in mechanically loaded melts It was observed in many cases
that melts exposed to shear (or tensile) stress crystallized into jS-phase at a
higher proportion than the average [76-79]. Although this method is not suitable
for the preparation of pure /MPP, it is capable of enhancing the local
concentration of /?-phase. The characteristics of shear-induced crystallization of
/J-IPP are surveyed in Section 3.4.2. 3.3.2 Crystallization of /MPP Kinetic
studies of crystallization of /MPP were opened up by the emergence of /J-
nucleating agents [43, 57, 59, 63, 65]. In their presence, the product

75 β-Modification of IPP formed consists mostly of the /^-modification.


However, the simultaneous formation of α-phase and its interference was
avoided in only two cases [63,65]. The elementary steps of crystallization
(characteristics of nu- cleation and growth) are covered by many papers.
Nucleations The fundamental empirical observations with ^-nucleating agents
[36-75] were surveyed in the previous section. Evaluation of these data is
hindered by a lack of informative facts and data about the exact compositions of
the nucleating agents [50-53, 58, 67, 71], since the relevant data might be
utilized by industry. Interpretation of the selective /^-nucleating activity requires
more detailed knowledge. However, it can be established from the experimental
results to date that the effect of jS-nucleating agents is consider¬ ably suppressed
when α-nucleating agents such as talc [67,74] and self¬ nuclei [66] remaining
‘intact’ after melting are present simultaneously. When this occurs, there is
competition between the two types of nucleation, a and β. It is a basic aspect that
competition between the two types of nucleation, a and β. It is a basic aspect that
α-nucleating agents are effective in the whole temperature range below Tm°(a),
while the ^-nucleating agents are active only below Τ(βοί). Kinetics of growth
Some characteristics of kinetics of growth were discovered for jS-IPP rather
early [9,15,17,28] because these features can be studied in conventional IPP
samples with low ^-modification content. More recent investigations
[59,65,68,70] have been conducted with pure jS-IPP or with samples of high /?-
phase content. Padden and Keith [9] first pointed out that the growth rate of jS-
spherulites was higher than that of α-spherulites. Lovinger, Chua and Gryte [28]
demonstrated that the ratio of growth rates of the two modifications (Gp/Ga),
while decreasing with elevating 7^, was always higher than 1 in the temperature
range studied. Using the method of stepwise isothermal crystallization, the
temperature range for the determination of Gp was considerably extended by
Varga [15], reaching a critical temperature, denoted as T(/?a), above which
Ga>Gp. During the stepwise crystallization, /?-spherulites are first prepared at
lower Tc and then, after heating to a higher temperature, the rate of further
growth is measured. It was established by more detailed investigations [68] that:
1. The ratio Gp/Ga decreases steadily with increasing Tc and, above Τ(βα) =
413-414 K, Ga>Gp (Figure 3.8). 2. At a temperature of the second step above
Τ(βa), punctiform a-nuclei are formed on the surface of /?-spherulites, growing
into a-spherulite segments which finally encompass the basic /?-spherulite. This

76 Crystallization, melting and supermolecular structure Figure 3.8 Temperature


dependence of linear growth rate (Ga and Gp) of a- and /^-modifications of an
IPP homopolymer [68, 70]. phenomenon wat· named /^-bifurcation of growth
and interpreted on the basis of the secondary /?a-nucleation [68]. A further
increase in Tc in the range of TC>T(βα) enhances the probability of /fa-
transitions (while Ga and Gp decrease steadily) becoming predominant as
approaching Tm(/?). The /^-bifurcation of growth was documented by optical
microscopy in our earlier works [68,70]. It is worth mentioning that Shi, Zhang
and Qui [59] again proved recently that Ga>Gp above 413 K, but did not
mention the j8a-transition. 3. The critical temperature, Τ(βa), can be accepted as
an upper limit temperature of formation of β-ΙΡΡ since above this temperature
forma¬ tion of jS-IPP is not possible. The formation of /MPP, however, also has
a lower limit temperature, Τ{αβ), as demonstrated by Lotz et al [69]. They
revealed that partially crystallized samples were subject to a /ta-transition on the
surface of growing /?-spherulites in the course of cooling at about 378 K. This
transition is similar to that observed above Τ(βa) [68] but is difficult to recognize
because of the rapid crystalliza¬ tion and the small spherulite sizes. The
existence of the lower limit
β-Modification of IPP 77 temperature had been suggested earlier [62,67] on the
basis of the melt¬ ing memory effect. 4. It was established by evaluating the
kinetic data [59,65,68] according to the kinetic theory of crystallization [98, 159]
that the growth of jS-IPP in the temperature range studied proceeds in the
Hoffman’s regimes II and III. The transition temperature, 7"(II—III), is about
406 K, according Varga and colleagues [65,68], while Shi, Zhang and Qui [59]
obtained 406-410 K for various samples. The ratio of slopes of the straight lines
for the two regimes, m(III)/m(II), was lower than the theoretical value of 2,
varying at around 1.5 [59,65,68]. For this reason, <re values calculated from the
two regimes are different from each other. Varga and Garzo [65] obtained
σ6(ΙΙ)= 112 x 10”3 J/m2 and σβ(ΙΙΙ) = 85.4 x 10“3 J/m2, while Shi, Zhang and
Qui [59] found values between 48.2 x 10“3 and 67 x 10“3 J/m2 for <re(III). It is
worth comparing the above values with data determined by independent
methods. By processing functions through Equation 2, Ulmann and Wendorff
[39] obtained σβ = 35 x 10“3 J/m2, while the result of Fujiwara [35], σβ = 77 x
10“3 J/m2, was close to those calculated from the kinetics of growth. The
differences in σβ can be explained partly by the fact that the various researchers
did not use identical material constants for the calculations. Despite the
differences in the numerical values of <re, it can be concluded that σβ values of
the /^-modifications are always lower than those of the a-modifications
[35,39,59,68]. The tendency of various IPP grades to undergo ^-crystallization
can be predicted by comparison of the related Ga and values. Our results on
different RC-PP grades showed that the tendency of random ethy-
lene/propylene copolymers with low (1.8-2.5%) ethylene content (such as
Moplen EP1X35F) to /^-crystallization is negligible. On the other hand, the
incorporation of an efficient ^-nucleating agent might force their crystalliza¬
tion into the /^-modification. According to the kinetic measurements of
isothermal growth, this is attributed to the fact that the growth rate of the α-phase
is always higher than that of the /?-phase (Figure 3.9). In addition, a spontaneous
/?a-bifurcation of growth may appear during the isothermal crystallization. The
onset of /fa-bifurcation of growth increases with Tc to such a degree that above
403 K determination of Gp is not possible. It can be concluded, therefore, that a
disturbance in the regularity of the IPP chain results in a reduced tendency to jS-
crystallization. Overall crystallization Crystallization of samples containing ^-
nucleating agents was investigated under non-isothermal [39, 41-43, 60, 63, 64]
and isothermal [12, 57, 65]

78 Crystallization, melting and supermolecular structure 20 CO 15 10 5 0 395


WO 405 TC[K] 410 Figure 3.9 Temperature dependence of linear growth rate
(G) of a- and ^-modifi¬ cations for a random propylene copolymer (Moplen
(G) of a- and ^-modifi¬ cations for a random propylene copolymer (Moplen
EP1X35F). conditions. These studies were aimed principally at the
determination of the optimum thermal conditions for the formation of /J-IPP.
Turner Jones, Aizlewood and Beckett [12] reported that the highest amount of
/J-IPP was formed between 373 and 403 K. Other authors [39, 41-43] indicated
wider ranges, from room temperature to 400 K. Based on literature data and
experimental data of his own, the present author takes the following view. The
formation of /J-IPP has a theoretical upper and lower limit tempera¬ ture, as
stated above [62, 67, 69]. Within these limits, a narrower range may be allocated
where pure /J-IPP is produced. This range is bordered by an upper (T*) and
lower (T**) critical temperature of crystallization. Whether or not this
temperature range (T*-T**) exists depends on the type of IPP, and on the
selectivity and concentration of β-nucleating agents [65, 67]. Correspondingly,
pure jS-IPP is produced if Tc or the temperature range of a non-isothermal
crystallization is between Τ*(=Τ(β<χ)) and T**(=T(a/?)) [63, 65]. Thus, for a
given system of IPP with ^-nucleating agents, an optimum range of cooling rates
may exist when the above thermal expecta¬ tion is fulfilled. On the other hand,
when ^-nucleating agents that are not completely selective (e.g. y-quinacridone)
are used, it is advisable to cool rapidly into the range between T(β<χ) and Τ(<χβ)
in order to avoid the onset of crystallization above Τ(βa).

79 β-Modification of IPP Very few reports deal with the isothermal


crystallization of /MPP. Moos [42] and Moos and Jungnickel [43] attempted to
determine the kinetic characteristics of crystallization of /MPP. They processed
the crystallization curves recorded at constant cooling rate through the Avrami’s
equation for y-quinacridone containing samples. From the linearized data, the
Avrami’s exponent was approximately n = 3 at low cooling rates, suggesting
athermic nucleation. However, due to the non-isothermal external conditions and
to the simultaneous formation of a- and /^-modifications, both the qualitative
and the quantitative conclusions derived from the experimental data should be
considered with criticism. Shi et al [57] revealed that crystallization isotherms of
samples crystallized predominantly in the /?-form could not be linearized
through the Avrami’s equation. This was explained by the simultaneous
formation of a- and ^-modifications. Isothermal crystallization into pure /MPP
was studied by Varga and Garzo [65] in the temperature range between T* and
T**. They estab¬ lished that the crystallization isotherms could be linearized by
the Avrami’s equation and that the Avrami’s exponent was around 3. The rate
constant of crystallization was also linearized in the lg(K) = /(Γ0ΔΓ)_1 form.
The type of nucleation was determined by maching K and Gp values. The
nucleation was proved to be athermal but the calculated density of nuclei
nucleation was proved to be athermal but the calculated density of nuclei
decreases with increasing Tc. Reliability of the experimental conclusions is
reduced by the fact that the investigations covered a relatively narrow range of
Tc. 3.3.3 Melting characteristics and tendency to recrystallization of /MPP
Padden and Keith [9] observed melting of spherulites of ^-modification by PLM
and demonstrated that the spherulites recrystallize into the «-modifi¬ cation (jSa-
recrystallization) when heated and, finally, melt from this form. This observation
was later confirmed by several authors [26-28, 33, 39, 40, 44, 47-49, 51, 55-59]
and used to support the supposed thermodynamical instability of the /^-
modification. However, it was later discovered that /MPP is a unique polymer
with respect to its melting characteristics. Varga [62] revealed that the melting
characteristics of /MPP were highly depen¬ dent on the thermal post-history of
the crystalline sample. In other words, an expressed memory effect of /MPP was
detected. It was stated that /MPP samples are susceptible to /fa-recrystallization
only if they were cooled below a critical temperature before melting. This
critical ‘recooling tempera¬ ture’ was T$ = 373-383 K [61-63, 67]. On the other
hand, /MPP samples that were not cooled below do not recrystallize into the a-
modification when heated; instead, they melt separately like the
thermodynamically stable modifications. Consequently, the tendency to )Sa-
recrystallization is not a general feature of /MPP; it is associated only with a
given thermal history, namely, cooling below .

80 Crystallization, melting and supermolecular structure The melting memory


effect is illustrated by the calorimetric curves (Figure 3.10) for a pure and a low
α-containing /MPP. In order to focus on the time-dependent recrystallization
process, the melting curves were recorded at an unusually low heating rate (uh =
2.5 K/min). Despite this low vh9 even traces of /?a-recrystallization could not be
detected when the sample had not been recooled. In contrast, the endothermic
melt¬ ing process of recooled /MPP samples was superimposed by an exo¬
thermic /?<x-recrystallization (βα-RC). It can also be seen (Figure 3.10) that the
new α-phase (a2) formed by the /fa-recrystallization has a higher melting point
than the original α-content of the sample (at), as explained by its higher
temperature of formation (recrystallization temperature). The melting memory
effect and the tendency to )Sa-recrystallization of /MPP cooled below can be
attributed to the formation of a very low Figure 3.10 Melting curves of pure
/MPP and chalk-filled (2.5 w/w %) predomi¬ nantly /MPP samples crystallized
at TC = 401K (vh = 2.5 K/min): (1) non-recooled pure /MPP (TR=TC); (2)
recooled pure /MPP (TR = TRT); (3) non-recooled chalk- filled /MPP (TR =
TC); (4) recooled chalk-filled 0-IPP (TR = TRT).
81 I^-Modification of IPP amount of α-phase within the β-spherulites due to a
secondary crystall¬ ization in the cooling process below T$. This α-phase acts as
an a- nucleating agent during the partial melting of the /?-phase and induces a
jSa-recrystallization, meeting both preconditions of the /?a-transition, namely,
the mobility of molecules and the presence of nuclei of the new phase. It is also
apparent from the above interpretation of the memory effect that the critical
recooling temperature corresponds to the lower limit temperature, (Ταβ), of the
formation of /MPP [62, 67]. It should be emphasized, however, that α-
spherulites in mixed polymorphic samples (Figure 3.10) and the α-phase formed
by /?a-bifurcation of growth [62, 70] will not induce /?a-recrystallization. This
effect is exerted only by an α-phase, assumed to be very finely distributed,
formed within the /?-phase below T£. In addition to the scientific interest, the
melting memory effect involves an important methodical consequence: the
calorimetric melting curves of non- recooled samples permit an easy and exact
determination of the polymor¬ phic composition (^-proportion). This may result
in the replacement of expensive and laborious X-ray diffractometry by simple
calorimetric measurements. Calorimetry provides a simple monitoring of the
composi¬ tion-dependence of polymorphic IPP during crystallization [61-63] as
a function of the thermal profile applied [64], and of the type and amount of
additives - nucleating agents [71], fillers [74] or other polymeric additives [67,
74, 75]. In such a way, information can be obtained about the relative activities
of some jS-nucleating agents (see Figures 3.6 and 3.7). The tendency to /?a-
recrystallization and the presence of a greater or lesser amount of α-phase (i.e. a
mixed polymorphic composition) disturb the experimental determination of two
crucial thermodynamical parameters of /MPP, Tm° and AH{°. This is the reason
for the wide variation in the literature data determined before the recognition of
or without consider¬ ation of the melting memory effect. For Tm°, values of 420
K [26], 443 K [200], 450 K[39], 456-457 K [35, 62, 65], 465 K [54] and 473 K
[28] have been given. The author accepts 457 ± 4 K [65] as most reliable since it
was determined using pure /MPP and eliminating any disturbance of /fa-
recrystallization. For AHf°, values of 100 J/g [26], 113 J/g [65] and 193 J/g [54]
were found. The value of 113 + 11 J/g was determined for pure /MPP by the
method of Monasse and Haudin [160]. Just as for a-IPP, this method gives lower
values than the procedure based on the depression of melting point [54]. Shi,
Huang and Zhang [54] used mixed polymorphic samples, which impaired the
reliability of their values. The characteristics of the recrystallization of /MPP are
worth surveying more comprehensively. It has been pointed out in connection
with the melting of a-IPP that an αα-recrystallization may take place within the
α-phase as a structural perfection and stabilization during the heating of this
82 Crystallization, melting and supermolecular structure non-equilibrium
structure. The scenario for /MPP is more complicated because two kinds of
recrystallization should be distinguished: 1. /?/?-recrystallization (ββ-RC) within
the ^-modification leading to a stabilization (perfection) of the structure of the
/^-modification. 2. /fa-recrystallization (/fa-RC) involving a β to a phase
transition. In the course of heating of non-recooled /J-IPP samples, they behave
as thermodynamically stable modifications. In their melting process, a ββ-
recrystallization takes place (i.e. without transition in the crystalline form),
resulting in a duplication of the melting peak of jS-phase (Figures 3.6 and 3.11).
The tendency of /?/?-recrystallization is stronger at higher struc¬ tural
instability, i.e. when the sample was crystallized at higher supercooling Figure
3.11 Effect of recooling temperature (TR) on the melting curves of /1-IPP
samples crystallized at a cooling rate yc=10K/min. The melting curves were
registered at a low heating rate (t>h = 2.5 K/min) [63].

83 β-Modification of IPP (low Tc, high vc) within the range between Τ(<χβ) and
Τ(β<χ). A reduction in the heating rate is also favourable for /?/?-
recrystallization [61-63]. For determination of the thermal characteristics of /?-
IPP, the disturbance of /?/?-recrystallization should be eliminated by using high
heating rates (i;h^10K/min). /fa-recrystallization is a characteristic feature of /J-
IPP cooled below . The tendency to /fa-recrystallization is influenced by the
aforementioned parameters (Tc, vC9 vh) according to the regularities of
polymers [61-63]. The tendency to /fa-recrystallization is supposed to be
induced by the α-phase formed during the secondary crystallization below T£.
The amount of the α-phase increases with decreasing TR. For this reason, the
/fo¬ recrystallization is influenced by the TR level (i.e. the amount of a-phase)
besides the above parameters (Tc, vc, vh). As the factors controlling the
recrystallization are changed, the tempera¬ ture range of /fa- and /?/?-
recrystallization varies, while intermittently one or the other process becomes
predominent, as illustrated by the experimental results shown in Figure 3.11. The
sample recooled to a low temperature (TR = 323 K) has a high tendency to /fa-
recrystallization and the /fa-transi- tion proceeds at an early stage of the partial
melting of the /^-modification. Consequently, the /?/?-recrystallization becomes
impossible. At low recooling (TR> T$), no α-phase is formed; thus, only /?/?-
recrystallization can occur. At an intermediate recooling temperature (Figure
3.11; TR = 363 K), both types of recrystallization appear, clearly separated from
each another. In this case, the temperature range of /fa-recrystallization exceeds
that of the /?/?-recrys- tallization (due to the relatively lower amount of α-phase)
and the transition is preceded by the perfection within the j8-phase. The
assumption that the α-phase produced by recooling is responsible for the
assumption that the α-phase produced by recooling is responsible for the
enhanced tendency to /fa-recrystallization is confirmed by the data in Figure
3.11. Varga and Toth [84] conducted comprehensive annealing studies in order
to decide whether the tendency of β-ΙΡΡ to /fa-recrystallization caused by
recooling can be stopped and how the structural stability of jS-phase can be
enhanced. They demonstrated that the feature of structural changes induced by
annealing is highly dependent on the thermal history of the samples. In other
words, a strong ‘annealing memory effect’ was observed. When annealing is
started from the temperature of crystallization, β-ΙΡΡ behaves like a
thermodynamically stable modification, i.e. transforms into a more stable form
with higher melting point as a result of the structural perfection processes (ββ-
RC). After annealing, the tendency to /fa-re- crystallization practically
disappears. It is remarkable that, at annealing temperatures above Tm(β) (but
Tm(/?)<Ta <Tm(a)), when /J-IPP has com¬ pletely melted, on cooling the
samples recrystallize in the a-modification [66]. This recrystallization into the α-
modification is a sensitive indicator to the entire melting of /?-IPP. With a step-
wise decreasing temperature programme for annealing, jS-IPP can be resolved
into several stabilized

84 Crystallization, melting and supermolecular structure β-crystallite fractions


with different melting points whereby the tendency to jSa-recrystallization due
to the recooling is eliminated. It is notable that the separate melting of ^-
crystallite fractions can be attributed to the different lamella thicknesses (cf.
Equation 2) rather than to the diverse geometry of the elementary cell supposed
by Garbarczyk [48]. On the other hand, /J-IPP cooled before annealing below
transforms, partly or completely, into the α-modification (/?a-RC) since, in this
case, the nucleating effect of the α-phase formed during the recooling is exerted.
It can be established from the foregoing that melting and recrystallization of /?-
IPP is an extremely complex process influenced by a number of factors, most
essentially by the conditions of crystallization and fusion and the post-
crystallization thermal history, including not only the actual values of recooling
and annealing temperatures but also their sequence. All these factors can be
summarized as the ‘overall thermal history’. It results in very diverse melting
curves even for pure j?-IPP [61-67]. The characteristics of melting curves for jS-
IPP reported in the literature are complicated further by the fact that they were
recorded for mixed polymorphic samples [27, 28, 39, 49, 54-56, 60]. For this
reason, conclusions from the calorimetric measurements should be treated
carefully and critically in most cases. The contradictory conclusions and
observations may also be attributed to the various ‘overall thermal histories’ of
the samples investigated [27, 39, 40, 48, 49, 55, 56]. The melting characteristics
the samples investigated [27, 39, 40, 48, 49, 55, 56]. The melting characteristics
of /?-IPP and the phenomenon of the melting memory effect are summarized in
Figure 3.12. Limit temperatures of formation of pure /J-IPP (T*, Γ**), the
critical recooling temperature (T&), and the melting processes of the individual
modifications (M(α), Μ(β), Μ(β')) in the corresponding melting ranges, Tm(a)
and Tm(/?) are indicated in the figure. The amorphous isotropic molten state
(Α{) is distinguished from the amorphous states formed immediately after
melting (Αα,Αβ) since they comprise partial ordering and residual self-nuclei
from the original crystal structure up to a certain temperature, (T£>Tm). During
melting of non-recooled jS-IPP (C^) crystallized isothermally or non-
isothermally between T* and T**, no modification transition occurs (1-3-9 or 1-
2-3-9 in Figure 3.12). The phase Οβ either melts immediately (1-3-9) or is
superimposed by a structural perfection within the j?-phase (/?/?-RC), resulting
in a phase Οβ with a higher melting point (1-2-3-9). During melting of /Ϊ-ΙΡΡ
cooled below T£, a /fa-modification transition takes place in any case (1-4-6-7-
8-9). The /?a-recrystallization during the partial melting of jS-IPP (6-7) is
induced by the α-phase (Na) formed by the secondary crystallization below Tj£.
According to our experiments, /fa- recrystallization is the predominant process
in samples cooled below Tj£ although, under certain conditions (cf. the curve
corresponding to the recooling temperature TR = 363K in Figure 3.11), it may
be preceded by a ^-crystallization (1-4-5-6-7). During melting of a sample
transformed into

Supermolecular structure of IPP 85 Figure 3.12 The scheme of melting feature


and melting memory effect of /MPP (see text). the α-modification (Ca), a
partially ordered amorphous phase (Aa) is formed first ‘resembling’ the original
structure of the α-modification, followed by a transition into the isotropic
amorphous state (.A{) above TjL 3.4 SUPERMOLECULAR STRUCTURE OF
IPP Aggregates of folded-chain primary crystallites (fibrils or lamellae) ar¬
ranged in definite geometry are designated supermolecular structures. The most
frequent supermolecular formations in melt crystallized polymer are spherulites
but, under special conditions, cylindrites, axialites, quadrites, hedrites and
dendrites may also be formed [5, 8, 70, 216, 217]. The formation and structural
features of various types of supermolecular structures are influenced essentially
by the thermal conditions of crystal¬ lization, by the mechanical stresses to the
melt and by the foreign materials present. In a quiescent melt, crystallization is
generally spheru- litic while under mechanical load, cylindritic structures are
formed. In the presence of extraneous materials (heterogeneous surfaces),
characteristic transcrystalline structures may formed. In the following sections,
character¬ istics of the spherulitic, cylindritic and transcrystalline structures of
character¬ istics of the spherulitic, cylindritic and transcrystalline structures of
IPP are discussed. Since several reviews [70, 216-218] are available for this
topic, only the general aspects of the supermolecular structures are covered.

86 Crystallization, melting and supermolecular structure 3.4.1 Spherulitic


structure A spherulite is a spheriform cluster of primary crystallites with
spherical symmetry. Spherulites develop from crystallites starting from a central
nucleus and growing uniformly in all spatial directions radially, with small-angle
(non-crystallographic) branching in between. The branching of growing
crystallites provides complete space filling. The phenomenological theory of
spherulitic crystallization of polymers was substantiated by Keith and Padden
[219] and subsequently modified [220] on the basis of criticism by Bassett and
Vaughan [221]. According to the phenomenological theory, the small-angle
fibrillar branching essential in the formation of spherulitic structures is induced
by impurities (with low, if any, tendency to crystalliza¬ tion) segregating in the
vicinity of the growing crystal front. It can be observed when a crystallization in
a thin film of melt (essentially a two-dimensional crystallization) is monitored
by PLM that, on randomly located thermal or athermal nuclei, disk-like
birefringent spherulites are produced, growing radially at a constant rate (linear
growth rate) under isothermal conditions. Due to the impingement of growing
spherulitic fronts, the structure formed will consist of polygonal formations
confined by straight or curved lines after the complete crystallization. The
possible types of boundary lines between adjacent spherulites and their
mathematical descriptions have been analysed in detail [70]. Unfortunately, this
work overlooks an earlier work [222]. Based on all of these works [16, 18, 70,
222-224], the following statements can be made: between two spherulites, A and
B, started at different moments (ίΑ#ίβ) and growing at different rates (KA> KB),
the boundary line can be described by a fourth-degree equation [16]. If growing
is undisturbed, the spherulite growing at a higher rate (A) encircles the other one
(B), which remains as an inclusion [16, 18, 70, 222-224]. In this case, the
boundary consists of two parts: it is a Descartes’ oval up to the ‘shielding point’
[18, 222-224] followed by a logarithmic spiral beyond [18, 222]. If a nucleus of
a spherulite originates on the surface of the growing spherulite B and grows at a
higher rate [15, 16], the boundary relative to the straight line between A and B
will consist of two symmetrical logarithmic spirals [18]. The boundary between
spherulites generated at the same time but growing at different rates is circular
(Appollonius circle) up to the shielding point [16], followed by a logarithmic
spiral [18] behind. Between spherulites growing at identical rates, a hyper¬ bolic
(tA^tB) or linear (iA = £B) boundary is formed [16]. A general statistical
description for the formation and morphology of the spherulitic structure was
given by Piorkowska and Galeski [225]. Several works deal with the statistical
given by Piorkowska and Galeski [225]. Several works deal with the statistical
characterization of the shape and size distributions of spherulites (see [226] and
its references). Several characteristic types of spherulites have been detected by
PLM [5, 8, 9, 70, 216-218]. Since spherulites can be regarded as optically
uniaxial

87 Supermolecular structure of IPP crystals, PLM micrographs show a dark


cross through the centre of the spherulite (the so-called Maltese cross) with
wings in the direction of the planes of the analyser and of the polarizer. The
spherulites can be radial or ringed (banded), depending on the shape of the fibrils
they contain. Radial spherulites are constructed by straight fibrils, while in
banded ones, fibrils are twisted along their longitudinal axes (helical shape),
creating concentric dark circles in the PLM micrograph. The spherulites can be
positive or negative based on the sign of their birefringence. Birefringence (Δη)
of spherulites is defined as [218]: An = nT — nl (4) where nr and nt are the
refractive index in the radial and tangential directions, respectively. The type of
birefringence of spherulites is determined [218] by a first- order red filter
positioned diagonally between crossed nichols (polarizers). During the
crystallization of IPP, a well-developed, optically easily resolved spherulitic
structure is formed. The spherulitic crystallization and structure of IPP have
several particular features, as described in the following sections. Types of
spherulites During the crystallization of IPP, various types of spherulites are
formed that consist of a- and /^-modifications. No supermolecular formations of
the y-modification have been found. In general, y-modification may occur as a
component of α-spherulites in the form of wide-angle fibrillar branches [4, 90,
97]. Under given thermal conditions, a- and jS-spherulites develop
simultaneously. Using knowledge of the critical temperatures of formation of /J-
IPP (Section 3.3.2), it can be established that this is possible in the range
between Τ{αβ) and Τ( βa). It is notable that a simultaneous formation of two
modifications of different stabilities and free enthalpies under identical thermal
conditions can hardly be interpreted on a thermodynamic basis [216] although
similar phenomena were observed for other polymers (e.g. [222, 224]). Padden
and Keith [9] demonstrated that three types of a-spherulites might be formed,
depending on the temperature of crystallization: positive radial (I) below about
406 K, negative radial (II) above about 410 K, and a mixed type in between. The
latter type has no definite birefringence and the Maltese cross does not appear in
the polarized optical micrograph. Forma¬ tion of α-spherulites has a lower limit
temperature. At lower temperatures (TC<350K), the spherulites are replaced by
a smectic structure [50, 194]. Two types of β-spherulites formed sporadically
among the a-modification were demonstrated [9]: a highly birefringent negative
among the a-modification were demonstrated [9]: a highly birefringent negative
radial one (III) below 401 K and negative ringed ones (IV) between 401 and 405
K. Structural

88 Crystallization, melting and supermolecular structure characteristics of /?-


spherulites were discovered by Padden and Keith [9] and by Samuels [13] and
Samuels and Yee [14]. Padden and Keith’s observations were later confirmed
[19, 227], with some alterations in the temperature ranges of the formation of the
individual types, presumably due to the differences in the molecular structures of
IPP grades studied. In our experience, the mixed spherulites of IPP are formed in
a much wider temperature range and the transition temperature between types III
and IV, T(III-IV), is lower. Considering the recent results relating to the ^-
modifica¬ tion, in accord with the experimental observations, it can be
generalized that /^-spherulites of types III and IV are formed between T(<χβ)
and T(III-IV) and between T(III-IV) and Τ(βa), respectively. However, it should
be noted that instead of sharp dividing lines between the temperature ranges of
formation of the different types of spherulites, rather wide temperature zones
exist. The spherulitic structure of IPP varies considerably with Tc and vc, as
illustrated by a schematic diagram in [70]. During the crystallization of samples,
different from the conventional commercial IPP products, new types of
spherulites were also observed [70, 138, 166]. A way a [228] crystallized low
molecular IPP at high temperature (Tc = 420-430 K) and found highly dissimilar
types of spherulites (pseudo¬ positive, pseudo-negative, neomixed, flowerlike,
etc.). Janimak and Cheng [138] observed that the spherulitic crystallization of
low molecular IPP transferred to axialitic when Tc was elevated. Janimak et al
[166] also demonstrated that reducing isotacticity of IPP resulted in a more and
more open and coarse spherulitic structure, which transformed into a dendritic
one at a higher temperature. Other types of unconventional spherulites were
obtained by high temperature self-seeded crystallization [50, 70, 229, 230]. The
spherulitic structure of crystallizable propylene copolymers is some¬ what
different from that of conventional IPP [70]. The types of spherulites and the
temperature range of their formation in BC-PP grades are approxi¬ mately
identical to those in IPP but in the optical pictures finely divided drop-like
formations can be seen on the surface of spherulites. This can be explained by a
phase separation due to the crystallization of chain segments of ethylene
sequences [70, 231]. RC-PP grades are highly supercoolable. During their
crystallization, predominantly positive α-spherulites are formed [70]. The
tendency to ^-crystallization practically ceases due to the disturbed chain
structure (see Section 3.3.2). For this reason, unlike IPP and BC-IPP, only traces
of, if any, j?-spherulites may occur in them. The spherulitic crystallization of
of, if any, j?-spherulites may occur in them. The spherulitic crystallization of
RC-PP was studied more comprehensively by Monasse and Haudin [161]. The
influence of thermal conditions on spherulitic structure The characteristics of the
spherulitic structure of IPP are highly dependent on the thermal pre- and post-
history of the sample. This will be illustrated

Supermolecular structure of IPP 89 by some experimental observations, as


follows. It was demonstrated above that type of IPP spherulites is controlled by
Tc. It was also revealed that the optical character might change even within the
growing spherulites if Tc is not constant during crystallization [9, 15, 70, 229,
233]. For example, a step-wise temperature profile of crystallization may lead to
spherulites comprising gloriolae (annuli) of different optical character and
melting point [70]. Such a sensitive response of the spherulitic structure to the
changes in Tc can be utilized for marking the time schedule of growing crystal
fronts (thermic marking) by a momentary elevation of Tc [232]. The spherulitic
structure is also modified by the temperature changes after crystallization; e.g.
due to heating or annealing, the positive and mixed a-spherulites transform into
negative form [9, 70]. Negative ringed β-spherulites cooled below T£ transform
into positive ringed α-spherulites during heating [61, 62, 70]. An essential
thermal factor to influencing the course of crystalliza¬ tion and the character of
spherulitic texture is the fusion temperature prior to crystallization [50, 66, 70,
115, 194, 229, 230], i.e. the thermal prehistory of the melt. The effects of this
factor were illustrated by several examples in [70] and were attributed to the role
of self-nuclei remained in the melt. Cross-hatching and autoepitaxy during the
growth of a-spherulites The formation of α-spherulites having positive or mixed
birefringence seems to contradict the model for the spherulitic texture.
According to this model, based on the radial growth of fold-chain fibrils
(lamellae) and on the small-angle branching of fibrils, the polymer chains are
nearly or exactly perpendicular to the radii of spherulites. Since the refraction
index is higher in the chain direction than perpendicularly to it, only polymeric
spherulites with negative birefringence would be expected. The structure of
positive and mixed spherulites is interpreted by the formation of a cross-hatched
texture [19, 50, 227, 233-236] consisting of radial (R) and high-angle branched
tangential (T) lamellae. T-lamellae in α-spherulites were detected by electron
microscopy [19, 50, 230, 233-237] and it was determined that they meet the
radius of spherulite at an angle of about 80° [19]. The sign of birefringence of α-
spherulites is controlled by the proportion of T-lamellae. On the basis of the
calculations of Binsbergen and de Lange [50], the optical character of α-
spherulites converts into positive if the proportion of T-lamellae exceeds one-
third. Changes in the fraction of T-lamellae also play a role in the dependence of
third. Changes in the fraction of T-lamellae also play a role in the dependence of
type and birefringence of α-spherulites on Tc. It was actually detected by
electron microscopy that increasing Tc led to a reduced amount of T-lamellae; in
fact, their formation ceased above Tc of about 430 K [19, 166, 227, 230]. At the
same time, the optical character of spherulites shifts in the positive -► mixed -»
negative direction. The transformation of birefringence of positive and mixed

90 Crystallization, melting and supermolecular structure α-spherulites during


heating in the positive -> mixed -> negative and mixed -> negative directions,
respectively, can be explained by partial melting of T-lamellae [9]. It may be
associated with the smaller thickness of T-lamellae as compared to radial ones
[19, 166]. Khoury [235], Sauer, Morrow and Richardson [236], Padden and
Keith [234], Binsbergen and de Lange [50] and Lotz and colleagues [237-239]
have made important contributions to the interpretation of molecular origin of
wide-angle branching of lamellae and the consequent cross-hatched texture. In
particulars, the critical analyses of Lotz and colleagues [237-239] clarified that
the course of branching in a-IPP lamellae was a unique case of autoepitaxy. By
solution crystallization Khoury [235] and Sauer, Morrow and Richardson [236]
obtained a cluster of characteristic twin crystals consisting of intercrossed
elongated lath-like lamellae. This struc¬ ture was named quadrite [236]. In
quadrites, the angle between the intercrossed lamellae is 80°40'. They are located
according to definite crystallographic orientation, namely, axes a± and cx of one
(parent) lamella are parallel with those (c2 and a2) of the other (daughter) one,
(i.e. ax \\c2 and Ci\\a2). The generation of new branches of lamellae (twinning)
takes place on the (010) lateral plane of the lath-like lamellae [233]. (It is notable
that this physical picture is more realistic than that supposed by others [19, 228,
235] which refers to branching of the daughter lamella from the fold surface of
the parent one.) The helical handedness of chains and a perfect interdigita- tion
of methyl groups in the contact planes play important roles in the formation of
the epitaxial branching [50, 237-239]. During normal crystal growth, the
interdigitation of methyl groups cannot be realized unless an antichiral layer is
deposited on the crystal surface involved. If chirality of the deposited layer is
identical to that in the (0 10) plane, the new layer should turn by about 80° or
about 100° to attain a perfect interdigitation of methyl groups. Due to this
rotation of the new isochiral layer, the a1\\c2 and a2\\c1 parallelisms are realized
and a normal further growth will produce the wide-angle lamella branching
(rotation twinning) [238]. It is worth mention¬ ing that y-IPP can crystallize
epitaxially on the lateral plane of a-IPP lamellae in a way similar to the above
(ay-lamellar branching). In this case, however, the angle of branching will be
markedly lower (40°) and the crystallographic orientation of parent and daughter
lamellae will be altered [4, 90]. In /J-IPP, the wide-angle branching of fibrillae
lamellae will be altered [4, 90]. In /J-IPP, the wide-angle branching of fibrillae
do not exist, thus, no cross-hatched texture is formed [19]. Transition in
crystalline modification and anti-epitaxy during spherulitic growth As an
interesting feature of crystal growth of IPP, modification transitions may
proceed on the growing crystal front (bifurcation of growth) under appropriate
conditions. These phenomena were treated in Section 3.3.2 and

Supermolecular structure of IPP 91 will be referred to in the characterization of


polymorphism (Section 3.5), so only an introduction is given here. The transition
of a to β is observed during isothermal spherulitic growth [15, 28] and during
crystallization by the temperature gradient method [25-28]. The «/^-bifurcation
of growth during the shear-induced crystalliza¬ tion is analogous with the above
process (see Section 3.4.2). The β to ol transition on the surface of growing jS-
spherulites may generally occur if Tc is either higher than Τ(βa) [15, 68, 70] or
lower than T(«/?) [69]. On the other hand, in RC-PP with a lower tendency to ^-
crystallization, ^«-bifurcation of growth takes place even under isothermal
conditions. This is well illustrated in Figure 3.13. On the surface of a central /?-
spherulite, «-nuclei are formed in the course of growth which set out to grow
into α-spherulite segments. They can be clearly resolved after the separated
melting of the j?-phase (Figure 3.13b). It can be established from this figure that
the effect of the ^-nucleating agent (which induced the formation of 0-
spherulites) on the course of crystallization stops at a certain stage of growth. In
other words, this is an inverse case of epitaxy, known as anti-epitaxy. The a to y
transition may also be classified into this group of phenomena. It takes place on
the lateral plane of «-lamellae when the wide-angle a-y branches of lamellae are
formed [4, 90]. The general interpretation of these transitions to various
directions upon thermodynamic, kinetic or even structural bases is still open and
unresolved, and only some attempts have been made [68, 238]. 3.4.2 Cylindritic
Structure The cylindritic supermolecular structure is formed by crystallization of
a melt under mechanical load (shear, strain). Under a shear or strain stress, the
molecular chains are extended and ordered into bands. These preordered bands
oriented in the direction of the mechanical load can act as a nucleus of
crystallization. This characteristic, mostly linear type of self-nuclei is named row
nuclei [241]. Row nuclei induce crystallization at higher temperatures than the
heterogeneous and homogeneous nuclei. Due to the expitaxial growth of the
folded-chain fibrils on the row nuclei, characteristic supermolecular formations
of cylindric symmetry, known as cylindrites, are formed [240, 241]. The centre
of a cylindrite is a row nucleus and its structure consists of folded-chain fibrillar
crystallites positioned perpendicu¬ larly to the row nucleus. During the growth,
fibrils branch, which provides complete space filling. Unlike spherulitic
fibrils branch, which provides complete space filling. Unlike spherulitic
crystallization, this fibrillar branching and splaying can only be realized in
planes perpendicular to the row nucleus for steric reasons. From IPP, well-
developed cylinditic structures can be produced, which may be studied by
optical microscopy. This novel supermolecular structure

92 Crystallization, melting and supermolecular structure Figure 3.13 Isothermal


/to-bifurcation of growth in the case of a random propylene copolymer (Moplen
EP1X35F); (a) sample crystallized at TC = 394K; (b) after separate melting of /?
-phase.

93 Supermolecular structure of IPP formed in sheared melts of IPP was first


discovered by Kargin and Andrianova [242] and Binsbergen [243]. Optical
micrographs show these structures to consist of cylindrites of α-modification.
The formation of cylindritic structure was observed even under a small
mechanical load (e.g. by shifting the cover plate over the melt between glass
plates [17, 76] or in flow-through viscometers [77]). A new method for creating
shear stress is fibre pulling [78, 79, 244, 245] where a fibre inactive to
crystallization (e.g. a glass fibre) is inserted into the melt and its pulling
generates the row nuclei. This technique provides an exact adjustment of the
thermal (Tc, vc) and mechanical conditions (Tpull). It is especially useful for
studying the crystalli¬ zation induced by row nuclei formed in situ. Formerly,
the cylinditic structure was produced on thermally stable row nuclei that had
previously undergone a cycle of melting/recrystallization [17, 70]. It is
reasonable to discuss the terminology of cylindritic crystallization. Since the row
nuclei for the cylindritic crystallization are created most often by shear stresses,
the terms ‘shear’ or ‘flow induced crystallization’ or ‘row nucleated
crystallization’ are used [76, 77, 245] as synonyms for cylindritic crystallization.
As the cylindritic crystallization and the transcrystallization [246] show very
similar optical pictures, these phenomena are often not distinguished [78,244].
The principal difference between the two processes will be elucidated later. The
cylindritic crystallization of IPP was studied in detail [17, 70] and the following
characteristic features of this phenomenon were emphasized. Firstly, row nuclei
inducing the cylindritic crystallization have consider¬ able thermal stability,
even at temperatures much higher than the melting point, up to about 470 K.
Consequently, the cylindritic structure is re¬ covered firmly during the repeated
melting/crystallization cycles (structural memory effect). The thermo-optical
method [70] for structural investiga¬ tions of IPP end-products is also based on
the high thermal stability of row nuclei. The stability of row nuclei depends on
the intensity of the mechanical load that induced their formation [17, 70] and,
the intensity of the mechanical load that induced their formation [17, 70] and,
according to recent observations, also on the loading temperature [79]. Secondly,
the sign of optical birefringence of the cylindrites is a function of the
temperatures of crystallization and of annealing, as in the case of spherulites.
The sign of optical birefringence of cylindrites is suitably defined by: Δ n = nT-
na (5) where nr and na are the refraction indices in the radial and axial directions
to the cylindrite, respectively. (In the case of transcrystallization, the sign of
birefringence can be defined again by the difference between the refraction
indices in the perpendicular and in the parallel directions to the surface of
heterogeneous nucleating activity.) The above definition is reasoned by the fact
that in this case, the signs of birefringence of IPP spherulites and

94 Crystallization, melting and supermolecular structure cylindrites and


transcrystalline structure formed under the same thermal conditions are identical.
This is due to the convention that the sign of birefringence comes from the
difference between the refraction index in the direction of the growth of crystal
front and that perpendicular to it. The optical character of cylindrites can be
determined by means of a first-order red filter positioned diagonally between
crossed nichols [17]. Positive cylindrites in the direction of the filter (diagonal)
are yellow, negative ones are blues, and, in mixed cylindrites, blue and yellow
stripes perpendicular to the row nuclei alternate randomly. It was established by
studies on the crystallization induced by thermally stable self-nuclei [17] that the
optical character of α-cylindrites, as a function of Tc, changes identically with
that of spherulites [9, 19]: above 410 K it is positive, below 390 K, it is nearly
positive, and in between optically mixed cylindrites are formed. Cylindrites
inevitably positive in optical character were successfully prepared in RC-PP with
low comonomer content at Tc levels below 400 K. However, the region of
cylindrites in the close vicinity of an α-row-nucleus is negative in birefringence.
This can be explained by the fact that the crystallization on the row nuclei
generally starts at higher temperatures, before reaching Tc. The sign of
birefringence of cylindrites follows the same pattern as that of spherulites during
heating, annealing and recrystallization after fusion at TF (Tm<TF<T£) [17, 70].
Similarly, cylindrites crystallized at a step-wise temperature programme form
strips parallel to the row nucleus of different optical character and melt step-wise
[17]. Thirdly, in general, in the vicinity of a sheared layer, the crystal structure
formed is rich in ^-modification. This modification appears either as
characteristic strips or as triangular segments growing off from the row nuclei.
Leugering and Kirsch [76] first drew attention to the promoting effect of shear
stress on the formation of /^-modification. They demonstrated that the greatest
amount of ^-modification was formed during the crystalli¬ zation of sheared
melt around 393 K. The subsequent investigations, mainly by the fibre pulling
melt around 393 K. The subsequent investigations, mainly by the fibre pulling
technique, defined the thermal and mechanical condi¬ tions of formation of ^-
modification more exactly. The observations of Devaux and Chabert [78] by
PLM were fundamental in this respect. They demonstrated that, at Tc<411 K, ^-
modification was formed both in iso¬ thermal and in non-isothermal
crystallization while at higher temperatures α-crystal fronts were produced. The
latter observation was supported by Thomason and Van Rooyen [245] who
conducted isothermal crystalliza¬ tions only at higher temperatures (Tc^413 K).
The effects of crystallization (Tc) and pulling temperatures (Tpull) on the
formation of /?-phase were also studied [79]. In accord with the statements about
the theoretical upper limit temperature of formation of /MPP [70], it was found
that the isothermal crystallization above Γ(/?α) = 413Κ resulted in α-cylindrites.
However, un¬ der the thermal conditions where Tc = ^pull <Τ(βa), banded
cylindrites,

95 Supermolecular structure of IPP apparently consisting of β-ΙΡΡ, with highly


negative birefringence are formed in the vicinity of pulled fibre. This
characteristic banded structure is shown in Figure 3.14. In samples crystallized
under the above conditions, the optical view of the structure remaining after the
separate melting of the β-modification led to the inevitable conclusion that the
close vicinity of the drawn fibre consisted of α-modification. Consequently,
shear stress gener¬ ated by the fibre pulling produces row nuclei of α-
modification that start to grow a cylindritic front. However, on the surface of the
row nuclei, a great many β-nuclei appear, growing into segments of β-spherulites
(αβ-bifurca- tion of growth). Since the growth rate of β-phase is higher than that
of α-phase (at Tc<T(/?a)), the β-segments block the growth of a-cylindritic
fronts, thus, they remain as inclusions inside the developed β-cylindritic texture.
Between the α-cylindrite and the β-phase developed from puncti- form β-nuclei
on the α-row-nuclei, the boundary lines are straight [17]. They are clearly
perceptible in the micrographs of partially molten samples (Figure 3.14). It is
notable that no banded structure is produced at low Tc values (<395 K). Instead,
negative radial β-cylindrites are formed [79]. The appearance of the banded
structure can be attributed to twisting the fibrils along their longitudinal axis
while growing perpendicularly to the row nuclei, just as in the case of ringed β-
spherulites [9, 14]. An interesting feature of α-row-nuclei created by shear stress
is that they may undergo αβ-transition of modification on their surface under
suitable thermal conditions (Tc = 7pull < Τ(βα)). Inducing /^-modification due to
a shear load is surprising because it is known that mechanical effects on β-ΙΡΡ
samples result in just the opposite, namely the β to a transition. In the necking
stage of drawing, for example, the β-phase transforms into a-phase, as
stage of drawing, for example, the β-phase transforms into a-phase, as
demonstrated by X-ray diffractometry [30-32, 59] and calorimetry [67].
Presumably, the role of α-row nuclei in inducing the formation of /?-phase may
be explained by the special structureof the lateral surface of row nuclei or of
extended-chain crystallites. As a result, the existing energetic condi¬ tions are
favourable for the formation of secondary β-nuclei (secondary αβ-nucleation)
and for the αβ-phase transition on the surface. Consequently, it can be expected
that two different surfaces exist: one of them is involved in the normal growth,
while the other contributes to the processes resulting in the αβ-phase-
transformation. It was also detected [79] that if Tc< Τ(βα)< Tpull, then the
proportion of β-phase will be reduced in the cylindritic front; in fact, at a
properly high shear temperatures where Tpull>450K, only α-cylindrites are
formed after cooling even at Tc < Τ(βa). It can also be established from Figure
3.14 that the row nuclei formed in situ below Τ(βa) are most active at inducing
β-crystallization. The decrease in the activity of row nuclei formed above Τ(βa)
can be explained by the fact that their active surface is covered with layers of α-
modification crystallized above Τ(βa), where only a-IPP can be formed. For
similar reasons, the otherwise thermally stable row nuclei

96 Crystallization, melting and supermolecular structure Figure 3.14 Optical


micrograph of cylindritic crystallization of IPP induced by pulling of two glass
fibres at different temperatures: A, Tpun = 431 K>T(/fa); B, Tpull = Tc =
4O7K<T(0a) [79]. (a) The structure formed at TC = 407K; (b) the structure after
separate melting of 0-phase at T= 433 K.

Supermolecular structure of IPP 91 Figure 3.15 (a) α-transcrystalline structure


formed on quiescent PET fibre at Tc = 398 K; an (b) β-cylindritic structure
induced by pulling of PET fibre at Tc = 408 K.

98 Crystallization, melting and supermolecular structure practically lose their


capability of inducing /^-crystallization (denaturaliz¬ ation) after a
melting/recrystallization cycle (Tm(a)<7V<T£). During rec¬ rystallization after
repeated melting, the row nuclei develop a cylindritic structure of a-modification
[17]. In summary, samples rich in ^-modification may form in sheared melts if
the thermal and mechanical condition of T(a/?)<TC <Tpull <T(/?a) is met. It is
worth mentioning that the high-size supermolecular formations of ^-
modification can be prepared very easily using the fibre pulling technique. In
samples containing both cylindrites and spherulites, the boundary lines between
the two kinds of supermolecular formations have characteristic shapes [17].
Between a cylindrite and a spherulite of identical modification (growing at the
Between a cylindrite and a spherulite of identical modification (growing at the
same rate), parabolic boundary lines are formed. The shape of the parabola is
determined by the relative position of the cylindritic front at the moment of the
formation of the nucleus of spherulite and is independent of the absolute level of
growth rate (consequently, of Tc, too). Between a cylindrite and a spherulite of
different modifications, growing at different rates (Vc and Vs, respectively), the
boundary line is either a hyperbola or an ellipse when Vc < Vs or Vc > V89
respectively. In the latter case, the spherulite, growing at the lower rate, remains
as an inclusion. Because of the shadowing effect of the spherulite formed, the
elliptic shape extends only up to the shielding points. Behind them, in accord
with the analysis of Schulze and Wilbert [18], the shape of the boundary line is a
logarithmic spiral. Quantitative analysis revealed [17] that the boundary line
between the two supermolecular formation is straight if the higher-rate
spherulitic nucleus is formed on a row nucleus or on a growing cylindrite front.
This is the case when a/?- or /^-bifurcation of growth appears on the cylindrite
fronts (cf Figure 3.14b). The angle between the straight line and the row nucleus
depends on the relative growth rate, VT=VC/VS. The transition of modification
on the growing crystal fronts has the kinetic condition that the growth rate of the
new phase is higher than that of the basic front. Consequently, a/?-bifurcation
may proceed only between Τ(βa) and T(ajS) while /fa-bifurcation is possible
below T(ajS) and above T(jSa) in accord with the experimental results discussed
above. It should be pointed out in this respect that sheared no β-cylindritic band
formed in RC-PP having low tendency to ^-crystallization due to the lack of the
kinetic preconditions, i.e. Ga > Gp corresponding to the data in Figure 3.9. 3.4.3
Transcrystalline structure* Crystalline supermolecular structures caused by
oriented growth from heterogeneous surfaces are designated transcrystalline
structures [246]. * Treated also in Chapter 10, Volume 3.

Supermolecular structure of IPP 99 higher on a solid surface in contact with a


polymer melt than in the bulk of Transcrystallization appears when the density
of crystal nuclei is markedly the melt itself. Since spherulitic segments from
nuclei spaced on the surface at high density mutually hinder their lateral growth,
the crystallites are allowed to grow only perpendicularly to the surface
(orientation). As a result, a transcrystalline front will emerge parallel to the
surface. Transcrys¬ tallization is possible if the energetic conditions of
nucleation are more favourable on the surface than in the bulk of the melt. In
other words, the surfaces should have a heterogeneous nucleating activity.
Consequently, in the presence of additives with heterogeneous nucleating
activity, transcrys¬ tallization may occur. In fact, the formation of a
transcrystalline layer was detected experimentally [247, 248] on the surface of α-
nucleating additives with lamellar structure, such as talc [126, 127] or mica
nucleating additives with lamellar structure, such as talc [126, 127] or mica
[148]. This means that the structure and properties of crystalline matrix will be
different whether nucleating or inactive fillers are present. Transcrystallization
of IPP was studied comprehensively in the presence of various fibres.
Transcrystall¬ ization of IPP is induced by polyethylene terephthalate (PET)
[117, 249- 256], polyamide-6 and polyamide-66, [251, 253, 255, 256], aramide
(Kevlar, Twaron) [253, 256-258], some grades of high-modulus carbon [253,
256], cellulose, cotton [259] and some other polymeric fibres [255, 257, 260]. It
should be emphasized that the above fibres are selective α-nucleating agents; for
this reason, transcrystalline bands of α-modification are formed in their vicinity,
as proved by optical micrographs, by the melting point and by the growth rate
values [256]. It is worth noting that the most frequently used grades of glass
fibres induce no transcrystallization in IPP [260]. There is a principal difference
in the feature of nucleation between the transcrystallization and the cylindritic
crystallization discussed in Section 3.4.2. The transcrystallization is induced by
heterogeneous nuclei, while the cylindritic crystallization originates from self-
nuclei (i.e. row nuclei having identical chemical and physical structure to the
crystalline phase to be formed). This difference in the feature of nucleation is
reflected by the formation process of the structure, as demonstrated for IPP
samples crystallized isothermally between T( βά) and Τ(αβ) in Figure 3.15. A
PET fibre with nucleating activity induces (in the quiescent state) the forma¬
tion of an α-transcrystalline front around the filament (Figure 3.15a). Under
identical thermal conditions, if the PET fibre is pulled during the crystalliza¬
tion (Tc < Tpull < Τ(βa)), row nuclei will form in the vicinity of the filament,
resulting in a j?-cylindritic band (Figure 3.15b). This band comprises the region
of the α-row-nuclei responsible for the formation of the cylindrite, as evidenced
by the melting experiments [261]. These experimental observa¬ tions emphasize
the distinction between transcrystallization and row nu¬ cleated cylindritic
crystallization. It is worth noting that no jS-transcrystal- line structure has been
found to date, since this would require the presence of a surface having selective
^-nucleating activity.

100 Crystallization, melting and supermolecular structure 3.5


POLYMORPHISM IN IPP There are two types of polymorphism: monotropy
and enantiotropy. The difference between them is usually illustrated by the
temperature depend¬ ence of the free enthalpy, G [262, 263] Figure 3.16 shows
the function G=/(T) linearly for the sake of simplification for both the
amorphous and crystalline modifications (i.e. a and β) of the material. In the case
of monotropy, only one of the modifications (a) can be stable thermodynami¬
cally since its G value is always lower below the melting point, Tm(a), than that
cally since its G value is always lower below the melting point, Tm(a), than that
of the other one (β). Enantiotropy permits thermodynamically stable states (with
lower G) for both modifications as a function of temperature. It can be seen from
Figure 3.16b that, in the case of enantiotropy, the amor¬ phous state is stable
thermodynamically above Tm(a) while a- and β- modifications are stable from
Tm(a) to T(/?a) and below Τ(βa), respectively. Τ(βa) is the equilibrium
temperature of the solid-state transition between the β and the a phase. _j | | j x
TM Tffiot) Tm(fi) Tm(ot) T Figure 3.16 Possible temperature dependences of
free enthalpy (G) of the α-, β- and amorphous phase in the case of (a) monotropy
and (b) enantiotropy.

7 101 Polymorphism in IPP Unequivocal determination of the type of


polymorphism is difficult for polymers since the solid-solid phase transitions are
kinetically hindered and equilibrium states and structures usually fail to form.
This means that non¬ equilibrium, thermodynamically unstable structures may
exist permanently. Determination of the thermodynamical stability of various
IPP modifi¬ cations is still unknown. According to much of the literature, out of
the various modifications of IPP, only « is thermodynamically stable. This
assumption is based on the observation that the directions of polymorphic
transitions are generally β to a [9, 30-32, 49, 58] or y to a [85-87]. We are
unaware of any experimental data in the literature about the relative stabilities of
β- and y-modifications and about their transformation into each other. It may be
attributed to the quite different molecular structural conditions from the point of
view of β- and y-crystallization. The tendency to y-crystallization is
characteristic of IPP samples of more irregular structure [4], such as degraded
IPP or RC-PP with 4-8% comonomer content. On the other hand, the tendency
to ^-crystallization may be suppressed by disturbance of chain regularity (Figure
3.9) /^-transitions observed earlier during melting [9, 27, 40], annealing [12, 47]
or deformation [30-32, 58] of the ^-modification seemed to support the opinion
that the thermodynamically stable state is the α-modification. This suggests that
IPP is monotropic. In view of the more recent observations, however, this
question should be analysed critically. As was shown above, jS-IPP has a
characteristic dual feature during melting and annealing, i.e. it behaves like
either a thermodynamically stable or an unstable modification, depending on its
thermal history. These features are manifested by the melting [62] and annealing
[64] memory effect. The observations on the upper and lower limit temperatures
of formation of jS-IPP and its dual feature lead to the concept that the ^-
modification is thermodynamically stable between Τ(βά) and Τ(αβ). In other
words, it is a peculiar case of enantiotropy, as illustrated schematically in Figure
3.16. This may be realized if the free enthalpy of the a-modification follows the
3.16. This may be realized if the free enthalpy of the a-modification follows the
broken line in the temperature range T< Τ{βa). Accordingly, the ^-modification
has a minimum free energy and is thermodynamically stable between Τ(αβ) and
Τ(βa). On the other hand, the α-modification is stable below Τ(οίβ) and between
Τ(βa) and Tm(a). It is worth noting that the β to a transition, which is the
apparent evidence of monotropy, is observed only in the aforementioned
stability range of the α-modification. Actually, during partial melting or high
temperature annealing, the ^«-transition takes place above Τ(βa) (if nuclei of the
a-phase formed below T$ are present). The ^«-transition in the necked part of a
cold-drawn jS-IPP sample was observed when deformed below T(«/?). The
occasional « to β transition on the growing crystal fronts (a/?- bifurcation of
growth) is not in accord with the supposed monotropy. This phenomenon was
found during spherulitic growth [15, 28], crystallization

102 Crystallization, melting and supermolecular structure by the temperature


gradient method [25-28] and row nucleated cylindritic crystallization [76-79].
Although these phenomena were discussed above (Section 3.4.2), we designed a
special thermo-optical experiment in order to demonstrate the β a transitions
jointly. The results of this are shown in Figure 3.17. In this experiment, row
nucleated cylindritic crystallization was induced by pulling a glass fibre
embedded in the molten polymer. Tempera¬ tures of crystallization and pulling
were below Τ(βa). Under these condi¬ tions, an a/?-bifurcation of growth took
place on the primary α-row nuclei and a growth of a j8-cylindritic front started
(Figure 3.17a). After that, Tc was elevated above T(/fa), which caused a /fa-
transition of modification on the β-cylindritic crystal front (Figure 3.17b). The
structure of a-modification, remained after the two-step melting of the ^-
modification, clearly illustrates that a/?-transition occurred below T(/fa),
whereas /fa-transition occurred above Τ(βa). Series of experiments proved the
onset of a to β and β to <x transitions around T(/fa), which is characteristic to
enantiotropy. Accordingly, the temperature T(/fa) can be regarded as a
thermodynamic equilibrium tem¬ perature of a solid-solid phase transition.
Unfortunately, similar tests cannot be conducted around T{<χβ) because of the
high supercooling and the rapid crystallization. The author is aware of the
objections to the above hypothesis. It may be a straightforward complaint that
the results for the apparent evidence for enantiotropy were obtained from
samples containing ^-nucleating agent or from crystallization in sheared melts
(though the accompanying jS-IPP in the conventional IPP products behaved
similarly). However, the author knows of no experimental observations that
would contradict unequivocally the assumption of enantiotropy. The validity of
this theory should be determined in further investigations. The author attempted
this theory should be determined in further investigations. The author attempted
to prove it by direct experiments. Started from the fact that the ^-modification is
stable between Τ(αβ) and T(/fa), it would be expected that, by drawing the
sample in this temperature range, /?/?-recrystallization would take place in the
orientation stage and /fa-recrystallization at other Td temperatures of
deformation. In Figures 3.18 and 3.19 calorimetric melting curves are shown as
a function of the relative elongation for samples obtained from the necking part
of a specimen subjected to drawing and orientation recry¬ stallization at room
temperature [Td<T( a/?)] and at 120 °C [Τ(αβ)<Τά<Τ(β<χ)]. It is worth noting
that compression moulded speci¬ mens were used and susceptibility to /fa-
recrystallization due to recooling was eliminated annealing [64], started from the
temperature of compression moulding. It can be seen that, by increasing the
relative elongation, an increasing proportion of /MPP transforms into α-
modification and the /fa-transition becomes almost complete at break. It is also
clear from the data of Figure 3.19 that, regardless of the temperature of drawing,
in contrast to our presumption, a /fa-recrystallization takes place in the

Polymorphism in IPP 103 Figure 3.17 The α,β- and /fa-transition around Τ(βa)
(see text): (a) shear induced isothermal crystallization at Tc = Tpull = 408 K; (b)
further growth at TC = 418K; (c) the structure remained after separate melting of
the /?-phase.

104 Crystallization, melting and supermolecular structure Figure 3.17 Contd.


necking stage. Although these experimental results do not prove enantiot- ropy,
neither do they deny it. Stress as a thermodynamic parameter alters the course of
the function G=/(T). For this reason, Τ(<χβ) and Τ(βa) levels may be modified
under stress or the stability range of the ^-modification may even disappear. It
was demonstrated that disturbing the chain regularity of IPP (copolymerization)
reduced the tendency to β-crystallization. For this reason, our assumption of
enantiotropy may be perfectly valid for com¬ pletely regular, ideally isotactic
polymers only. According to our present knowledge, monotropy can also explain
the presence of both a- and ^-modifications. The apparent stability of non-
recooled β-ΙΡΡ is attributed to the kinetic and energetic hindrance of the /fa-
phase transition, since an appropriate molecular mobility and the pre¬ sence of
α-nuclei are necessary for the β to a phase transition. These requirements are met
during the melting and annealing of samples cooled to T&, since below T£ α-
nuclei are formed due to a secondary crystallization and the molecular mobility
is provided by a partial melting. In the case of monotropy, the kinetic concept of
the characteristic Τ(βa) and Τ(αβ) temperatures becomes conspicuous, since
below Τ(οίβ) and above Τ(βοΐ) the essential kinetic precondition of the /fa-
transition, namely, Gflt>G/?, is fulfilled. It can be concluded, therefore, that the
transition, namely, Gflt>G/?, is fulfilled. It can be concluded, therefore, that the
type of polymorphism in poly¬ propylene requires further investigation.

Appendix 105 Figure 3.18 Melting curves of necked /MPP samples after cold-
drawing (Td — Trt) at different tensile strains (%). 3.6
ACKNOWLEDGEMENTS This work was supported by the Hungarian Science
Foundation (OTKA) and by the Alexander von Humboldt Foundation, in the
form of a common research project between the Department for Plastics and
Rubber Industry at the Technical University of Budapest and the Institute for
Composite Materials at the University of Kaiserslautern. 3.7 APPENDIX:
SYMBOLS AND ABBREVIATIONS A\ A<X9 Αβ Isotropic amorphous phase
(self-nuclei free) Amorphous phase with residual orderedness of a- or /Lohase

106 Crystallization, melting and supermolecular structure Figure 3.19 Melting


curves of necked β-IPP samples after hot-drawing (Td = 393 K) at different
tensile strains (%). BC-PP Propylene block copolymer Ca, Cf Crystalline a- or
j8-phase DTA Differential thermal analysis G Free enthalpy G Growth rate G„,
Gfi Linear growth rate of a- or jS-phase IR Infrared K Rate constant of
crystallization k Empirical value to measure of the /^-content l Lamellar
thickness n Avrami’s exponent

References 107 Refractive index in radial, tangential and axial directions,


respectively Birefringence Density of athermal nuclei Nuclear magnetic
resonance Polyethylene terephthalate Polarized light microscopy Propylene
random copolymer Time Annealing temperature Crystallization temperature
Upper and lower temperature of the formation of pure /MPP, respectively Initial
and final temperature of anisothermal crystallization, respectively Final
temperature of fusion Fusion temperature of the formation of self-nuclei free
melt Experimental melting temperature of a-modification Experimental melting
temperature of ^-modification Equilibrium melting temperature of a-
modification Equilibrium melting temperature of /^-modification Recooling and
critical recooling temperature Theoretical lower limit temperature of the
formation of 0-IPP Theoretical upper limit temperature of the formation of /MPP
Transition temperature between growth regimes I and II Transition temperature
between growth regimes II and III Cooling and heating rate, respectively Growth
rate of spherulite and cylindrite, respectively Relative portion of crystallized
material Proportion of crystallized material at time t and ^fter primary
crystallization α-modification of isotactic polypropylene a to β recrystallization
^-modification of isotactic polypropylene β to a recrystallization ^-modification
of isotactic polypropylene Experimental heat of fusion Equilibrium heat of
of isotactic polypropylene Experimental heat of fusion Equilibrium heat of
fusion of a-phase Equilibrium heat of fusion of β-phase Supercooling; AT =Tm°
— Tc Temperature range of anisothermal crystallization Lateral surface free
energy Fold surface free energy

108 Crystallization, melting and supermolecular structure 3.8 REFERENCES 1.


Vieweg, R., Schley, A. and Schwarz, A. (1969) Polyolefine, Kunststoff
Handbuch Vol. IV, Hasder-Verlag, Munich. 2. Jezl, J. L. and Honeycutt, E. M.
(1969) Propylene polymers, in Encyclopedia of Polymer Science and
Technology, Vol. II (eds H. F. Mark and N. G. Gaylord), Wiley-Interscience,
New York. 3. Barbe, P. C. and Lieberman, R. B. (1988) Propylene polymers, in
Encyclopedia of Polymer Science and Engineering, 2nd edn, Vol. 13, Wiley-
Interscience, New York, pp. 464-531. 4. Bruckner, S., Meille, S. V., Petraccone,
V. and Pirozzi, B. (1991) Progress in Polymer Science, 16 361-404. 5.
Wunderlich, B. (1973) Macromolecular Physics, Vol. 1, Academic Press, New
York. 6. Wunderlich, B. (1976) Macromolecular Physics, Vol. 2, Academic
Press, New York. 7. Wunderlich, B. (1980) Macromolecular Physics, Vol. 3,
Academic Press, New York. 8. Geil, P. H. (1963) Polymer Single Crystals,
Wiley-Interscience, New York. 9. Padden, F. J. and Keith, H. D. (1959) Journal
of Applied Physics, 30, 1479-84. 10. Keith, H. D., Padden, F. J, Walter, N. M.
and Wyckoff, H. W. (1959) Journal of Applied Physics, 30, 1485-91. 11.
Addink, E. J. and Beintema, J. (1961) Polymer, 2, 185-93. 12. Turner Jones, A.,
Aizlewood, J. M. and Beckett, D. R. (1964) Makromolekulare Chemie, 75, 134-
54. 13. Samuels, R. J. (1967) Journal of Polymer Science, Part C, 20, 253-84. 14.
Samuels, R. J. and Yee, R. J. (1972) Journal of Polymer Science, Polymer
Physics Edition, A2, 10. 385-432. 15. Varga, J. (1982) Angewandte
Makromolekulare Chemie, 104 79-87. 16. Varga, J. (1983) Angewandte
Makromolekulare Chemie, 112, 161-72. 17. Varga, J. (1983) Angewandte
Makromolekulare Chemie, 112, 191-203. 18. Schulze, G. E. and Wilbert, Η. P.
(1989) Colloid and Polymer Science, 267, 108-15. 19. Norton, D. R. and Keller,
A. (1985) Polymer, 26, 704-16. 20. Geil, P. H. (1962) Journal of Applied
Physics, 33, 642-3. 21. Olley, R. H. (1986) Scientific Progress (Oxford), 70, 17-
43. 22. Merajev, S. D., Wunder, S. L. and Wallace, W. (1985) Journal of
Polymer Science, Polymer Physics Edition, 23, 2043-57. 23. Beckett, D. R.,
Chalmers, J. M. Mackenzie, M. W. et al. (1985) European Polymer Journal, 21,
848-52. 24. Gomez, M. A., Tanaka, H. and Tonelli, A. E. (1987) Polymer, 28,
2227-32. 25. Crissman, J. M. (1969) Journal Polymer Science, A2 (7), 389-404.
26. Kamide, K. and Nakamura, K. (1969) Sen-i Gakaishi, 25, 53-9. 27. Fujiwara,
Y. (1975) Colloid and Polymer Science, 253, 273-82. 28 Lovinger, A. J., Chua,
J. O. and Gryte, C. C. (1977) Journal of Polymer Science, Polymer Physics
Edition, 15 641-56. 29. Fujiwara, Y. (1991) A survey of temperature slop
Edition, 15 641-56. 29. Fujiwara, Y. (1991) A survey of temperature slop
crystallization for the oriented crystallization of polymer lamellae, Shizuoka
University, Japan. 30. Asano, T. and Fujiwara, Y. (1978) Polymer, 19, 99-108.
31. Asano, T., Fujiwara, Y. and Yoshida, T. Polymer Journal, 11, 383-90. 32.
Yoshida, T., Fujiwara, Y. and Asano, T. (1983) Polymer, 24, 925-9. 33.
Fujiwara, Y. (1987) Polymer Bulletin, 17, 539-43. 34. Fujiwara, Y., Goto, T. and
Yamashita, Y. (1987) Polymer, 28, 1253-6.

References 109 35. Fujiwara, Y. (1987) Colloid and Polymer Science, 265,
1027-8. 36. Leugering, H. J. (1967) Makromolekulare Chemie, 109, 204-16. 37.
Turner Jones, A. and Cobbold, A. J. (1968) Polymer Letters, 6 539-46. 38.
Duswalt, A. (1970) American Chemical Society, Division of Organic Coatings,
30, 93-6. 39. Ullmann, W. and Wendorff, J. H. (1979) Progress in Colloid and
Polymer Science, 66, 25-33. 40. Forgacs, P., Tolochko, B. P. and Sheromov, M.
A. (1981) Polymer Bulletin, 6, 127-33. 41. Moos, K. H. and Tilger, B. (1981)
Angewandte Makromolekulare Chemie, 94, 213-25. 42. Moos, K. H. (1983)
Nukleierung und Polymorphie in isotaktischem Polypropy¬ lene, Dissertation,
Darmstadt. 43. Moos, K. H. and Jungnickel, B. J. (1985) Angewandte
Makromolekulare Chemie, 132, 135-60. 44. Jacoby, P., Bersted, B. H., Kissel,
W. J. and Smith, C. E. (1986) Journal of Polymer Science, B24, 461-91. 45.
Garbarczyk, J. and Paukszta, D. (1981) Polymer, 22, 562-4. 46. Garbarczyk, J.
(1985) Makromolekulare Chemie, 186, 2145-51. 47. Garbarczyk, J. and
Paukszta, D. (1985) Colloid and Polymer Science, 263, 985-90. 48. Garbarczyk,
J. (1989) Polymer Communications, 30, 153-7. 49. Rybnikar, F. (1991) Journal
of Macromolecular Science, Physics, B30, 201-23. 50. Binsbergen, F. L. and de
Lange, B. G. M. (1968) Polymer, 9, 23-40. 51. Morrow, D. R. (1969) Journal of
Macromolecular Science, Physics, B3, 53- 65. 52. Hughes, J. K. (1970)
Laminates of similarly constituted films of different crystal structure, US Patent
3540979, 17th November. 53. Shi, G., Zhang, J. and Jins, H. (1987) Chemical
Abstracts, 106. 5666P. 54. Shi, G., Huang, B. and Zhang, J. (1984)
Makromolekulare Chemie, Rapid Communications, 5, 573-8. 55. Gui-en Zhau,
Zhi qun He, Jian-min Yu and Zwe-wen Han (1986) Mak¬ romolekulare Chemie,
187, 633-42. 56. Shi, G., Huang, B., Cao, Y. et al. (1986) Makromolekulare
Chemie, 117, 643-52. 57. Shi, G., Huang, B., Zhang, J. and Cao, Y. (1987)
Scientia Sinica, Series B, 30, 225-33. 58. Shi, G., Chu, F., Zhou, G. and Han, Z.
(1989) Makromolekulare Chemie, 190, 907-913. 59. Shi, G., Zhang, X. and Qiu,
Z. (1992) Makromolekulare Chemie, 193, 583-91. 60. Jingjiang, L., Xiufen, W.
and Qipeng, G. (1990) Journal of Applied Polymer Science, 41, 2829-35. 61.
Varga, J. and Ille, J. (1986) Magyar Kemiai Folydirat, 92,180-6, (in Hungarian).
62. Varga, J. (1986) Journal of Thermal Analysis, 31, 165-72. 63. Varga, J.,
62. Varga, J. (1986) Journal of Thermal Analysis, 31, 165-72. 63. Varga, J.,
Garzo, G. and Ille, A. (1986) Angewandte Makromolekulare Chemie, 142, 171-
81. 64. Varga, J. and Toth, F. (1986) Makromolekulare Chemie,
Makromolecular Symposia, 5, 213-23. 65. Varga, J. and Garzo, G. (1991) Acta
Chimica Hungarica, 128, 303-17. 66. Varga, J., Schulek-Toth, F. and Ille, A.
(1991) Colloid and Polymer Science, 269, 655-64. 67. Varga, J. (1989) Journal
of Thermal Analysis, 35, 1891-1912.

110 Crystallization, melting and supermolecular structure 68. Varga, J.,


Fujiwara, Y. and Ille, A. (1990) Periodica Polytechnica, Chemical Engineering,
34, 256-71. 69. Lotz, B., Fillon, B., Therry, A. and Wittmann, J. C. (1991)
Polymer Bulletin, 25, 101-5. 70. Varga, J. (1992) Journal of Materials Science,
27, 2557-79. 71. Varga, J., Schulek-Toth, F. and Pati, M. (1992) Method for the
preparation of ^-modification of isotactic polypropylene (homopolymers,
copolymers and compounds containing its) in pure form, Hungarian Patent
Application, P 92. 01422, 29th April. 72. Fujiyama, M., Kawamura, Y., Wakino,
T. and Okamoto, T. (1988) Journal of Applied Polymer Science, 36, 985-93,
995-1009, 1011-23, 1025-34, 1035^18, 1049-59, 1061-6. 73. Varga, J. and
Garzo, G. (1990) Angewandte Makromolekulare Chemie, 180, 15-33. 74. Varga,
J. and Schulek-Toth, F. (1991) Angewandte Makromolekulare Chemie, 188, 11-
25. 75. Varga, J. (1991) Journal of Polymer Engineering, 1, 231-51. 76.
Leugering, H. J. and Kirsch, G. (1973) Angewandte Makromolekulare Chemie,
33, 17-23. 77. Grubb, D. T., Odell, J. A. and Keller, A. (1975) Journal of
Materials Science, 10 (1975) 1510-18. 78. Devaux, E. and Chabert, B. (1991)
Polymer Communications, 32, 464-8. 79. Varga, J. and Karger-Kocsis, J. (1993)
Composites Science and Technology, 48, 191-8, and (1993) Polymer Bulletin,
30, 105-10. 80. Pae, K. D., Morrow, D. R. and Sauer, J. A. (1966) Nature, 211,
514-15. 81. Kardos, J., Christiansen, A. W. and Bear, E. A. (1966) Journal of
Polymer Science, A2 (4), 777-88. 82. Awaya, H. (1966) Journal of Polymer
Science, Polymer Letters, 4, 127-30. 83. Kojima, M. (1967) Journal of Polymer
Science, B5, 245-50. 84. Kojima, M. (1968) Journal of Polymer Science, A2 (6),
1255-71. 85. Pae, K. D. (1968) Journal of Polymer Science, A2 (6), 657-63. 86.
Morrow, D. R. and Newman, B. A. (1968) Journal of Applied Physics, 39, 4944-
60. 87. Newman, B. A. and Song, S. (1971) Journal of Polymer Science, A2 (9),
181-6. 88. Turner-Jones, A. (1971) Polymer, 12, 487-508. 89. Kajima, M. and
Mori, K. (1972) Journal of Polymer Science, Polymer Physics Edition, A2 (10),
1171—4. 90. Lotz, B., Graff, S. and Wittmann, J. C. (1986) Journal of Polymer
Science, Polymer Physics Edition, 24, 2017-32. 91. Bruckner, S. and Meille, S.
V. (1989) Nature, 340, 455-7. 92. Bruckner, S., Meille, S. V., Sozzani, P. and
Torri, G. (1990) Makromolekulare Chemie, Rapid Communications, 11, 55-60.
Torri, G. (1990) Makromolekulare Chemie, Rapid Communications, 11, 55-60.
93. Meille, S. V., Bruckner, S. and Porzio, W. (1990) Macromolecules, 23,
4114-21. 94. Rieger, B., Mu, X., Mallin, D. T. et al. (1990) Macromolecules, 23,
3559. 95. Marigo, A., Marega, C., Zanetti, R. et al. (1989) Macromolekulare
Chemie, 190, 2805-13. 96. Marigo, A., Marega, C., Zanetti, R. et al. (1990)
Makromolekulare Chemie, 191, 1967-71. 97. Nedkov, E. and Krestev, V. (1990)
Colloid and Polymer Science, 268, 1028-35. 98. Hoffman, J. D., Davies, G. T.
and Lauritzen, J. J. (1976) The rate of crystalliza¬ tion of linear polymers with
chain folding, in Treatise on Solid State Chemistry, Vol. 3 (ed. N. B. Hannay),
Plenum Press, New York, pp. 497-614.

References 111 99. Armistea, K., Goldbeck, G. and Keller, A. (1992) Advances
in Polymer Science, 100, 221-312. 100. Marker, L., Hay, F. M., Tilley, G. P. et
al (1959) Journal of Polymer Science, 38, 33-43. 101. Griffith, J. H. and RAnby,
B. G. (1959) Journal of Polymer Science, 38, 107-16. 102. Von Falkay, B. and
Stuart, H. A. (1959) Kolloid Zeitschrift, 162, 138-40. 103. Von Falkay, B.
(1960) Makromolekulare Chemie, 41, 86-109. 104. Mayer, J. (1960)
Kunststojfe, 50, 565-7. 105. Parrini, P. and Corrier, C. (1961) Makromolekulare
Chemie, 62, 83-96. 106. Magill, J. H. (1962) Polymer, 3, 35-42. 107. Majer, J.
(1963) Kunststoff-Rundshau, 10, 9-11. 108. Donald, H. J., Humes, E. S. and
White, L. W. (1964) Journal of Polymer Science, C6, 93-9. 109. Hoshino, S.,
Meinecke, E., Powers, J. et al (1965) Journal of Polymer Science, A3, 3041-65.
110. Gordon, M. and Hillier, H. (1965) Polymer, 6, 213-19. 111. Beck, Η. N.
(1967) Journal of Applied Polymer Science, 11, 673-85. 112. Beck, Η. N. and
Ledbetter, H. D. (1965) Journal of Applied Polymer Science, 9, 2131-42. 113.
Godowsky, J. K. and Slonimsky, S. L. (1966) Vysokomolekulyarnie
Soedineniya (Polymer Science USSR), 8, 403-10 (in Russian). 114. Slonimsky,
G. L. and Godowsky, J. K. (1966) Vysokomolekulyarnie Soedineniya (Polymer
Science USSR), 8, 718-21 (in Russian). 115. Collier, J. R. and Neal, L. M.
(1968) Annual Technical Conference of the Society for Plastics Engineering
Technology, Paper 26, pp. 63-8. 116. Rybnikar, F. (1969) Journal of Applied
Polymer Science, 13, 827-33. 117. Binsbergen, F. L. and de Lange, B. G. M.
(1970) Polymer, 11, (309-28). 118. Johnsen, U., Spilgies, G. and Zachmann, H.
G. (1970) Kolloid Zeitschrift und Zeitschrift fur Polymere, 240, 762-5. 119.
Johnsen, U. and Spilgies, G. (1972) Kolloid Zeitschrift und Zeitschrift fur
Polymere, 250, 1174-1174. 120. Godowsky, J. K. and Slonimsky, G. L. (1974)
Journal of Polymer Science, 12, 1053-80. 121. Pratt, C. F. and Hobbs, S. J.
(1976) Polymer, 17, 12-16. 122. Ishizuka, O. and Koyama, K. (1977) Polymer,
18, 913-18. 123. Crispino, L., Martuscelli, E. and Pracella, M. (1980)
Makromolekulare Chemie, 181, 1747-55. 124. Varga, J., Menczel, J. and Solti,
Makromolekulare Chemie, 181, 1747-55. 124. Varga, J., Menczel, J. and Solti,
A. (1981) Journal of Thermal Analysis, 20, 23-32. 125. Wlochowicz, A. and
Eder, M. (1981) Polymer, 22, 1285-7. 126. Rybnikar, F. (1982) Journal of
Applied Polymer Science, 27, 1479-86. 127. Menczel, J. and Varga, J. (1983)
Journal of Thermal Analysis, 28, 161-74. 128. Avella, M., Martuscelli, E. and
Pracella, M. (1983) Journal of Thermal Analysis, 28, 237-48. 129. Martuscelli,
E., Pracella, M. and Crispino, L. (1983) Polymer, 24, 693-9. 130. Eder, M. and
Wlochowicz, A. (1983) Polymer, 24, 1593-5. 131. Wlochowicz, A. and
Malinowska-Grabos, Z. (1983) Acta Polymerica, 34, 166-8. 132. Varga, J.,
Menczel, J. and Solti, A. (1984) Vysokomolekulyarnie Soedineniya (Polymer
Science USSR), 26A, 2467-76. 133. Chalvert, P. D. and Rian, T. G. (1984)
Polymer, 25, 921-6. 134. Carfagna, C., De Rosa, C., Guerra, C. and Petraccone,
V. (1984) Polymer, 25, 1462-4.

112 Crystallization, melting and supermolecular structure 135. Kowalewski, T.


and Galeski, A. (1986) Journal of Applied Polymer Science, 32, 2919-34. 136.
Acosta, J. L., Ojeda, M. C., Morales, E. and Linares, A. (1986) Journal of
Applied Polymer Science, 32, 4119-26. 137. Wlochowicz, A. and Eder, M.
(1989) Angewandte Makromolekulare Chemie, 171, 79-89. 138. Janimak, J. J.
and Cheng, S. Z. D. (1989) Polymer Bulletin, 22, 95-101. 139. Martuscelli, E.,
Silvestere, C., Canetti, M. et al (1989) Makromolekulare Chemie, 190, 2615-25.
140. Mitsuishi, K., Ueno, S., Kodama, S. and Kawasaki, H. (1991) Journal of
Applied Polymer Science, 43, 2043-9. 141. Horng-jer Tai, Wen-Yen Chiu, Leo-
Wang Chen and Line-Hwa Chu (1991) Journal of Applied Polymer Science, 42,
3111-22. 142. Young Chul Kim, Chung Yup Kim and Sung Chul Kim (1991)
Polymer Engineering Science, 31, 1009-14. 143. Jurado, J. R., Maure, C.,
Durana, P. et al (1991) Journal of Materials Science, 26, 4022-5. 144. Janimak,
J., Cheng, S. Z. D., Zhang, A. and Hsieh, E. T. (1992) Polymer, 33, 728-35. 145.
Varga, J., Biro, F., Schulek-Toth, F. et al. (1990) Miianyag is Gumi, 27, 133-42
(in Hungarian). 146. Monasse, B. and Haudin, J. M. (1986) Colloid and Polymer
Science, 264,117-22. 147. Kargin, V. A., Sogolova, T. J., Rapoport, N. J. and
Kurbanova, J. J. (1967) Journal of Polymer Science, Polymer Symposia, 16,
1609-17. 148. Garton, A., Kim, S. W. and Wiles, D. M. (1982) Journal of
Polymer Science, Polymer Letters, 20 273-8. 149. Al-Ghazawi, M. and Sheldon,
R. P. (1983) Journal of Polymer Science, Polymer Letters, 21, 347-51. 150.
Bartczak, Z. and Galeski, A. (1990) Polymer, 31, 2027-38. 151. Binsbergen, F.
L. (1973) Journal of Polymer Science, Polymer Physics Edition, 11, 117-35.
152. Limbert, F. J. and Baer, E. (1963) Journal of Polymer Science, Al, 3317-31.
153. Keith, H. D. and Padden, F. J. (1964) Journal of Applied Physics, 35, 1286-
96. 154. Chang, E. P. (1977) Journal of Applied Polymer Science, 21, 937-42.
155. Goldfarb, L. (1978) Makromolekulare Chemie, 179, 2297-303. 156.
155. Goldfarb, L. (1978) Makromolekulare Chemie, 179, 2297-303. 156.
Martuscelli, E., Silvestere, C. and Abate, G. (1982) Polymer, 23, 229-37. 157.
Galeski, A., Pracella, M. and Martuscelli, E. (1984) Journal of Polymer Science,
Polymer Physics Edition, 22, 739-47. 158. Bartczak, Z. and Galeski, A. (1984)
Polymer Engineering and Science, 24,1155-65. 159. Clark, E. J. and Hoffman, J.
D. (1984) Macromolecules, 17, 878-85. 160. Monasse, B. and Haudin, J. M.
(1985) Colloid and Polymer Science, 263, 822-31. 161. Monasse, B. and
Haudin, J. M. (1988) Colloid and Polymer Science, 266, 679-87. 162. Bartczak,
Z., Galeski, A. and Krasnikova, N. P. (1987) Polymer, 28, 1627-34. 163. Allen,
R. C. and Mandelkern, L. (1987) Polymer Bulletin, 17, 473-80. 164. Atanassov,
A. M. (1987) Polymer Bulletin, 17, 445-1. 165. Pawlak, A. and Galeski, A.
(1990) Journal of Polymer Science, Polymer Physics Edition, B28, 1813-21.
166. Janimak, J. J., Cheng, S. Z. D., Giusti, P. A. and Hsieh, E. T. (1991)
Macromolecules, 24, 2253-60. 167. Ghijsels, A., Groesbeek, N. and Yip, C. W.
(1982) Polymer, 23, 1913-16. 168. Martuscelli, E., Silvestre, C. and Bianchi, L.
(1983) Polymer, 24, 1458-68.

References 113 169. Wenig, W., Fiedel, H. W. and Scholl, A. (1990) CoHmd
and Polymer Science, 268, 528-35. X 170. Wenig, W. and Fiedel, H. W. (1991)
Makromolekulare Chemie, 192, 191-9. 171. Pukanszky, B, Tiidds, F., Kallo, A.
and Bodor, G. (1989) Polymer, 30, 1399- 1406. 172. Martuscelli, E., Pracella,
M., Volpe, G. D. and Greco, P. (1984) Mak¬ romolekulare Chemie, 185, 1041-
61. 173. Eder, M. and Wlochowicz, A. (1984) Acta Polymerica, 35, 548-53. 174.
Galeski, A., Batrczak, Z. and Pracella, M. (1984) Polymer, 25, 1323-6. 175.
Avrami, M. (1939-1941) Journal of Chemical Physics, 7, 1103-12; 8, 212-24; 9,
177-84. 176. Grenier, D. and Prudhomme, R. E. (1980) Journal of Polymer
Science, Polymer Physics Edition, 18, 1655. 177. Cheng, S. Z. D. and
Wunderlich, B. Macromolecules, 21, 3327-8. 178. Perez-Cardenas, F. C.,
Castillo, L. F. and Vera Graziano, E. (1991) Journal of Applied Polymer
Science, 43, 779-82. 179. Malkin, A. J., Beghisev, V. P. and Keapin, J. A.
(1983) Polymer, 24, 81-4. 180. Ozawa, T. (1971) Polymer, 12, 150-8. 181.
Nakamura, K., Katayama, K. and Amano, T. (1973) Journal of Applied Polymer
Science, 17, 1031^11. 182. Wasiak, A. (1991) Chemtracts, Macromolecular
Chemistry, 2, 211-41. 183. Douillard, A. (1990) Compte rendu hebdomadaire
des seances de I’Acadimie des sciences, Paris, Series II, 311, 1405-10. 184.
Lauritzen, J. J. and Hoffman, J. D. (1960) Journal of Research of the National
Bureau of Standards, A64, 73-102. 185. Hoffman, J. D. and Weeks, J. J. (1962)
Journal of Research of the National Bureau of Standards, 66A, 13—28. 186.
Kamide, K., Ohno, K. and Kawai, T. (1970) Makromolekulare Chemie, 137,1-7.
187. Khanna, Y. P. and Kumar R. (1989) Journal of Polymer Science, Polymer
187. Khanna, Y. P. and Kumar R. (1989) Journal of Polymer Science, Polymer
Physics Edition, 27, 369-79. 188. Cheng, S. Z. D., Janimak, J. J. and Zhang, A.
(1991) Polymer, 32, 648-55. 189. Ke, B. (1960) Journal of Polymer Science, 42,
15-23. 190. Wilkinson, R. W. and Dole, M. (1962) Journal of Polymer Science,
58, 1089-1106. 191. Wyckoff, H. W. (1962) Journal of Polymer Science, 62, 83-
114. 192. Passaglia, E. and Kevorkian, Η. K. (1963) Journal of Applied Physics,
34, 90-7. 193. Kriegbaum, W. R. and Uematsu, J. (1965) Journal of Polymer
Science, A3, 767-76. 194. Cox, W. W. and Duswalt, A. A. (1967) Polymer
Engineering Science, 7, 309-16. 195. Danusso, F. and Gianotti, G. (1968)
European Polymer Journal, 4, 165-70. 196. Kamide, K. and Nakamura, K.
(1968) Sen-i Gakkaishi, 24, 486-91. 197. Kamide, K. and Toyama, K. (1969)
Sen-i Gakkaishi, 25, 49-51. 198. Gee, D. R. and Melia, T. P. (1970)
Makromolekulare Chemie, 132, 195-201. 199. Fatou, J. G. (1971) European
Polymer Journal, 7, 1057-64. 200. Samuels, R. J. (1975) Journal of Polymer
Science, Polymer Physics Edition, 13, 1417-46. 201. Mucha, M. (1981) Journal
of Polymer Science, Polymer Symposia, 69, 79-89. 202. Miller, R. L. and
Seeley, E. G. (1982) Journal of Polymer Science, Polymer Physics Edition, 20,
2297-307. 203. Bu, H. S., Cheng, S. Z. D. and Wunderlich, B. (1988)
Makromolekulare Chemie, Rapid Communications, 9, 76-7. 204. Kamide, K.
and Yamaguchi, K. (1972) Makromolekulare Chemie, 162, 219-33.

114 Crystallization, melting and supermolecular structure 205. Yadav, Y. S. and


Jain, P. C. (1986) Polymer, 27, 721-7. 206. Corradini, P., Napolitano, R., Oliva,
L. et al. (1982) Makromolekulare Chemie, Rapid Communications, 3, 753-6.
207. De Rosa, C., Guerra, G., Napolitano, V. et al. (1984) European Polymer
Journal, 20, 937-41. 208. Petraccone, V., De Rosa, C., Guerra, G. and Tuzi, A.
(1984) Makromolekulare Chemie, Rapid Communications, 5, 631-4. 209. Guera,
C., Pertaconne, V., Corradini, P. et al. (1984) Journal of Polymer Science,
Polymer Physics Edition, 22, 1029-39. 210. Petraccone, V., Guerra, G., De Rosa,
C. and Tuzi, A. (1985) Macromolecules, 18, 813-14. 211. Guerra, G., De Rosa,
C., Petraconne, V. and Tuzi, A. (1985) Journal of Thermal Analysis, 30, 1337-
42. 212. De Rosa, C., Guerra, G., Napolitano, R. et al. (1985) Journal of Thermal
Analysis, 30, 1331-6. 213. De Rosa, C., Guerra, G., Petraccone, V. and Tuzi A.
(1987) Polymer Com¬ munications, 28, 143. 214. Pae, K. D. and Sauer, J. A.
(1968) Journal of Applied Polymer Science, 12, 1901-19. 215. Hikosaka, M. and
Seto, T. (1972) Polymer Journal, 5, 111-127. 216. Fox, D., Labes, Μ. M. and
Weissberger, A. (1965) Physics and Chemistry of Organic Solid State, Vol. 1,
Wiley-Interscience, New York, Chapter 8. 217. Vaughan, A. S. and Bassett, D.
C. (1989) Crystallization and morphology, in Comprehensive Polymer Science,
Vol. 1. (eds G. Allen and J. C. Bevington), Pergamon Press, Oxford, Chapter 12.
Vol. 1. (eds G. Allen and J. C. Bevington), Pergamon Press, Oxford, Chapter 12.
218. Haudin, J. M. (1986) Optical studies of polymer morphology, in Optical
Properties of Polymers (ed. G. H. Meeten), Elsevier, London, Chapter 4. 219.
Keith, H. D. and Padden, F. J. (1963) Journal of Applied Physics, 34, 2400-
2419. 220. Keith, H. D. and Padden, F. J. (1986) Polymer, 27, 1463-71. 221.
Bassett, D. C. and Vaughan, A. S. (1986) Polymer, 27, 1472-6. 222. Point, J. J.
(1955) Bulletin de VAcadimie Royale de Belgique, Science, Cl, 974-81. 223.
Lednicky, F. (1984) Polymer Bulletin, 11, 579-84. 224. Lednicky, F. (1987)
Makromolekulare Chemie, 188 619-27. 225. Piorkowska, E. and Galeski, A.
(1985) Journal of Polymer Science, Polymer Physics Edition, 23, 1723-48. 226.
Schulze, G. E. W. and Willers, R. (1987) Journal of Polymer Science, Polymer
Physics Edition, 25, 1311-24. 227. Idrissi, Μ. B. O., Chabert, B. and Guillet, J.
(1985) Makromolekulare Chemie, 186 881—92 228. Awaya, H. (1988)
Polymer, 29, 591-6. 229. Idrissi, Μ. B. O., Chabert, B. and Guillet, J. (1986)
Makromolekulare Chemie, 187, 2001-10. 230. Olley, R. H. and Bassett, D. C.
(1989) Polymer, 30, 399-409. 231. Prentice, P. (1982) Polymer, 23, 1189-92.
232. Schulze, G. E. W. and Wilbert, Η. P. (1989) Journal of Materials Science,
8, 71-4. 233. Bassett, D. C. and Olley, R. H. (1984) Polymer, 25, 935-43. 234.
Padden, F. J. and Keith, H. D. (1973) Journal of Applied Physics, 44, 1217-23.
235. Khoury, F. J. (1966) Journal of Research of the National Bureau of
Standards, 470, 29-40.

References 115 236. Sauer, J. A., Morrow, D. R. and Richarson, G. C. (1965)


Journal of Applied Physics, 30, 3017-21. 237. Lotz, B. and Wittmann, J. C.
(1986) Journal of Polymer Science, Polymer Physics Edition, 24, 1541-58. 238.
Wittmann, J. C. and Lotz, B. (1990) Progress in Polymer Science, 15, 909-48.
239. Lotz, B., Wittmann, J. C., Stocker, W. et al. (1990) Polymer Bulletin, 26,
209-14. 240. Peterlin, A. (1975) Molecular aspects of oriented polymers, in
Structure and Properties of Oriented Polymers (ed. J. M. Ward), Elsevier,
London, pp. 46-48. 241. Keller, A. (1977) Routes to high modulus by
ultraorientation of flexible molecules, in Ultra-High Modulus Polymers (eds A.
Cifferi and J. M. Ward), Elsevier, London, Chapter 11. 242. Kargin, V. A. and
Andrianova, G. P. (1962) Doklady Akademii Nauk SSSR, 146, 1337-9 (in
Russian). 243. Binsbergen, F. L. (1966) Nature (London), 211, 516-17. 244.
Gray, D. G. (1974) Polymer Letters, 12, 645-50. 245. Thomason, J. L. and Van
Rooyen, A. A. (1992) Journal of Materials Science, 27, 897-907. 246. Jenckel,
E., Teege, E. and Hinrichs, W. (1952) Kolloid Zeitschrift, 129, 19-24. 247.
Fujiyama, M. and Wakino, T. (1991) Journal of Applied Polymer Science, 42, 9-
20. 248. Xavier, S. F. and Sharma, Y. N. (1984) Angewandte Makromolekulare
Chemie, 127, 145-52. 249. Fitchmun, D. R. and Newman, S. (1969) Polymer
Chemie, 127, 145-52. 249. Fitchmun, D. R. and Newman, S. (1969) Polymer
Letters, 7, 301-5. 250. Fitchmun, D. R. and Newman, S. (1970) Journal of
Polymer Science, Part A2, 8, 1545-64. 251. Kantz, M. R. and Cornelliussen, R.
D. (1973) Polymer Letters, 11, 279-84. 252. Chatterjee, A. M., Price, F. P. and
Newman, S. (1975) Journal of Polymer Science, Polymer Physics Edition, 13,
2391-400. 253. Campbell, D. and Qayyum, Μ. M. (1980) Journal of Polymer
Science, Polymer Physics Edition, 18, 83-93. 254. Folkes, M. J. and Hardwick,
S. T. (1987) Journal of Materials Science Letters, 6 656-658. 255. Huson, M. G.
and McGill, W. J. (1984) Journal of Polymer Science, Polymer Chemistry
Edition, 22, 2571-80. 256. Thomason, J. L. and Van Rooyen, A. A. (1992)
Journal of Materials Science, 27, 889-96. 257. Avella, M., Dellavol, G.,
Martuscelli, E. and Raimo, M. (1992) Polymer Engineering Science, 32, 376-82.
258. Avella, M., Martuscelli, E., Pascucci, B. and Raimo, M. (1992) Polymer
Engineering Science, 32, 383-91. 259. Gray, D. G. (1974) Polymer Letters, 12,
509-15. 260. Devaux, E. and Chabert, B. (1990) Polymer Communications, 31,
391-4. 261. Vargor, J. and Karger-Kocsis, J. (1994) Journal of Materials Science
Letters, 13. 262. Fox, D., Labes, Μ. M. and Weissenberger, A. (1965) Physics
and Chemistry of the Organic Solid State, Vol. II, Wiley-Interscience, New
York, Chapter 6. 263. Keller, A., Hikosaka, M., Rastogi, S., Toda, A., Barham,
P. J. and Goldbeck- Wood, G. (1994) Journal of Materials Science, 29, 2579-
2604.

4 Nucleation of polypropylene A. Galeski 4.1 NUCLEATION IN POLYMERS


The classical concept of crystal nucleation based on the assumption that
fluctuations in the undercooled phase can overcome the energy barrier at the
surface of the crystal was first developed by Gibbs and modified by Kossel and
Volmer (see general surveys of nucleation [1] and for macro¬ molecules [2]).
The rate of nucleation /* has been derived as [3]: I* = (NkT/h) exp[ —(AG* +
AG„)//cT] (1) where N is related to the number of crystallizable elements, AG*
is the energy of formation of a nucleus of critical size and AG^ is the activation
energy for chain transport. The formula derivation is based on the above
assumptions using the absolute rate theory. Generally, in polymers as the
temperature is lowered from the melting temperature a rapid decrease in AG*
and a slow increase in AG,, occur, causing /* to increase. As the temperature is
lowered even further, the decrease in AG* becomes moderate but the increase in
AG,, more significant, resulting in a decrease in /*. Therefore, a maximum in /*
exists, which is related to the ease with which crystallizable elements can cross
the phase boundary. The theories of polymer crystallization are still very
controversial. The behavior of polymer melts on the molecular level at melting
temperature and below it is not yet fully understood and creates problems with
interpreting crystallization on higher levels. The basic problem lies in the
interpreting crystallization on higher levels. The basic problem lies in the
quantification of processes that are only qualitatively understood. For this reason
one will find in the literature a variety of expressions for a given parameter for
polymer nucleation and crystallization. Although the Polypropylene: Structure,
blends and composites. Edited by J. Karger-Kocsis. Published in 1995 by
Chapman & Hall, London. ISBN 0 412 58430 1

Nucleation in polymers 117 expressions may differ only slightly one has a
choice in selecting an expression to use for one’s own data. The calculations
may result in a scatter of 10-20% or even greater from the published data. Since
the smallest value of AG* is related to the size and shape of the nucleus in such
a way that it has the minimum surface free energy, the critical dimensions of the
nucleus can be calculated for the anti¬ cipated geometry of the nucleus. The
expressions for critical sizes of nucleus can be obtained by zeroing the first
derivative of AG* (the sum of changes in bulk and surface energies of the
nucleus) with respect to the dimensions of the nucleus. Similarly, several
expressions can be found for the free energy of fusion Ag{. Together with
various geometries of the nucleus, this gives rise to a range of equations that
relate important parameters. Hoffman and colleagues [4-7] presented extensive
work in this area, and the expressions derived or used by them will be quoted in
this review. For example, for the case of homogeneous primary nuclea¬ tion and
for rectangular shape of the nucleus the free enthalpy of the formation of a
nucleus of critical size can be described by the following expression (see Ref. 4):
AG* = (32<7<xe)/(A0f)2 = [32σ2σβ(Τ^°)2]/[(Δ/ιΓ)2/2(ΔΤ)2] (2) where σ and
<xe are the side and end surface free energies of the crystal, Ag{ is the free
enthalpy of fusion of the crystal of the chosen geometry, Tm° is the equilibrium
melting temperature, AT=Tm°—T, and/ = 27y(T+Tm°). The term AG,, is
usually approximated by the William-Landels-Ferry (WLF) equation for the
viscous flow: A G,/fcT=G*/[R(T-T00)] (3) Based upon these types of
calculations, one will find that the typical homogeneous nucleus dimensions are
about 102 nm3, while a typical polymer chain volume is between 102 and
104nm3. Thus, only a small portion of the polymer chain is involved in forming
a nucleus. One of the two types of nuclei - fringed micelle - is thought to be a
bundle of polymer chains with long sections remaining uncrystallized. There are
restrictions to the formation of fringed micelle. As shown by Flory [8], the strain
generated at the crystalline-amorphous interface by polymer molecules that cross
the interphase boundary must limit the nucleus and crystal dimensions. A single
molecule, therefore, must fold in order to reach the proper dimensions for the
formation of a nucleus and further growth. Thus, the crystal continues to grow in
chain folded fashion with constant lamellar thickness. Chain folded nuclei are
chain folded fashion with constant lamellar thickness. Chain folded nuclei are
more probable than fringed micelle nuclei in all cases where the segmental
mobility of the macro- molecular chain is high. Another path for primary
nucleation is the heterogeneous nucleation, studied extensively by Binsbergen
and de Lange [9] and Binsbergen

118 Nucleation of polypropylene [10-14]. The experimental part of these works


is concerned mostly with the heterogenous nucleation in isotactic polypropylene.
On the basis of a vast number of his own and other authors’ experimental
observations, Binsber- gen derived a theory of heterogeneous nucleation of
crystallization in polymers [14]. The formation of a nucleus on a foreign surface
involves creation of a new interface, similarly to the case of homogeneous
nucleation. However, the pre-existing foreign surface greatly reduces the free
enthalpy of the formation of a critical nucleus, AG*. This lowers the critical size
of the nucleus and results in the formation of heterogeneous nuclei at lower
undercooling. Again, assuming rectangular shape of a nucleus lying flat on a
foreign surface, one can obtain the expression for the free enthalpy of the
formation of a nucleus of the critical size: AG* =(16Δσσσε)/(Δ^)2 = [16
Δσσσ6(Τ„°)2]/(Δ^/ΔΓ)2 (4) where Ασ is the specific interfacial free energy
difference for the nucleus- foreign surface interface. As for homogeneous
nucleation, the Equation 4 is proportional to 1/(ΔΤ)2. However, for very active
foreign surfaces charac¬ terized by low values of Δσ, the critical thickness of a
nucleus approaches the molecular thickness, and Equation 4 transforms to: AG*
= (4iWe Tm°)/(Ah{fAT—AaTm°/b0) (5) where b0 is the molecular thickness.
As Δσ tends to zero Equation 5 approaches the enthalpy barrier characteristic for
the secondary nucleation mechanism which is proportional to 1 /AT. The kinetic
nucleation theory with chain folding provides now the best general tool for
understanding the primary nucleation and the growth of polymer crystals at
isothermal conditions from unstrained melt [4]. The reptation concept proposed
originally by de Gennes [15] was also adapted for the description of chain
motion and transport in the melt [5, 16, 17]. The reptation theory leads to more
accurate expressions for the pre¬ exponential factor in I* and for the activation
energy for chain transport, AGni in Equation 1. It also predicts the dependence
of the crystallization rate on molecular weight in different regimes of
crystallization [6, 7]. For example, the appropriate expression for the secondary
nucleation process predicted by the kinetic nucleation theory with reptation is
[5]: I=(N0pt Pi)/(a0ns) exp[-4b0ffaeTm7(ATA/if/cT)] (6) where N0 is the
number of reacting species at the growth front, /?g=(fc/n)(/c7y/z)exp[ —
Qxy/RT\ n is the number of macromolecule sege- ments in the melt, ns is the
number of stems of width a0, h is the Planck constant and Qd is the activation
number of stems of width a0, h is the Planck constant and Qd is the activation
energy for reptation. κ is a constant, usually of the order of unity as determined
from the experiments. The most spectacular prediction of the reptation concept
concerns the mean time of reeling out from the melt an entire macromolecule
with one

Nucleation in polymers 119 end attached to the nucleus on the growing front and
pulled by crystalliza¬ tion forces [6]: i=1.9xlO"V[s] (7) (n is the number of
chain units, other required parameters taken for polyethylene), which is very
short (order of 10“2 s). For comparison, the time for establishing intermolecular
entanglements in a polymer melt is approximately four orders of magnitude
longer [18] (intermolecular en¬ tanglements were removed by dissolution in a
solvent followed by quick evaporation of a solvent and careful drying). The time
for restoring the intermolecular entanglements is of the order of tens of minutes
for poly¬ propylene melt. These two results show that macromolecules in the
melt are relatively immobile in contrast to a crystallizing macromolecule pulled
at one end by attractive forces of crystallization. It is also clear in view of these
results that the nucleation of new crystalline layers on an existing crystal is the
controlling factor of the crystal growth. Direct evidence for a further nucleation
step beyond primary nucleation was demonstrated by Wunderlich and Cormier
[19] from obser¬ vation of crystallization of linear polyethylene melt seeded
with extended chain crystals. The observable crystal growth is a result of two
processes: the nucleation of initiating stems on the surface of the crystal, and the
coverage of the surface by new stems beginning at the initial stem. It must be
emphasized here that the crystal grows macroscopically in the direction normal
to its surface while on the molecular level the elementary growth mechanism is
the growth along the crystal surface. 4.2 PRIMARY SPHERULITE
NUCLEATION IN POLYPROPYLENE 4.2.1 General remarks on primary
nucleation From the expressions for AG* for homogeneous (Equation 2) and
hetero¬ geneous (Equations 4 and 5) nucleation, a constant nucleation rate in
isothermal conditions is expected. However, in polymer samples there are
usually heterogeneous seeds with a broad spectrum of Δσ values, resulting in
various nucleation rates /*. Also, a limited number of those seeds present in
samples leads to differentiated exhaustion of particular fraction of nuclei.
Moreover, the self-seeding gives rise to almost instantaneous nucleation. All of
these attributes of nucleation events mean that the real nucleation process in
polymers, and particularly in polypropylene, is a complicated function of time,
not just of temperature: J* = /*(T(i), t). The theories of homogeneous,
heterogeneous and self-seeded nucleation describe the mechanisms and show the
tendency but do not predict the real habit of nucleation in polymers. Hence,
tendency but do not predict the real habit of nucleation in polymers. Hence,
experimental methods of determining nucleation are of particular importance.
Knowledge of nucleation data is often essential for

120 Nucleation of polypropylene controlling the physical properties of polymer;


mechanical properties de¬ pend to a great extend on the average size of
spherulites, size distribution and the size of so-called ‘weak spots’ - defects of
spherulitic structure including cavities and frozen stresses which result from
volume contraction during crystallization [20, 21] - all of which are determined
by the primary nucleation process. For some applications it is sufficient to
determine only the total number of nuclei activated during the crystallization.
The simplest way of obtain¬ ing this value is from the average spherulite size for
samples that are filled with spherulites. The average spherulite size can be
obtained from first moment of the size distribution or of the distribution of chord
inter¬ cepts with spherulite boundaries [22]. Other average spherulite sizes can
be obtained from higher moments of spherulite size distribution. The higher
moments of spherulite size distribution can be determined on the basis of direct
characterization of spherulite patterns (under polarized, scann¬ ing or
transmission electron microscopy of thin films or sections of bulk samples,
second or third moment), on the basis of the small angle light scattering (fourth
or fifth moment of the spherulite size distribu¬ tion) [23-25] or using the light
depolarization technique (second or third moment) [26, 27]. However, if the time
dependence of activation of nuclei is required, other methods must be used. The
data on time distribution of primary nucleation are usually obtained by direct
microscopic observation of a crystallizing sample. The drawbacks of this method
are the necessity of using thin samples and the need for crystallization conditions
that allow for counting the spherulite centers. These limitations can be overcome
by applying a method of reconstruct¬ ing the sequence of nucleation events from
the shapes of spherulite bound¬ aries in previously crystallized films [28] and in
thin sections for bulk samples [29]. The time lag between the nucleation of two
neighboring spherulites can be found from the curvature of their common
boundary, and if this procedure is repeated for the chain of neighboring
spherulites data on the time distribution of the activation of nuclei can be found.
The time distribution of activation of nuclei should be expressed as the number
of activated nuclei per volume unit of untransformed fraction of the sample.
Calibration of the time axis in this method is made by measurement of the
spherulite growth rate. Usually, a given brand of polypropylene is characterized
by an average number of primary nuclei per volume unit at certain crystallization
condi¬ tions. The average spherulite size in bulk is determined by the number of
nuclei per volume unit. For thin films, however, the spherulites as seen in plan
are larger. For a sample below the average spherulite size in bulk, the thinner the
are larger. For a sample below the average spherulite size in bulk, the thinner the
sample, the larger the spherulites. This apparent increase in spherulite sizes in
thin films results from the constant average number of

Primary spherulite nucleation 121 nuclei per volume unit. The factor
complicating this simple relation is the nucleating ability of sample surfaces. The
Avrami type of analysis is often erroneously applied to obtain nuclea¬ tion data
from differential scanning calorimetry (DSC) using isothermal crystallization
conditions and from dilatometry. The reason for this is that the conversion of
melt to spherulites is assumed to follow pure sporadic or pure instantaneous
modes, the only modes described correctly by the Avrami equation: α(ί)=1—
exp(—Kin) (8) where a(i) is the degree of the conversion of melt to spherulites.
However, the general form of the equation for the conversion kinetics of melt to
spherulites for isothermal experiments, which was first developed by Av¬ rami,
is as follows: α(ί)=1— exp[ — nG2 /(τ) (t — τ)2 dr] for films (9a) a(t) = 1 - exp[
- (4/3)nG31/(τ)(ί - τ)3 dr] for bulk samples (9b) 0 where G is the spherulite
growth rate constant at a given temperature of crystallization and /(τ) is the rate
of nucleation. It is evident from Equation 9a and b that for I* = I0S(t) (<5(i)
being the delta Dirac function, representing the instantaneous mode of
nucleation) one obtains the Avrami exponent, t2 and i3 for two- and three-
dimensional cases, respectively, while for I* = I0tc (c being a constant, for
sporadic nucleation c = 0) the Avrami exponent is proportional to ic+4 and tc+4
for the two- and three- dimensional cases, respectively. Therefore the plot of ln[
—ln(a(f))] against ln(i), as required for the Avrami type of analysis, can lead to a
straight line only in two limiting cases, instantaneous and sporadic nucleation
modes, both of which occur very seldom in the crystallization of polymers. The
proper way of analysing the data is by solving the integral equation which
follows from Equation 9a and b: t /* /(£)(£—τ)2 dr = — l/(nG2) ln[l —α(ί)] for
films (10a) % 0 t j*/(t)(£—τ)3 dr = — 3/(4nG3)ln[l — a(t)] for bulk samples o
(10a)

122 Nucleation of polypropylene The simplest way of solving Equation 10a and
b is by Laplace trans¬ formation, or equivalent, by third (for films) or forth (for
bulk) order differentiation against time, t: /(£) = — l/(2nG2) d3[ln(l — a(i))]/d£3
for film (11a) /(£) = —1/(8nG3)d4[ln(l — a(£))]/d£4 for bulk samples (lib) In
the literature there are many examples of the treatment of the problem of
nonisothermal solidification (for a broad review see [30]. Most of them are
lacking in a firm theoretical background. However, the probabilistic approach to
the description of spherulite patterns [31, 32] now provides a convenient tool for
the description of the coversion of melt to spherulites. In the case of
the description of the coversion of melt to spherulites. In the case of
nonisothermal crystallization the conversion of melt to spherulites is described
by [31-33]: <x(t)= 1 — exp t * — π Ι(τ) 0 t jc(s)ds * for films (12a) α(ί)= 1 -
exp< — (4π/3) /(τ) Η ΐ 0 τ G(s) d s άτ for bulk samples (12b) Since the growth
rate is unambiguously determined in all three regimes of crystallization by the
secondary nucleation process and the completion rate of the nucleated layer, it
could be measured precisely in isothermal experi¬ ments in thin films as a
function of temperature and could then be easily transformed to the function of
time, provided that the temperature change is monitored during nonisothermal
solidification of a polymer. The solution of Equation 12a and b is also by
Laplace transformation, or equivalent, by differentiation [33]: /(£)=-1/(2π)^
/(£)=-1/(8π)^ for films l/G(i)^[l/G(i)^{ln[l-«(£)]} (13a) for bulk samples (13a) In
all of the above equations the conversion degree of melt to spherulites must not
be mistaken for the crystallinity degree. The difficulty in all nonisothermal
experiments is that the fractions of the material crystallized at different
temperatures differ in the degree of crystallinity. It is usually considered that
nuclei are spread randomly over the sample, except for nuclei formed on outer
surfaces in three-dimensional samples, on

Primary spherulite nucleation 123 surfaces allowing for transcrystallinity and on


surfaces of the second dispersed component, e.g. short fibers. However, that is
not necessarily true. If the nucleation events occur not only at the very beginning
of crystalliza¬ tion but also during crystallization then the volume occupied by
already advanced spherulites is excluded from further nucleation. The close
vicinity of an arbitrarily chosen nucleus contains fewer nuclei than more distant
regions. Although the nucleation itself is a spatially random process it is limited
to the uncrystallized portion of the sample. Such an exclusion always produces a
kind of a distance correlation, if the nucleation process is prolonged in time. The
spatial correlation of spherulite centers was first observed by Misra,
Prud’homme and Stein [34], while mathematical for¬ mulas for the description
of the distance correlation of nuclei for model modes of primary nucleation were
derived in [32] and [33] for the isothermal and nonisothermal cases,
respectively. Polypropylene has been shown to exhibit several crystalline
modifications [35-38] in addition to the most common a monoclinic structure
reported first by Natta and Corradino [36]. The β-phase crystallizes from
primary nuclei in a hexagonal fashion, although, it can also be initiated along a
growing front of the α-phase in the temperature gradient. The y triclinic phase
was first discovered in low molecular fractions of isotactic polypropy¬ lene
solidified by slow cooling [37], in commercial polypropylene crystal¬ lized
under high pressure [38] and in samples of propylene-ethylene copolymer at a
under high pressure [38] and in samples of propylene-ethylene copolymer at a
low ethylene content, which exhibited the peculiarity of complete crystallization
in the y form [39, 40]. There are no reports on primary nucleation of the y form
spherulites in the literature, however, the y form is reported to exhibit spherulitic
lamellar structure and undergoes the γ-a transition on annealing at 147 °C at
atmospheric pressure [41]. 4.2.2 Nucleation of the a form of isotactic
polypropylene The first consistent microscopic data on the primary nucleation of
isotactic polypropylene were obtained by von Falkai and Stuart [42] and von
Falkai [43]. Their data for samples melt annealed at 180 °C prior to
crystallization and then crystallized isothermally in the temperature range from
122 to 145 °C are presented in Table 4.1. The nucleation in these experiments in
polypropylene was found to be heterogeneous and instantaneous with calculated
Avrami exponent close to 3. (There is some doubt about the three-dimensionality
of their samples for microscopic observation of crystal¬ lization). Since then,
many researchers have published nucleation and crystallization data for isotactic
polypropylene [e.g. 26, 27, 44—48, see also 49]. The change from instantaneous
to sporadic primary nucleation is re¬ ported in the literature for samples heated
to 200 °C and above [44, 47]. Under such conditions, polarized light microscopy
examination revealed an

124 Nucleation of polypropylene Table 4.1 Basic crystallization parameters of


isotactic polypropy¬ lene of Mw = 51200, heptane extracted, melt annealed
before crystallization at 180 °C for 15 min [42] Crystallization temperature (°C)
Nucleation density (106/cm3) Growth rate (μπι/ιηίη) 122.0 2.36 18.0 125.0 1.56
12.0 127.5 1.02 7.0 130.0 0.85 4.3 132.5 0.73 2.6 135.0 0.65 1.6 138.0 0.58 0.86
140.0 0.53 0.59 145.0 0.47 0.27 initial constant nucleation rate, decreasing for
longer crystallization time. The Avrami type of fit to the integral exponent was
not satisfactory in this case. The course of primary nucleation in isotactic
polypropylene down to 70 °C was first demonstrated by Burns and Turnbull [50]
and by Koutsky, Walton and Baer [51], employing the droplet technique
originally develop¬ ed by Vonnegut for tin and water droplets [52]. Four distinct
regions can be recognized in the nucleation of isotactic polypropylene melt: 1.
Immediately below the DSC determined melting point (165-167 °C) there is a
gap where the crystal nucleation and growth hardly occurs. Neither the present
heterogeneities nor introduced nucleating agent can accelerate the nucleation. 2.
Most of the published nucleation data concern the region of tempera¬ ture below
150 °C but above 115 °C, where regular spherulites are nucleated [53, 54]
(although Binsbergen and de Lange [55] observed negatively birefringent
sheaves of crystalline lamellae crystallized iso- thermally at as high as 160 °C
which is well above the estimated regime II-regime I transition temperature for
isotactic polypropylene at 155 °C [54]). This region is the extended region of
isotactic polypropylene at 155 °C [54]). This region is the extended region of
activity of hetero¬ geneous nuclei. The number of heterogeneous nuclei is
limited during the crystallization. 3. Some of the heterogeneous nuclei become
active at even lower tempera¬ ture, which follows from their smaller size or
lower perfection. These nuclei are also limited in number. 4. Finally at around
80-85 °C and below there is the region of homogene¬ ous nucleation. The
number of nuclei in this region increases rapidly with the decrease of
temperature.

Primary spherulite nucleation 125 Except for very thin specimens, because of the
low thermal diffusivity of polymers, intense nucleation and fast spherulite
growth at lower tempera¬ ture, it is difficult to reach beyond the upper range of
activity of heterogene¬ ous nuclei. In addition, the latent heat of fusion liberated
by the rapid crystallization during quenching tends to maintain the temperature
during crystallization in the upper range of activity of heterogeneous nuclei. This
and the instantaneous character of most heterogeneous nuclei mean that the
homogeneous nucleation range is rarely reached, and many polymeric objects in
technological applications crystallize only from heterogeneous nuclei. Figure 4.1
illustrates the nucleation activity in isotactic polypropylene (RAPRA, iPPl; Mw
= 3.07 x 105 g/mol, Mn = 1.56 x 104 g/mol, density = 0.906 g/cm3, melt flow
index = 0.39 g/min). The data were taken from [56- 58] and differentiated to
represent the contribution of new nuclei activated by the temperature decrease
by 1 °C (data for crystallization temperatures Figure 4.1 Nucleation activity in
isotactic polypropylene (by RAPRA iPPl). The number of nuclei are
differentiated to represent the contribution of new nuclei activated by the
temperature decrease by 1°C. (O) Data obtained in isothermal crystallization on
a microscopic hot stage; (·) nucleation data in isothermal crystallization in a
specially designed crystallization cell. of 90, 100 and 110°C were obtained by
the author for the purpose of this review, using the method of crystallization
described in [58]. The samples in the form of thin films (20-30 pm) were
crystallized isothermally on a microscopic hot stage for temperatures above 115
°C, while for temperatures below 115 °C they were obtained by isothermal
crystallization in a specially

126 Nucleation of polypropylene designed crystallization cell, enabling


isothermal conditions to be reached within the sample volume in less than 0.5 s.
It can be seen that the number of nuclei increases initially as the temperature of
crystallization decreases. At the temperature of crystallization of 132 °C a
change of slope of the Δ//ΔΤversus Tcurve is seen, which is apparently
associated with the regime II-regime III transition in crystal growth kinetics
associated with the regime II-regime III transition in crystal growth kinetics
(reported to be at 135-137 °C for other brands of isotactic polypropylene, as
determined in [7] on the basis of data from [9, 42 59], see also [54]). The regime
II-regime III transition is in fact the change in the intensity and the habit of
secondary nucleation, which may be considered as heterogeneous nucleation on
the polymer crystal surface; similar, though not identical, transition (slightly
different transition temperature) should be expected for nucleation on surfaces of
other heterogeneous nuclei. Further reduction of the crystallization temperature
below 115 °C results in saturation of the Δ//ΔΤ value. Apparently all
heterogeneous nuclei present in the sample are able to show up within the time
of crystallization below 115 °C. At temperatures below 85 °C a new intense
process of homogeneous nucleation takes place. A rapid increase in the number
of nuclei formed with the decrease of crystallization temperature is observed
(see Figure 4.1). Early droplet experiments during isothermal crystallization [50,
51] also showed that the nucleation in droplets of isotactic polypropylene is
thermally activated and that the droplets crystallize sporadically in time. Inves¬
tigations of nucleation performed during continuous cooling could, how¬ ever,
resolve only a single large peak of nucleation at one particular undercooling.
Annealing of the melt has a great influence on primary nucleation in isotactic
polypropylene. However, knowledge of the behavior of primary nucleation
during melt annealing in polypropylene was acquired gradually as understanding
of nucleation processes in polymers became better. The first extensive studies of
the effects of thermal history on crystallization of isotactic polypropylene were
conducted by Pae and Sauer [60] and Sauer and Pae [61]. Further studies
included direct microscopic observations of the formation of spherulites in
polypropylene melt subjected to various thermal treatment. Annealing of
polypropylene melt prior to crystallization decreases the active fraction of
primary nuclei. The crucial factor is the temperature of melt annealing - below or
above the equilibrium melting temperature, T^. At 190 °C a vast number of
those thermally sensitive nuclei remains untouched while even short exposure to
temperature around 220 °C decreases the number of active nuclei by orders of
magnitude. (There is still some controversy about the proper value of the
equilibrium melting temperature for the α-phase of polypropylene which follows
from various extrapolations of melting data; if one applies a higher
crystallization temperature range for extrapolation the equilibrium melting
temperature is

Primary spherulite nucleation 127 found at 208 °C [62], 220 °C [63] and
208±8°C [54]. Mucha [64] found 212 °C independent of the crystallization
temperature range from the extrapolation of small angle X-ray scattering
temperature range from the extrapolation of small angle X-ray scattering
(SAXS) data on lamellae thickness.) The nuclei disappearing due to annealing
are thermally activated heterogeneous nuclei and so-called self-nuclei [65]. Self-
nucleation is a general term describing nucleation of a melt or solution by its
own crystals grown previously. Self-nucleation in polymers is particularly strong
because of the large temperature range where crystals do not melt entirely. It has
been suggested that most of the previous obser¬ vations of the behavior of
primary nucleation concerns self-seeding. The equilibrium melting temperature
must be exceeded on melt annealing in order to remove entirely the self-seeded
nuclei from the melt. A technique of self-nucleation was developed [66] based
on the observa¬ tion that the critical nucleus size decreases with decreasing
temperature. After melting the temperature is reduced below the melting point,
preferably 10-15 °C higher than the required crystallization temperature, and
main¬ tained for a period of time to produce embryos. The temperature is then
lowered to the crystallization temperature, at which most of the embryos reach
the critical size and become stable nuclei. Using this technique one can increase
the number of nuclei by a few orders of magnitude and significantly reduce the
crystallization time. The other source of thermally sensitive nuclei are polymer
crystal frag¬ ments in small cracks and cavities of a foreign surface, which
survive melting because of an increase of their melting temperature due to the
stiffening effect. At moderate undercooling they initiate the growth of
spherulites. The structure of early stages of the growth of spherulites in
polypropylene was studied by Bassett and Olley [67-69]. Figure 4.2 shows a cut
and etched surface of an isotactic polypropylene sample crystallized at 155 °C.
The nucleus is built from more or less regular lamellae showing a multilayer
arrangement. Early objects develop by branching, usually at rather large angles,
and splaying apart the dominant lamellae. A sheaf-like centre is usually seen for
crystallization temperatures above 155 °C or a cross- hatched structure when
viewed in plan for crystallization temperatures of 155 °C and below (see Figure
4.2). The change of morphology of spherulite centers at 155 °C is apparently
connected with regime I-regime II transition, expected at around 155 °C. The
multilayer arrangement of the very center of a spherulite is explained by Bassett
[68] as the result of a shish-kebab type of nucleation on straightened fragments
of macromolecules, which can always be found in a polymer melt. This
conclusion is based on electron microscopy observations of nucleation in
isotactic polystyrene. It means that the core of a nucleus is build of a single or a
bundle of elongated macromolecules. The elongated fragment extends as far as
several lamellae thicknesses i.e. 50-100 nm. The volume of such nuclei agrees
with the estimation of the volume of a homogeneous nucleus based on
calculations
calculations

128 Nucleation of polypropylene Figure 4.2 Two views of early growth at 155
°C in isotactic polypropylene. In the top half of the micrograph one object would
appear cross-hatched in plan. In the bottom half a similar object assumes a sheaf-
like shape when vieewed sideways (from [68], courtesy of Elsevier Applied
Science, London). of the critical nucleus size. However, such elongated shape of
the nucleus is inconsistent with the intuitive assumption that the nucleus is
limited to a single lamella thickness which was made in most calculations and
modelling concerning primary nucleation. 4.2.3 Nucleating agents for the α-form
of isotactic polypropylene The crystallization of isotactic polypropylene from
melt could be enhanced in the region of the temperature where heterogeneous
nucleation is observed by adding some extra heterogeneous nuclei. Interest in
such experiments was stimulated by industrial efforts to decrease the size of
spherulites to improve optical and mechanical properties. By adding finely
subdivided foreign material it was shown that solids, liquids and even gas
bubbles are able to nucleate polypropylene spherulites (for a list of older patents,
see [70]). However, Binsbergen [11] found, contrary to some patent claims, that
most inorganic salts and oxides are inactive in nucleating polypropylene.

Primary spherulite nucleation 129 Beck and Ledbetter [71], Beck [72],
Binsbergen and de Lange [9] and Binsbergen [11] tested a large number of
substances for their possible nucleating effect on the crystallization of
polypropylene. The most active nuclei for isotactic polypropylene are sodium t-
butyl benzoate, monohyd- roxylaluminum p-t-butyl benzoate, sodium p-methyl
benzoate, sodium benzoate, colloidal silver, colloidal gold, hydrazones,
aluminum salts of: aromatic and cycloaliphatic acids, aromatic phosphonic acids,
phosphoric acid, phosphorous acid, several salts of Ca2+, Ba2+, Cu2+, Co2+,
Ga3+, In3+, Ti4 + and V4+ with the above mentioned acids, indigo and air
bubbles. There are also other reports on nucleation activity of certain seeds:
indigo [73], talc [74], and certain crystallographic planes of calcite because of
alternating polar-nonpolar rows on some cleavage planes [75]. Good nucleating
agents are insoluble in the polymer or crystallized before crystal¬ lization of
polypropylene. The important feature of a good nucleating agent appears to be
the existence of alternating rows of polar and nonpolar groups at the seed
surface. It should be mentioned that the best cleavage planes of seed crystals are
not necessarily the crystallographic planes exposing alter¬ nating polar-nonpolar
rows, if they exist. The extensive data on nucleating agents for crystallization of
polypropylene do not completely conform to the theory of heterogeneous
nucleation based on the surface free energy consideration (compare Equations 4
and 5). The crystalline lattice matching type of epitaxy is also excluded as the
and 5). The crystalline lattice matching type of epitaxy is also excluded as the
major mechanism because of the large variety of nucleating seeds. For flat
surfaces the nucleation activity of seeds with a finite nucleation rate is expected,
rather than their instantaneous activity as observed experimentally. Self-
nucleation on the surface, in cracks and steps by residual polypropylene crystals
is also excluded because of the constant number of active seeds independent of
melt annealing in the range 175-280 °C. Heterogeneous nucleation on steps on
the seed crystal surface is able to explain the observable behavior if a proper
distribution of step length is assumed and if Ασ in Equations is sufficiently
small. The reason for this may be a good accommodation of polypropylene
crystals on surfaces of seeds. Binsbergen [10] introduced an accommodation
coefficient £(()<£< 1) accounting for reduction of interfacial free energy by
increased epitaxy through lattice matching and other effects. He proposed that
the existence of alternating rows of polar and nonpolar groups on the surfaces
may be the cause of a large value of the accommodation coefficient. The
presence of a large quantity of a filler in polypropylene changes drastically the
crystallization conditions. Some active fillers, such as talc and aluminum oxide,
act like strong nucleating agents while inert fillers such as chalk show little
nucleating ability. Because of strong nucleating activity of active fillers the
crystallization of bulk filled polypropylene occurs at higher temperature during
quenching. Usually the presence of a filler increases the thermal diffusivity of
the composition. It enables inert fillers to reach lower crystallization
temperature.

130 Nucleation of polypropylene Primary nucleation in polymer blends is a


special case in which other phenomena, such as migration of nuclei during
mixing, rejection of the second component, and occlusions, are playing an
important role because of high concentration of the second component. The
primary nucleation behavior of polypropylene in polypropylene-based blends are
reviewed separately in Chapter 2 in Volume 2. 4.2.4 Nucleation of /(-
modification in isotactic polypropylene More highly negative birefringent
spherulites of /(-phase crystallize in a hexagonal unit cell [35], rather than in
monoclinic fashion as for a form [36]. Nucleation of the β form occurs much
more rarely in bulk samples than the predominant a form. Keith and Padden [53]
observed the sporadic formation of spherulites of β during crystallization in the
range from 128 to 132 °C. Since then, the conditions for the formation of β form
have been studied intensively. It was found that the β form is nucleated pre¬
ferentially in the presence of shearing forces [76]. Quenching was used to
produce the β form of polypropylene in larger quantities. In contrast, Shi and
Zhang [77] and Varga, Garzo and Ille [78] reported an interest¬ ing observation
Zhang [77] and Varga, Garzo and Ille [78] reported an interest¬ ing observation
that the amount of a phase could be suppressed by slowing the cooling rate to
below 5°C/min. In this way β form of high purity can be obtained. Lovinger
Chua and Gryte [79] showed that a large amount of β phase can be obtained by
crystallization in a tem¬ perature gradient. Although the primary nucleation of
/(-phase spherulites is extremely rare in this case, the β phase is easily initiated
by the growth transition along the growing front of α-phase spherulites. The
characteristics of β phase formation were studied extensively by Varga and
colleagues [78, 80-85], who found that certain brands of poly¬ propylene, those
having high molecular weight, are more susceptible to /(-crystallization. It was
established that the nucleation of β form is instantaneous, and that the density of
nuclei decrease with increasing temperature of crystallization [85]. Those
observations and some others [e.g. 86-90] showed clearly that the β form
generally occurs at the level of only a few per cent unless certain heterogeneous
nuclei are present [91] or the crystallization occurs in a temperature gradient
[79], or in the presence of shearing forces [92]. Leugering [93] has demonstrated
that a certain quinacridone dye known as permanent red E3B is very effective in
generating spherulites of β form for isotactic polypropylene crystallized below
130 °C. However, its effectiveness depends on nucleant concentration,
dispersion and the cooling rate. Since then a series of other crystalline
substances have been found to nucleate β form in iso tactic polypropylene [94],
including, 2-mercaptobenzimidazole, phenothiazin, triphenodithiazine,
anthracene, phenanthrene (see Table 4.2 for the effectiveness of those nucleants)
and pimelic acid [95].

Secondary isotactic polypropylene 131 Table 4.2 Fraction of the β form in


nucleated isotactic polypropylene samples prepared by powder blending with 0.5
wt% of a nucleant, molten at 205 °C for 10 min, and then crystallized during
slow cooling [92] Nucleant β form (%) 2 — mercaptobenzimidazole 18
phenothiazine 16 triphenodithiazine 65 anthracene 35 phenanthrene 30
permanent red E3B 78 It has been suggested in the literature that observation of
the β form during slow crystallization results from the retardation of one
substage of a multistage process leading finally to the a form [96]. The β-α
transition observed at 145 °C [97] is supposed to be the confirmation of this
concept. 4.3 SECONDARY NUCLEATION IN ISOTACTIC
POLYPROPYLENE After completion of a folded layer on the surface of the
crystal a new surface nucleus must be created for further growth of the crystal.
This is called a secondary nucleation process. The most widely accepted
expression for the secondary nucleation rate is given by Equation 6. The /?g
factor in Equation 6 is the retardation factor because at a large undercooling
factor in Equation 6 is the retardation factor because at a large undercooling
polymers become very viscous and reptation is retarded. The temperature
dependence of the secondary nucleation rate at low and moderately high
undercoolings is determined by AG*, which is proportional to Tm°/(TAT). The
completion rate of the layer nucleated on the surface of the sub¬ strate, g, is
expressed as a difference of attachment and detachment rates of folds to the
nucleus on the surface of the substrate. Figure 4.3 illustrates the formation and
the growth of the surface nucleus as it spreads in the direction parallel to the
surface of the crystal. After the formation of a stable nucleus (in the case
illustrated in Figure 4.3 for embryos with more than four stems i.e. for the
negative free energy of formation) the layer will be completed with new stems
by the attachment-detachment mechanism. The expression for the completion
rate derived by Hoffman, Davies and Lauritzen [4] is: 9 = αοΟ,βι exp[ —
2a0b0at{\ - i//)/kT] (14) where φ is the fraction of the free energy of fusion for
the forward reaction and Q is a factor of the order of unity. As can be seen from
Equation 14, the completion rate is not a strongly dependent function of
temperature.

132 Nucleation of polypropylene o 2 3 4 5 6 NUMBER OF STEMS Figure 4.3


Free energy of the formation of a chain folded surface nucleus by the attachment
- detachment mechanism. Reaction A is the attachment of a new stem while
reactions B and Bi are the detachment of a stem [47]). There is competition
during crystallization between the secondary nu¬ cleation and the completion of
the layer that is nucleated by the secondary nucleation on the substrate. Three
cases can be distinguished: 1. The secondary nucleation process is slow,
allowing for completion of the nucleated layer before the next event of the
secondary nucleation. 2. The secondary nucleation events occur before the
completion of the nucleated layer. 3. The secondary nucleation occurs so often
that it does not allow for the completion of the nucleated layer with new folds.
These cases are the reason for the existence of three regimes of crystalliza¬ tion
and respective changes in polymer morphology. The transitions to different
regimes is possible because the completion rate, g, depends on temperature but
much less so than the nucleation rate, /. The dependence between the secondary
nucleation rate, /, and the observable growth rate, G, is described by a basic
relationship: 1. For regime I with low nucleation rate allowing a rapid
completion of the entire substrate length [98]: 2. For regime II with the
completion rate of the layers, g, allowing for multiple nucleation on the substrate
[99]: G„ = b0{Ig)u2 Gi=b0IL (15a) (15b)

In syndiotactic polypropylene 133 Table 4.3 Basic parameters for the description
of nucleation and crystallization of isotactic polypropylene a form Equilibrium
of nucleation and crystallization of isotactic polypropylene a form Equilibrium
melting point, T^ 208 + 8 °C [54,62] Enthalpy of fusion, Ah{ 209.3 ± 29.9 J/g
[101] Chain conformation 3i helix [36] a form unit cell parameters [36]
Monoclinic, Space group C‘R-C2/c[36] a 0.665 nm b 2.096 nm c 0.650 nm β
99.3° Main growth direction a* of reciprocal lattice [102] do 0.549 nm bo 0.626
nm Regime I—II transition (predicted) 155 °C [54, 103] Regime II—III
transition 137°C [7] L in regime I «0.11 pm [5, 100] L in regime III «2a0 =0.91
nm [5] Fold surface energy, σβ 1.22 x 10“5 J/cm2 [54] Work of chain folding, g
27.61 kJ/mol [7] Lateral surface free energy, σ [5, 54] (9.2-11.5) x 10'7 J/cm2
Activation energy for reptation, 6.276 kJ/mol [5] Too in the expression for
reptation activation energy 231.2 K [5] 13 form Equilibrium melting point, T°
176°C [95] Enthalpy of fusion, Δh{ 177 J/cm3 [95] Chain conformation 3! helix
Unit cell Hexagonal [35] Space group Dt-P3!21 [37] a 1.274 nm c 0.635 nm β
120° Main growth plane (300) 0.636 nm bo 0.551 nm Regime II—III transition
123-129.5 °C [95] Fold surface energy, σβ (4.82-5.52) xlO"6 J/cm2 3. For
regime III for which the crystallization occurs mainly by the intense nucleation
of new stems on the substrate rather than the completion of layers across the
surface of the substrate [7]: Gm=b0IL' (15c)

134 Nucleation of polypropylene where L is the length of a growth strip of


thickness b0i g is the completion rate of a strip and could be identified in terms
of the nucleation process g = a(A — B), A and B being attachment and detach¬
ment rates, respectively, L is the effective substrate length [100] which
corresponds to ~(2 —3)a0. In general, L' is considerably smaller than L. The
data for growth transitions in polypropylene obtained by Clark and Hoffman [7]
and by Monasse and Haudin [54] are collected in Table 4.3. The secondary
nucleation can be easily determined from the measure¬ ments of the spherulite
growth rate based on the knowledge of the basic crystallographic and
thermodynamic parameters characteristic of a given polymer. The necessary data
for the description of the processes involved in the crystallization of isotactic
polypropylene are collected in Table 4.3. 4.4 NUCLEATION IN
SYNDIOTACTIC POLYPROPYLENE Partially syndiotactic polypropylene has
been known since the 1960s. Only recently, due to the development of Zr and Hf
based catalysts [104], relatively pure syndiotactic polypropylene became
available in larger quan¬ tities. This is the reason why the physical properties
and nucleation behavior have been investigated less than those for the isotactic
form. The published data are often uncertain. It is now evident that syndiotactic
polypropylene can crystallize in several crystallographic forms [105-108] in
which the main difference is the chain conformation. The most common form
shows the orthorhombic unit cell with the cell constants «=1.45nm, h = 0.56nm
shows the orthorhombic unit cell with the cell constants «=1.45nm, h = 0.56nm
and c = 0.74nm [105] resulting from packing of uniformly rotated helices of
macromolecules. A new cell symmetry was proposed by Lotz, Lovinger and
Cais [106] in which the orthorhombic unit cell results from packing of opposite-
handed helices. Those two forms are believed to coexist. Other crystallographic
forms are obtained by solid-solid transformation with the help of various
solvents [107]. From the work of Miller and Seeley [109], it is known that
spherulites are formed in syndiotactic polypropylene. However, there are no
published data on the habit of primary nucleation in syndiotactic polypropylene.
The data on secondary nucleation were obtained by Miller and Seeley [107] and
analyzed by Clark and Hoffman [7]. The available data for the description of
secondary nucleation of syndiotactic polypropylene are presented in Table 4.4.
Recently, the influence of syndiotacticity of polypropylene on the ther¬
modynamic properties, structure, morphology, overall crystallization and
melting was extensively studied by Lovinger et al. [115-117] and Rodriguez-
Arnold et al. Γ1181.

Appendix 135 Table 4.4 Basic parameters for the description of nucleation and
crystallization of syndiotactic polypropylene Chain conformation Unit cell
parameters Space group a b c Space group a b c Equilibrium melting point,
Glass transition, Tg Fold surface energy, σβ Work of chain folding, q Enthalpy
of fusion, Ah{ Two fold helix S(2 1 1)2 [110] Orthorhombic [105] C 222! [105]
1.46 nm 0.56 nm 0.74 nm Ibca [106] 1.45 nm 1.12 nm 0.74 nm 151-161 °C
[109, 111, 112] 0°C [112] 4.99 x 106 J/cm2 [7] 24.27 kJ/mol of folds [7] 45-75
J/g [109, 113, 114], 196.6 J/g [112] 4.5 APPENDIX SYMBOLS AND
ABBREVIATIONS b0 DSC 9 G h /* k N (2d RAPRA SAXS t T TO m α(ί) β,
AG* AG, Δ<7 Ah{ σ σ* Width of a stem Molecular thickness Differential
scanning calorimetry Completion rate of the layer nucleated on the crystal
surface Spherulite growth rate Planck’s constant Nucleation rate Boltzmann’s
constant Number of crystallizable elements Activation energy for reptation
Rubber and Plastics Research Association Small angle X-ray scattering Time
Temperature Equilibrium melting temperature Conversion degree of melt to
spherulites Retardation factor due to melt viscosity Free enthalpy of fusion of a
crystal of certain geometry Energy of formation of a nucleus of critical size
Activation energy for chain transport Specific interfacial free energy difference
for interface Enthalpy of fusion Side surface free energy of a crystal End surface
free energy of a crystal

136 4.6 REFERENCES Nucleation of polypropylene 1. Zettlemoyer, A. C.


(1969) Nucleation, Marcel Dekker, New York. 2. Price, F. P. (1969) Nucleation
(1969) Nucleation, Marcel Dekker, New York. 2. Price, F. P. (1969) Nucleation
in polymer crystallization, in Nucleation (ed. A. C. Zettlemoyer), Marcel
Dekker, New York. 3. Turnbull, D. and Fisher, J. C. (1949) Journal of Chemical
Physics, 17, 71-3. 4. Hoffman, J. D., Davies, G. T. and Lauritzen Jr, J. I. (1976)
The rate of crystallization of linear polymers with chain folding, in Treatise on
Solid State Chemistry, Vol. 3 (ed. N. B. Hannay), Plenum Press, New York, pp.
497-614. 5. Hoffman, J. D. (1982) Polymer, 23, 656-70. 6. Hoffman, J. D. and
Miller, R. L. (1988) Macromolecules, 21, 3038-51. 7. Clark, E. J. and Hoffman,
J. D. (1984) Macromolecules, 17, 878-85. 8. Flory, P. (1962) Journal of the
American Chemical Society, 84, 2857-67. 9. Binsbergen, F. L. and de Lange, B.
G. M. (1970) Polymer, 11, 309-32. 10. Binsbergen, F. L. (1970) Kolloid
Zeitschrift und Zeitschrift fur Polymer e, 237, 289-97. 11. Binsbergen, F. L.
(1970) Polymer, 11, 253-67. 12. Binsbergen, F. L. (1970) Kolliod Zeitschrift und
Zeitschrift fur Polymere, 238, 389-95. 13. Binsbergen, F. L. (1972) Journal of
Crystal Growth, 16, 249-58. 14. Binsbergen, F. L. (1973) Journal of Polymer
Science, Polymer Physics Edition, 11, 117-35. 15. de Gennes, P.-G. (1971)
Journal of Chemical Physics, 55, 572-9. 16. Hoffman, J. D., Gutman, C. M. and
DiMarzio, E. A. (1979) Discussions of the Faraday Society, 68, 177-97. 17.
DiMarzio, E. A., Gutman, C. M. and Hoffman, J. D. (1979) Discussions of the
Faraday Society, 68, 210-17. 18. Andrianova, G. P. (1975) Journal of Polymer
Science, Polymer Physics Edition, 13, 95-112. 19. Wunderlich, B. and Cormier,
C. M. (1966) Journal of Physical Chemistry, 70, 1844-9. 20. Galeski, A. and
Piorkowska, E. (1983) Journal of Polymer Science, Polymer Physics Edition, 21,
1299-1312. 21. Galeski, A. and Piorkowska, E. (1983) Journal of Polymer
Science, Polymer Physics Edition, 21, 1313-122. 22. Schultze, G. E. W. and
Willers, R. (1987) Journal of Polymer Science, Polymer Physics Edition, 25,
1311-24. 23. Stein, R. S. and Rhodes, Μ. B. (1960) Journal of Applied Physics,
31, 1873-84. 24. Clough, S., van Aartsen, J. J. and Stein, R. S. (1965) Journal of
Applied Physics, 36, 3072-85. 25. Champion, J. V., Killey, A. and Meeten, G. H.
(1985) Journal of Polymer Science, Polymer Physics Edition, 23, 1467-76. 26.
Magill, J. H. (1961) Nature, 191, 1092-3. 27. Magill, J. H. (1962) Polymer, 3,
35-42. 28. Pakula, T., Galeski, A., Kryszewski, M. and Piorkowska, E. (1979)
Polymer Bulletin, 1, 275-9. 29. Galeski, A. and Piorkowska, E. (1980) Polymer
Bulletin, 2, 1-6. 30. Wasiak, A. (1991) Chemtracts, Macromolecular Chemistry,
2, 211-45. 31. Piorkowska, E. and Galeski, A. (1985) Journal of Physical
Chemistry, 89, 4700-3. 32. Piorkowska, E. and Galeski, A. (1985) Journal of
Polymer Science, Polymer Physics Edition, 23, 1723-48.

References 137 33. Piorkowska, E. (1994) Journal of Polymer Science, Part B,


Polymer Physics, in press. 34. Misra, A., Prud’homme, R. E. and Stein, R. S.
(1974) Journal of Polymer Science, Polymer Physics Edition, 12, 1235-8. 35.
(1974) Journal of Polymer Science, Polymer Physics Edition, 12, 1235-8. 35.
Keith, H. D., Padden Jr, F. J, Walter, Μ. M. and Wyckoff, H. W. (1959) Journal
of Applied Physics, 30, 1485-8. 36. Natta, G. and Corradini, P. (1960) Nuovo
Cimento, Suppl., 15, 40-51. 37. Addink, E. J. and Beintema, J. (1961) Polymer,
2, 185-93. 38. Kardos, J. L., Christiansen, A. W. and Baer, E. (1966) Journal of
Polymer Science, Part A2, 4, 777-88. 39. Marigo, A., Marega, C., Zanetti, R. et
al. (1989) Makromolekulare Chemie, 190, 2805-13. 40. Marigo, A., Marego, C.,
Zanetti, R. et al. (1990) Makromolekulare Chemie, 191, 1967-71. 41. Pae, K. D.
(1968) Journal of Polymer Science, Part A2, 6, 657-63. 42. von Falkai, B. and
Stuart, H. A. (1959) Kolloid Zeitschrift, 162, 138-40. 43. von Falkai, B. (1960)
Makromolekulare Chemie, 41, 86-109. 44. Majer, J. (1960) Kunststoffe, 50,
565-7. 45. Limbert, F. J. and Baer, E. (1963) Journal of Polymer Science, Part A,
1, 3317-31. 46. Donald, H. J., Hume, E. S. and White, L. W. (1964) Journal of
Polymer Science, Part C, 6, 93-9. 47. Hoshino, S., Meinecke, E., Powers, J. et al.
(1965) Journal of Polymer Science, Part A, 3, 3041-65. 48. Johnsen, U. and
Spilgies, G. (1972) Kolloid Zeitschrift und Zeitschrift fiir Polymere, 250, 1174-
81. 49. Wunderlich, B. (1976) in Macromolecular Physics, Vol. II: Crystal
Nucleation, Growth, Annealing, Academic Press, New York, pp. 235-9. 50.
Burns, J. R. and Turnbull, D. (1966) Journal of Applied Physics, 37, 4021-6. 51.
Koutsky, J. A., Walton, A. G. and Baer, E. (1967) Journal of Applied Physics,
38, 1832-9. 52. Vonnegut, B. (1948) Journal of Colloid Science, 3, 563-70. 53.
Keith, F. J. and Padden, H. D. (1959) Journal of Applied Physics, 30, 1479-84.
54. Monasse, B. and Haudin, J. M. (1985) Colloid and Polymer Science, 263,
822-31. 55. Binsbergen, F. L. and de Lange, B. G. M. (1969) Polymer, 9, 23-40.
56. Bartczak, Z., Galeski, A. and Pracella, M. (1986) Polymer, 27, 537-43. 57.
Galeski, A., Bartczak, Z. and Pracella, M. (1984) Polymer, 25, 1323-6. 58.
Bartczak, Z. and Galeski, A. (1990) Polymer, 31, 2027-38. 59. Goldfarb, L.
(1978) Makromolekulare Chemie, 179, 2297-303. 60. Pae, K. D. and Sauer, J. A.
(1968) Journal of Applied Polymer Science, 12, 1901-19. 61. Sauer, J. A. and
Pae, K. D. (1968) Journal of Applied Polymer Science, 12, 1921-38. 62. Fatou,
J. G. (1971) European Polymer Journal, 7, 1057-64. 63. Samuels, R. J. (1975)
Journal of Polymer Science, Polymer Physics Edition, 13, 1417-46. 64. Mucha,
M. (1981) Journal of Polymer Science, Polymer Symposia, 69, 79-89. 65.
Wunderlich, B. (1976) Macromolecular Physics, Vol. II: Crystal Nucleation,
Growth, Annealing, Academic Press, New York, pp. 52-69. 66. Vidotto, G.,
Levy, D. and Kovacs, A. J. (1969) Kolloid Zeitschrift und Zeitschrift fur
Polymere, 230 289-305. 67. Bassett, D. C. and Olley, R. H. (1984) Polymer, 25,
935-43.

138 Nucleation of polypropylene 68. Bassett, D. C. (1988) in Developments in


138 Nucleation of polypropylene 68. Bassett, D. C. (1988) in Developments in
Crystalline Polymers, Vol. 2 (ed. D. C. Bassett), Elsevier, London, p. 102. 69.
Olley, R. H. and Bassett, D. C. (1989) Polymer, 30, 399-409. 70. Wunderlich, B.
(1976) in Macromolecular Physics, Vol. 2: Crystal Nucleation, Growth,
Annealing, Academic Press, New York, pp. 44-52. 71. Beck, Η. N. and
Ledbetter, H. D. (1965) Journal of Applied Polymer Science, 9, 2131-42. 72.
Beck, Η. N. (1967) Journal of Applied Polymer Science, 11, 673-85. 73. Kargin,
V. A., Sogolova, T. T., Rapoport, N. Ya. and Kurbanova, I. I. (1967) Journal of
Polymer Science, Part C, 16, 1609-17. 74. Menczel, J. and Varga, J. (1983)
Journal of Thermal Analysis, 28, 161-74. 75. Kowalewski, T. and Galeski, A.
(1989) Journal of Applied Polymer Science, 32, 2919-34. 76. Dragaun, H.,
Hubeny, H. and Muschik, H. (1977) Journal of Polymer Science, Polymer
Physics Edition, 15, 1779-99. 77. Shi, G.-Y. and Zhang, J. -Y. (1982) Kexue
Tongbao, 27, 290-4. 78. Varga, J., Garzo, G. and Ille, A. (1986) Angewandte
Makromolekulare Chemie, 142, 171-81. 79. Lovinger, A. J., Chua, J. O. and
Gryte, C. C. (1977) Journal of Polymer Science, Polymer Physics Edition, 15,
641—56. 80. Varga, J. and Toth, F. (1986) Makromolekulare Chemie,
Macromolecular Symposia, 5, 213-23. 81. Varga, J. (1989) Journal of Thermal
Analysis, 35, 1891-1912. 82. Varga, J. and Garzo, G. (1990) Angewandt
Makromolekular Chemie, 180, 15-24. 83. Varga, J., Fujiwara, Y. and Ille, A.
(1990) Periodica Polytechnica, Chemical Engineering, 34, 255-71. 84. Varga, J.
and Garzo, G. (1991) Acta Chimica Hungarica, 128, 303-17. 85. Varga, J. and
Toth, F. S. (1991) Angewandte Makromolekulare Chemie, 188, 11-25. 86.
Turner Jones, A., Aizlewood, J. M. and Beckett, D. R. (1964) Makromolekulare
Chemie, 75, 134-58. 87. Fujiwara, Y. (1975) Colloid and Polymer Science, 253,
273-82. 88. Ullmann, W. and Wendorff, J. H. (1979) Progress in Colloid and
Polymer Science, 66, 25-33. 89. Moss, K. H. and Tilger B. (1981) Angewandte
Makromolekulare Chemie, 94, 213-25. 90. Shi, G.-Y., Huang, B. and Zhang, J.-
Y. (1984) Makromolekulare Chemie, Rapid Communications, 5, 573-8. 91.
Jacoby, P., Bersted, B. H., Kissel, W. J. and Smith, C. E. (1986) Journal of
Polymer Science, Polymer Physics Edition, 24, 461-91. 92. Leugering, H. J. and
Kirsch, G. (1973) Angewandte Markomolekulare Chemie, 33, 17-23. 93.
Leugering, H. J. (1967) Makromolekulare Chemie, 109, 204-16. 94. Garbarczyk,
J. and Paukszta, D. (1985) Colloid and Polymer Science, 263, 985-90. 95. Shi,
G.-Y., Zhang, X.-D. and Qiu, Z.-X (1992) Makromolekulare Chemie, 193, 583-
91. 96. Garbarczyk, J. (1985) Makromolekulare Chemie, 186, 2145-51. 97.
Forgacs, P., Tolochko, B. P. and Sheromov, M. A. (1981) Polymer Bulletin, 6,
127-33. 98. Lauritzen Jr, J. I. and Hoffman, J. D. (1973) Journal of Applied
Physics, 44, 4340-52.
References 139 99. Sanchez, I. C. and DiMarzio E. A. (1972) Journal of
Research of the National Bureau of Standards, 76A, 213-23. 100. Hoffman, J. D.
(1983) Polymer, 24, 3-23. 101. Krigbaum, W. R. and Uematsu, I. (1965) Journal
of Polymer Science, Part A, 3, 767-76. 102. Lovinger, A. J. (1983) Journal of
Polymer Science, Polymer Physics Edition, 21, 97-110. 103. Lauritzen, J. I.
(1973) Journal of Applied Physics, 44, 4353-9. 104. Ewen, J. A., Johns, R. L.,
Razavi, A. and Ferrara, J. D. (1988) Journal of the American Chemical Society,
110, 6255-6. 105. Corradini, P., Natta, G., Ganis, P. and Temussi, P. A. (1967)
Journal of Polymer Science, Part C, 16, 2477-84. 106. Lotz, B., Lovinger, A. J.
and Cais, E. (1988) Macromolecules, 21, 2375-82. 107. Chatani, Y. and
Maruyama, H. (1991) Journal of Polymer Science, Polymer Physics Edition, 29,
1649-52. 108. Chatani, Y., Maruyama, H., Noguchi, K. et al (1990) Journal of
Polymer Science, Polymer Letters, 28, 393-8. 109. Miller, R. L. and Seeley, E.
G. (1982) Journal of Polymer Science, Polymer Physics Edition, 20, 2297-2307.
110. Natta, G., Corradini, P. and Ganis, P. (1962) Journal of Polymer Science,
58, 1191-9. 111. Boor Jr, J. and Youngman, E. A. (1965) Polymer Letters, 3,
577-80. 112. Haftka, S. and Konnecke, H. (1991) Journal of Macromolecular
Science, Physics, B30, 319-34. 113. Boor Jr, J. and Youngman, E. A. (1966)
Journal of Polymer Science, Part Al, 4, 1861-84. 114. Aggarwal, S. L. (1975)
Physical constants of poly (propylene) in Polymer Handbook, 2nd edn (eds J.
Brandrup, E. H. Immergut and W. McDowell), Wiley, New York, pp. V23-8.
115. Lovinger, A. J., Davis, D. and Lotz, B. (1991) Macromolecules, 24, 552-60.
116. Lovinger, A. J., Lotz, B., Davis, D.D. and Padden Jr., F. J., (1993) Macro¬
molecules, 26, 3494-503. 117. Lovinger, A. J., Lotz, B. and Davis, D. (1990)
Polymer, 31, 2253-60. 118. Rodriguez-Arnold, J., Zhang, A., Chen, Z. D.,
Lovinger, A. J., et al., (1994) Polymer, 35, 1884-95.

5 Epitaxial growth on and with polypropylene J. Petermann 5.1


INTRODUCTION Epitaxy was discovered almost 150 years ago when it was
recognized by mineralogists that various natural minerals appear in unique
shapes, and those minerals were reproduced artificially by growing them out of
solutions [1]. With the discovery of X-ray diffraction, more understanding of the
atomistic nature of the interface leading to the oriented overgrowth was
obtained. Royer [2] introduced the term ‘epitaxy’ and postulated some rules for
its occurence, the most important of which is that the oriented over¬ growth of
one crystal onto another involves the parallelism of two crystal lattice planes that
have nearly identical arrangements of atoms. The upper limit is about 15% misfit
of the spacings of the atoms in the contact lattice planes of substrate and
overgrown material. Despite some exceptions, reported in the literature [3], this
rule is still used, but it is now based on potential energy calculations between the
rule is still used, but it is now based on potential energy calculations between the
atoms in the two contact planes, adding some defect structures (e.g. misfit dis¬
locations) for minimizing their potential energy. It was recently discovered that
even on surfaces of amorphous materials (Si02) crystals can grow epitaxially
when the surfaces of the substrates have a certain topology [4]. In this case, the
surfaces are micro-structured by lithographic methods, as they are known from
the production of microelec¬ tronic devices. In the literature, this phenomenon is
now called graphoepitaxy and has been demonstrated to occur even on
monatomic steps in the surfaces of ionic crystals [5]. The origin of this type of
epitaxy is believed to be the interface energy for the nucleation event of the
overgrowth. Polypropylene: Structure, blends and composites. Edited by J.
Karger-Kocsis. Published in 1995 by Chapman & Hall, London. ISBN 0 412
58430 1

Homoepitaxy in isotactic polypropylene 141 Epitaxial growth became


technically important with the development of semiconductor devices. In order
to study orientational effects during the crystallization on a substrate, the use of a
macroscopic single crystal with a well determined crystallographic characterized
surface is most appropriate. In general, the overgrowing material can be put onto
the substrate surfaces and crystallized from three different ‘states’: • Vapour
(molecular beam deposition, chemical vapour deposition, sput¬ tering,
evaporation) • Solution (solution deposition, electro-chemical deposition) • Melt
Experimental investigations of epitaxial effects involve X-ray diffraction,
transmision electron microscopy (TEM) and electron diffraction, surface specific
methods such as Auger spectroscopy, X-ray photon analysis and others. The
epitaxy of polymers were first observed by crystallization from solutions and
later during polymerization on ionic-crystal substrates [6, 7]. It was subsequently
found that a large variety of inorganic and low molecular weight organic
substrates can force polymer crystals into an epitaxial oriented crystallization.
Excellent reviews are given by Mauritz, Baer and Hopfinger [8] and Wittmann
and Lotz [9]. Experimental difficulties delayed the investigation of polymer
epitaxy on polymer substrates. The most severe obstacle is the preparation of
large crystallographic oriented surfaces of the polymer substrate, as polymer
single crystals are only of microscopic dimensions. Wunderlich and Melillo [10]
were the first to demonstrate that folded chain crystals of polyethylene (PE)
grow epitaxially on extended chain polyethylene crystals, and later it was
demonstrated that the lamellar overgrowth in the shish-kebab structure occurs
epitaxially [11, 12]. As the overgrowth is of the same material as the substrate,
this form of epitaxy was called homoepitaxy. Generally, the crystallographic
orientations between substrate and overgrowth are identical in homoepitaxies.
orientations between substrate and overgrowth are identical in homoepitaxies.
An exception is found for isotactic polypropylene (iPP), which is discussed in
Section 5.2. 5.2 HOMOEPITAXY IN ISOTACTIC POLYPROPYLENE iPP
can crystallize in three different crystal modifications. The dominant form is the
monoclinic a modification (α-PP) with lattice constants a = 0.666nm, h =
2.078nm, c = 0.6495nm and a = 99.62° [13]. The β modi¬ fication (jS-PP) can
be obtained by crystallization from the melt using special nucleation agents [14].
It has a pseudo-hexagonal unit cell with dimensions a = 1.908 nm and c =
0.649nm [15, 16]. The y modification (y-PP), which is very seldom observed,
has a triclinic unit cell [15]. The a modification exhibits a characteristic cross-
hatched texture, which has not been observed in other crystalline polymers
(Figure 5.1). The texture

142 Epitaxial growth Figure 5.1 TEM micrograph showing the cross-hatched
texture in a spherulite of iPP. can be obtained from solution crystallization of the
so-called quadrites [17], from spherulitic crystallization in the bulk [18] and
from fiber spinning [19]. The characteristic features of this texture are that it
consists of two sets of lamellae placed at an angle of 100° to each other, their
crystallographic b-axes are in common and the αγ- and coaxes of one set of
lamellae are parallel to the c2- and a2-axes of the other. The occurrence of this
texture is interpreted with the rather small mismatch between the a- and c-axes
(2.3%) in the two sets with the (010) planes in conduct. Models of the
homoepitaxy on the molecular level have been published by Khoury [17],
Padden and Keith [20] and Lovinger [21]. Recently, Lotz and Wittman [22]
presented a very elaborate suggestion, that the lamellar branching in iPP may be
the result of a crystal growth defect. If on a (0 1 0) face of the α-PP, consisting
of helices of the same handed¬ ness, an iPP helix with the same handedness is
deposited (for growth of α-PP single crystal a helix of opposite handedness is
required), then favourable interaction of the methyl groups is only obtained
when the helix axes are inclined about 100° (Figure 5.2). In other words, the
linear array of methyl side groups on the (0 1 0) plane of one set of the lamellae
fits into the channels between similar arrays in the (0 1 0) plane of the other set.
This kind of ‘stacking fault’ accounts for the experimental observation that
under rapid crystallization conditions lamellar branching in iPP is favoured.

Heteroepitaxy in isotactic polypropylene 143 C2 1 c 2 I I 3 2 c I 3 2' c I I a1 a1


Figure 5.2 Arrangement of the helical molecules of iPP at the interface in a
cross-hatched structure. 5.3 HETEROEPITAXY IN ISOTACTIC
POLYPROPYLENE The interest in heteroepitaxy of polyolefins on inorganic
and organic substances arose with the need for nucleating agent systems for
and organic substances arose with the need for nucleating agent systems for
these polymers. In particular, the crystallization of polyethylene on (0 0 1) faces
of alkali halides was successfully investigated [6, 23, 24]. Polypropylene
exhibited orientational growth on (00 1) faces on NaCl [25]. In general, it was
found that the polyolefin chains align along the <110) directions on the (001)
faces of the alkali halides. Lattice matching is unlikely, and the reason for the
alignment is rather an electrostatic interaction of ion rows of like charges with
the polymer chains [8, 26, 27]. Many polyolefins, aliphatic polyesters and
polyamides were found to crystallize epitaxially on the salts of benzoic acid [28]
and other organic salts [9, 29, 30]. In these cases, lattice matching between the
salts and the oriented grown polymers may play a more important role [9]. Early
investigations of polymer heteroepitaxies on polymer substrates were devoted
mainly to nucleation and transcrystalline growth. In contrast to experiments with
ionic and organic substrates, no large specific crystallo¬ graphic surfaces of the
substrates were available due to the lamellar nature of the polymer crystals. First
attempts were made using melt crystallized

144 Epitaxial growth non-oriented polymer films as substrates, and the


crystallization of the layered polymer films was followed by polarizing optical
microscopy [31, 32]. In several investigations it was found that polyamide
substrates nucleate iPP [31, 33]. With the increasing commercial success of
polymer blends, research activity concentrated on the field of polymer interfaces,
because it was quickly realized that interfaces are the mechanically weak
structures in the blends. Cold drawn blends of iPP and high density PE were
heated to a temperature between the melting points of PE (140 °C) and iPP (165
°C) [34, 35]. When recrystallizing the PE, a cross-hatched arrangement of the
lamellae was found using TEM (surface replication) [34] and scan¬ ning
electron microscopy (SEM) (ion etched surfaces) [35]. The cross- hatched
appearance of the PE lamellae was attributed to either thermal shrinking stresses
[34] or oriented spherulitic growth [35]. At about the same time, very similar
morphological structures were obtained, using quadrites of iPP as substrates and
depositing PE by vaporiza¬ tion under vacuum [36], and by using an ultra-thin
highly oriented polypropylene film, sandwiched with a non-oriented PE film,
sub¬ sequently heat treated at 150°C and recrystallized at various tempera¬ tures
[37-40]. Figure 5.3 shows a TEM micrograph of such a cross- hatched structure
and its corresponding electron diffraction pattern. In both cases it was supposed
that heteroepitaxy between the two com¬ ponents was the reason for the cross-
hatched structure. Shortly after these investigations it was found that nearly
identical morphological structures were obtained, when using irans-
polyoctenamer and trans- 1,4-poly butadiene instead of polyethylene [41, 42].
Also, highly oriented polyamides such as PA-11, PA-6, PA-6.6 and PA-12
Also, highly oriented polyamides such as PA-11, PA-6, PA-6.6 and PA-12
crystal¬ lize on iPP in a similar cross-hatched structure [43, 44]. The same
results were obtained not only from crystallization from a melt or from
vaporized polymers but also from crystallization from solutions [45, 46]. All
investigated systems to date have had two facts in common: • One component is
iPP and the other component contains aliphatic sequences with all trans-
conformations. • The chain directions of the two components have an inclination
angle of ±50°. Figure 5.4a-c shows TEM micrographs of the cross-hatched
structures of iPP as substrate with polyoctenamer and trans- 1,4-poly butadiene,
and of polyamide 12 as substrate with iPP, respectively. This kind of structure
formation is unique in polymer morphologies and further experimental and
theoretical work is required to interpret it.

Origins of the Heteroepitaxies 145 (a) (b) Figure 5.3 (a) TEM micrograph
showing the cross-hatched structure of PE lamel¬ lae on an uniaxially oriented
film of iPP. (b) Electron diffraction patterns of the uniaxially oriented
polypropylene film (a modification) and the epitaxially crystal¬ lized PE
lamellae. 5.4 ORIGINS OF THE HETEROEPITAXIES IN ISOTACTIC
POLYPROPYLENE Explaining the experimental observations on the basis of
an atomistic lattice matching theory has met with several experimental
difficulties. Polymer

Epitaxial growth a) (c) TEM micrographs of epitaxially crystallized (a) pol; 4-


polybutadiene on iPP and (c) iPP on polyamid 12.

Origins of the Heteroepitaxies 147 crystals have only microscopic dimensions,


which exclude them from most investigations aimed at determining the exact
crystal lattice orientations of substrate and layer (e.g. X-ray techniques). Only
electron diffraction is able to supply diffraction patterns from such small
crystals. Secondly, the epitaxial orientation relationships may be maintained
only in the very close vicinity of the interface due to disorientation of the
lamellar crystals during growth (e.g. lamellar twisting) and consequently may
not be detected by X-ray investigations. Very skillful experimental techniques
are needed to prepare ultra-thin (<0.1 pm) substrates and layers. In addition, the
crystal¬ lography of the polymer crystals also has to describe the positions of the
side groups on the polymer chains, as they contain the atoms which interact at
the interfaces, and this kind of crystallography needs a profound knowledge in
this field. Two different attempts were made to obtain ultra-thin samples for the
substrates and layers. Petermann and Gohil used a drawing technique [47] to
obtain ultra-thin uniaxial oriented films as a substrate and put solution cast films
obtain ultra-thin uniaxial oriented films as a substrate and put solution cast films
on them. After melting the solution cast films, they followed their crystallization
with TEM and electron diffraction. Lotz and Wittmann [36] crystallized the
polymer substrate epitaxially on benzoic acid and evaporated under vacuum the
layers on the substrates. While the epitaxially grown substrates have a single
crystal-like orientation, the drawn films have only a fiber texture, which means
that all crystalline lamellae have the <001) direction in common but their </ifco>
directions are rotated statistically around the <001) axis. Clearly, the substrates
with the single crystal-like orientations together with the layers cannot only give
informa¬ tion about the orientation relationships between both polymer chain
direc¬ tions, as can the drawn films as substrates, but also about the lattice
planes being in contact with each other. For the monoclinic a modification of iPP
as the substrates and PE as the layers it was found [36, 43, 9] that:
[001]pE//[101]iPP and (1 00)PE//(0 1 0)iPP and for polyamides are substrates
and iPP as layers [43, 9] [001]PA//[101]iPp and (0 1 0)PA//(0 1 0)iPP In both
epitaxies, the (0 1 0) lattice plane of the monoclinic iPP is involved (the same
lattice plane is involved in the formation of the polypropylene quadrites). A
characteristic geometry of this lattice plane is a dense row of methyl groups in
the [1 0 1] direction, sticking out of the plane. The distance between two
adjacent rows is 0.505 nm (Figure 5.5). Lotz and Wittman [43] explained the
epitaxy by a parallel alignment of the aliphatic sequences of the layered
polymers with the methyl rows. While there is no atomic matching between the
atoms of the aliphatic sequences and methyl groups along the rows, the
intermolecular distances between two adjacent rows (0.505 nm) and the distance
between the adjacent molecules in the contact planes of the layers match quite
well (0.494 nm for PE). Also, the angle

148 Epitaxial growth Figure 5.5 Schematic sketch of the methyl side group
packing (large circles) in the (0 1 0) lattice plane of the monoclinic phase of iPP.
between the direction of the methyl rows in the (0 1 0) lattice plane of the
monoclinic isotactic polypropylene and its chain direction (50°) fits the
experimental data. This geometrical model can be considered as a one¬
dimensional epitaxy. There are still some observations that cannot be explained
by pure geometrical considerations. Contrary to expectations, that the epitaxy
should be more perfect the more slowly the crystallization occurs, the reverse is
observed [48]. High crystallization rates lead to perfect parallel alignment of the
PE lamellae (Figure 5.6a); at slower crystallization rates (at elevated
temperatures) the alignment is less perfect (Figure 5.6b); and at very slow
crystallization rates the epitaxy even be suppressed (Figure 5.6c). For systems of
polyamides with iPP, epitaxy can only be observed after rapid quenching [44].
polyamides with iPP, epitaxy can only be observed after rapid quenching [44].
Furthermore, in the experimental set-up with the uniaxially oriented films having
a fiber texture, only a small fraction of the polypropylene lamellae are expected
to have a (0 1 0) surface for nucleating the PE epitaxially. But although the
polypropylene and PE lamellae have about the same size in Figure 5.6a, all PE
lamellae are oriented epitaxially, indicating that each polypropylene lamellae has
nucleated a PE lamellae epitaxially, even those that do not have a (0 1 0) surface.
Also, the direction of the methyl rows in the hexagonal iPP (β modification) is
expected to have a different inclination angle to the chain direction than in the
monoclinic a modification; the alignment of the PE chain directions compared to
the polypropylene chain direction is exactly the same in both the polypropylene
modifications [48] used as a substrate (Figure 5.7). The geometric aspects may
play an important role in the formation of epitaxial interfaces of polymer-
polymer systems, but the interactions of different chemical groups

Origins of the Heteroepitaxies :a) (b) (c) 6 TEM micrographs of epitaxially


polyethylene lamellae on u iPP, crystallized at (a) room temperature, (b) 115 °C
and (c) 125 as with cross-hatched structures crystallized when taking the si
iperature.)

150 Epitaxial growth (a) (b) Figure 5.7 (a) TEM micrograph showing epitaxially
crystallized PE on a uniaxially oriented film containing the hexagonal (/?) phase
of polypropylene, (b) Electron diffraction patterns of the uniaxially oriented
polypropylene film (β modification) and the epitaxially crystallized PE lamellae.

Epitaxies on isotactic polypropylene with normal paraffins 151 (such as methyl


or phenyl) may be as important. These interactions may already be strong
enough when the layered component is still in the molden state to lead to
monomolecular adsorbed layers at the interface, producing a nucleation-induced
orientation of the layered polymer during crystalliza¬ tion. Atomic force
microscopy [49, 50] and dynamic molecular modeling calculations may lead in
the future to better understanding of the origin of the polymer-polymer epitaxies.
5.5 HETEROEPITAXIES IN SYNDIOTACTIC POLYMERS Recently, new
catalysts were developed to polymerize highly stereoregular syndiotactic
polyolefins, especially syndiotactic polystyrene and syndiotactic polypropylene
(sPP). Both of these polymers differ in chain conformations from their isotactic
counterparts. The syndiotactic polystyrene has an all trans- (T) conformation
(linear zig-zag chain) with an identity period of 0.51 nm [51]. When it
crystallizes from solutions, a chain conformation of TTGG was observed [52].
The crystals from both chain conformations exhibit complex polymorphisms,
which are still under discussion [53, 54]. sPP has a TTGG chain conformation
which are still under discussion [53, 54]. sPP has a TTGG chain conformation
[55] and crystallizes in an orthor¬ hombic unit cell with a=1.45nm, 6=1.12nm
and c = 0.74nm [56, 57]. Epitaxial crystallization of iPP on uniaxially oriented
syndiotactic polysty¬ rene films was performed at different crystallization
temperatures [58]. Low crystallization temperatures (Tc< 90 °C) resulted in a
cross-hatched arrange¬ ment of iPP lamellae with the chain direction of
syndiotactic polystyrene and iPP ±40° apart. At elevated temperatures, lath-like
iPP crystals appear with their chain axis perpendicular to the substrate surface
and their crystallographic α-axis (from the monoclinic modification) parallel to
the syndiotactic polystyrene chains. No explanation for the two types of epitaxy
has been offered. Recently, epitaxial crystallization of sPP on highly uniaxially
oriented PE was reported [59]. The inclination of the chain directions of both the
polymers is ±37° and the (0 1 0) plane of sPP is in contact with the PE substrate
surface. A geometrical one-dimensional matching may explain the relation [60]:
(0 1 0)sPP^(l 1 0)PE and [1 0 l]sppy/ [0 0 1]PE Also, an epitaxial relationship of
the monoclinic iPP quadrites as substrate with sPP was reported [61] and
interpreted with a matching model. 5.6 EPITAXIES ON ISOTACTIC
POLYPROPYLENE WITH NORMAL PARAFFINS Highly oriented fiber
textured iPP films containing the monoclinic a modification were used as
substrates to study the epitaxial crystallization of normal paraffins of n-C20H42
to n-C50H102 [62]. Thin paraffin films were

152 Epitaxial growth deposited on the polypropylene and heat treated in order to
recrystallize the paraffins on the substrate. Different kinds of epitaxies were
observed depending on the paraffin chain length. All paraffins crystallized in the
orthorhombic modification. For the shorter paraffins (to C40H82) the paraffin
molecules are perpendicular to the substrate surface and three different kinds of
orientations with respect to the molecular direction [001] of the iPP were
observed. (0 0 1)P is the contact plane of the paraffin to the substrate and (0 1
0)iPP is assumed as the ‘nucleating’ plane of the iPP. The following
relationships were obtained: [110]P//[101]iPP [110]P//[001]„ [110]P//[101]iPP
Interestingly, these orientation relationships can be interpreted similar to the
interpretation by Wittmann and Lotz for the PE/iPP system (Figure 5.6).
Assuming the (010) plane of the isotactic polypropylene as the nucleating
contact plane, three different directions with close packed methyl side groups
can be seen: [1 0 1], [Ϊ 0 1] and [0 0 1]. While in the case of the epitaxy with PE,
the zig-zag chain of the PE alignes parallel to the closest direction of the methyl
side groups, [1 0 1], in the case of paraffins the closest packed direction of the
methyl end groups of the paraffins ([1 1 0]) aligns parallel to the close packed
directions of the methyl side groups in the (0 1 0) plane of the iPP. Hence,
directions of the methyl side groups in the (0 1 0) plane of the iPP. Hence,
interactions of methyl groups may also be considered as an important factor in
these epitaxies. When the length of the paraffin molecules exceeds C40H82, the
paraffins obey the same epitaxial relationship as the PE: [00 1]P//[1 0 l]iPP and
(1 00)p//(0 1 0)iPP 5.7 DECORATIONAL EFFECTS ON ISOTACTIC
POLYPROPYLENE Surface topologies can be studied by the decoration
technique [63, 64]. For decorating a surface, mostly metal atoms (Au) are
evaporated as an extremely thin layer onto the surface to be investigated. By
preferred nucleation of the metal atoms on surface irregularieties (e.g. steps),
topolo¬ gies of atomic height can be made visible. Using ultra-thin uniaxially
oriented iPP films as substrate with preparation conditions so that they have a
needle crystal morphology [65], and subsequently decorating them with tin (Sn),
the needle crystal morphology can be seen from the pearl-string like alignment
of the tin particles along the molecular direction (Figure 5.8). Furthermore,
electron diffraction patterns of the decorated films show a preferred orientation
of the tin particles. The [100] lattice direction of the tetragonal modification of
Sn is parallel to the molecular direction

Decorational effects on isotactic polypropylene 153 (a) (b) Figure 5.8 (a) TEM
micrograph of a uniaxially oriented polypropylene film decor¬ ated with tin
particles, (b) Electron diffraction pattern of the polypropylene film and the tin
particles.

154 Epitaxial growth of the iPP film (Figure 5.8b) [66, 67]. Similar results were
obtained when using tellurium (Te), bismuth (Bi) and indium (In) as the metals
for the decoration [68, 69]. The cause of the preferred orientation is not yet
established. Lattice matching between the polymer crystals and the metals seems
very unlikely. PE vapour had been used to decorate the fold surfaces of lath-
shaped iPP single crystals [70]. Chain folding directions on the fold surfaces of
polymer single crystals and sectorization were studied by this method. 5.8 THE
INFLUENCE OF EPITAXIAL INTERFACES ON THE MECHANICAL
PROPERTIES OF POLYPROPYLENE BLENDS AND LAMINATES In the
earlier stages of the investigations of polymer epitaxies, the main interest
resulted from their potential use as nucleating agents [28]. However, a
straightforward correlation was not established. More recently, the prob¬ lems of
interface adhesion in polymer blends triggered further research in the field of
polymer-polymer epitaxy. In the first studies of PE-iPP blends no significant
improvements in the mechanical properties were reported [71- 73], but in
subsequent works synergetic effects in yield strength and tensile strength were
found [74, 75]. At this time the epitaxial crystallization of PE on iPP was not
known. Gross and Petermann [37] prepared blends from a common solution by
known. Gross and Petermann [37] prepared blends from a common solution by
casting the solution onto a hot glass slide, evaporating the solvent and picking up
the remaining molten blend by a roller with a take-up speed of 4 cm/s (Figure
5.9). The resulting films were highly oriented in both the phases and the phase
dispersion was well below 1 pm. Fracture stress and Young’s modulus were
measured in the orientation direction. After heat treatment of the samples above
the melting point of PE but below that of polypropylene and subsequent
quenching to room temperature, Figure 5.9 Schematic diagram for the
preparation technique for the oriented polymer blends.

Influence of epitaxial interfaces 155 yield stress and Young’s modulus were
measured again (Figure 5.10a, b). A significant increase in both these properties
on the polypropylene-rich side were registered after the heat treatment, despite
the loss of the preferred uniaxial orientation of the PE molecules during the
melting. Corresponding electron microscope investigations revealed a cross-
hatched lamellar mor¬ phology of the recrystallized PE (Figure 5.11a, b). In a
different kind of Figure 5.10 (a) Young’s modulus and (b) fracture stress versus
composition in the PE-iPP blends ( ) before and (---) after heat treatment.

156 Epitaxial growth (a) (b) Figure 5.11 (a) TEM micrograph of an as-prepared
blend of 70% iPP and 30% PE. The darker ribbons are the PE. The arrow
indicates the orientation direction of the polypropylene and PE. (b) TEM
micrograph of a heat treated (15 min at 150 °C) and subsequently quenched
blend of the same composition as in (a).

Influence of epitaxial interfaces 157 experiment, ultra-thin alternating


multilayered (about 1000 layers) uniaxially oriented PE and polypropylene films
were prepared (Figure 5.12) and mechanically tested. The morphologies of the
films and the layers before and after heat treatment are shown schematically in
Figure 5.13 [39, 40]. Again, the Young’s modulus after heat treatment improved
(Figure 5.14). The improvement was attributed to a bridging of the mechanically
soft amorphous regions in the polypropylene lamellar structure by the epi¬
taxially recrystallized PE lamellae (Figure 5.15). For an effective bridging, good
adhesion between the polypropylene and PE lamellae is required. In order to test
the adhesion, thin PE films were sandwiched between two Figure 5.12
Schematic diagram for the preparation of multilayered films of PE and
polypropylene. 1000 layers • 100 μπη (a) Figure 5.13 Schematic diagram of (a)
the layered film samples, (b) their morpho¬ logy after preparation (as-spun) and
(c) their morphology after heat treatment. The arrow denotes the deformation in
the tensile test experiments.
the tensile test experiments.

158 Epitaxial growth (gpg) as-spun and at 155°C annealed films Figure 5.14
Young’s moduli of the layered films, having non-epitaxial (as-spun) and
epitaxial (annealed) interfaces. Figure 5.15 Schematic diagram of PE lamallae,
bridging the amorphous regions (grey areas) of the polypropylene. The
deformation direction is denoted by σ. polypropylene sheets and heat treated for
epitaxial recrystallization of the PE (high density PE) and non-epitaxial
recrystallization (low density PE). The 180° peel strength of the epitaxial
interface was about one order of magnitude higher than that without the epitaxial
interface (Figure 5.16). Lee and Schultz [76] bonded highly oriented sheets of
polypropylene by thin layers of PE using heat treatments for the epitaxial
crystallization. The epitaxial recrystallization was manifested by transmission
electron micro¬ scopy. Extraordinary strong bonding was obtained after
quenching the samples from 159 °C to room temperature. The authors suggested
a cohesive failure of the PE when debonding the sheets. This implies that the
cohesive strength of PE is smaller than the adhesive strength of both the
polymers in their epitaxial interface. Annealing the bonded sheets at elevated
tempera¬ tures (113 °C and 125 °C) reduced the bond strength by nearly a factor
of

Fracture stress (107 Pa) S Young’s modulus (10 Pa) Influence of epitaxial
interfaces (N cm"1) 159 5.16 180° peel strength of two uniaxially oriented
polypropylene sheets, a thin PE layer (~0.5 pm) as an adhesive. (b)
Concentration of TOR (wt %)

160 Epitaxial growth (C) Concentration of TOR (wt °/o) (d) Concentration of
TOR (wt %) Figure 5.17 (a) Young’s moduli versus concentrations of
polypropylent/trans- (O) and cis- (·) polyoctenamer blends (TOR). The blends
with irans-polyoctenamer contain epitaxial interfaces, (b) Fracture stresses for
the same blends as in (a). The strain direction in the tensile tests is parallel to the
chain direction of the polypropy¬ lene matrix, (c) Young’s moduli measured
perpendicular to the chain direction of the polypropylene matrix, (d) Fracture
stresses of the same blends as in (c). four. This is in accordance with the
observation that the epitaxial crystalliza¬ tion diminishes at elevated
crystallization temperatures [48]. Similar results concerning the adhesion of
laminates bonded by epitaxially crystallized thin interlaminate layers has been
observed for polypropylene bonded by trans- 1,4-poly butadiene and by trans-
polyoctenamer [77, 78]. Remarkable im¬ provements in mechanical properties
were obtained in blends of iPP and
References 161 ijms-polyoctenamer [77]. The Young’s modulus and fracture
stress increase with blending of 30wt% irans-polyoctenamer despite the fact that
pure irans-polyoctenamer has a Young’s modulus and fracture stress about one
order of magnitude less than those of oriented isotactic polypropylene (Figure
5.17). For comparison, these mechanical properties were also deter¬ mined in
corresponding blends containing polyoctenamer with higher ds-contents and not
crystallized epitaxially. No synergetic effects were observed in these blends.
Again, the improvements of mechanical properties in the blends with epitaxial
interfaces are explained by the bridging of the amorphous regions with trans-
polyoctenamer lamellae. Resulting from the geometry of the bridging lamellae, a
reinforcement of the blends perpendicu¬ lar to the main molecular direction of
the polypropylene molecules is also obtained (Figure 5.15). Until now, there was
not been enough work on structure-property relationships of polypropylene
blends containing epitaxial interfaces to enable conclusions to be drawn
regarding their importance in techno¬ logical applications. However, the
phenomenon of homoepitaxy and heteroepitaxy certainly plays an important
role, in scientific understanding of the mechanical properties of the isotactic
homopolymer and blends of polypropylene. 5.9 APPENDIX: SYMBOLS AND
ABBREVIATIONS iPP Isotactic polypropylene PE Polyethylene SEM
Scanning electron microscopy sPP Syndiotactic polypropylene Tc
Crystallization temperature TEM Transmission electron microscopy α-PP a-
phase polypropylene /?-PP jS-phase polypropylene y-PP y-phase polypropylene
5.10 REFERENCES 1. Frankenheim, M. L. (1986) Annals of Physics, 37, 516.
2. Royer, L. (1928) Bulletin de la Soci0ti frangaise de Mineral Crystal., 51, 7. 3.
Schult, L. G. (1951) Acta Crystallographa, 4, 483. 4. Geis, M. W., Flanders, D.
C. and Smith, Η. I. (1979) Applied Physics Letters, 35, 71. 5. Kasukabe, Y. and
Osaka, T. (1987) Thin Solid Films, 146, 175. 6. Willems, J. (1958) Discussions
of the Faraday Society, 25, 111. 7. Lando, J. B., Baer, E., Rickert, S. E. et al
(1983) American Chemical Society Symposia Series, 212, 89.

162 Epitaxial growth 8. Mauritz, K. A., Baer, E. and Hopfinger, A. J. (1978)


Journal of Polymer Science, Macromolecular Reviews, 13, 1. 9. Wittmann, J. C.
and Lotz, B. (1990) Progress in Polymer Science, 15, 909. 10. Wunderlich, B.
and Melillo, L. (1966) Science, 154, 1329. 11. Penning, A. J., van der Mark, J.
M. A. A. and Kiel, A. M. (1970) Kolloid Zeitschrift und Zeitschrift fur
Polymere, 237, 336. 12. Petermann, J. and Gleiter, H. (1978) Progress in Colloid
and Polymer Science, 64, 122. 13. Natta, G. and Corradini, P. (1960) Nuovo
Cimento, 15, 40. 14. Morrow, D. R. (1969) Journal of Macromolecular Science,
Physics, 3, 53. 15. Turner-Jones, A., Aizlewood, J. M. and Beckett, D. R. (1964)
Makromolekulare Chemie, 75, 134. 16. Samuels, R. J. and Yee, R. Y. (1972)
Makromolekulare Chemie, 75, 134. 16. Samuels, R. J. and Yee, R. Y. (1972)
Journal of Polymer Science, Part A2, 10, 385. 17. Khoury, F. (1966) Journal of
Research of the National Bureau of Standards, 70A, 29. 18. Binsbergen, F. L.
and de Lange, B. G. M. (1968) Polymer, 9, 23. 19. Katayama, K., Amano, T. and
Nakamura, K. (1968) Kolloid Zeitschrift und Zeitschrift fur Polymere, 226, 125.
20. Padden, F. J. and Keith, H. D. (1973) Journal of Applied Physics, 44, 1217.
21. Lovinger, A. J. (1986) Journal of Polymer Science, Polymer Physics Edition,
24, 1541. 22. Lotz, B. and Wittmann, J. C. (1986) Journal of Polymer Science,
Polymer Physics Edition, 24, 1541. 23. Fischer, E. W. (1958) Kolloid Zeitschrift
und Zeitschrift fur Polymeren, 159, 108. 24. Carr, S. H., Keller, A. and Baer, E.
(1970) Journal of Polymer Science, Part A2, 8, 1467. 25. Koutsky, J. A., Walton,
A. G. and Baer, E. (1966) Journal of Polymer Science, Part A2, 4, 611. 26.
Mauritz, K. A., Baer, E. and Hopfinger, A. J. (1973) Journal of Polymer Science,
Polymer Physics Edition, 11, 2185. 27. Ihn, K. J., Tsuji, M., Isoda, S.,
Kawagushi, A., Katayama, Κ.-L, Tanaka, Y. and Sato, H. (1989)
Makromoleculare Chemie, 190, 837. 28. Wittmann, J. C., Hodge, A. M. and
Lotz, B. (1983) Journal of Polymer Science, Polymer Physics Edition, 21, 2495.
29. Yamashita, Y., Shimamura, K., Kasahawa, H. and Monobe, K. (1987)
Synthetic Metals, 17, 253. 30. Isoda, S. (1984) Polymer, 25, 615. 31. Chatterjee,
A. M., Price, F. P. and Newmann, S. (1975) Journal of Polymer Science,
Polymer Physics Edition, 13, 2369. 32. Chatterjee, A. M., Price, F. P. and
Neuman, S. (1975) Journal of Polymer Science, Polymer Physics Edition, 13,
2385. 33. Seth, K. K. and Kempster, C. J. E. (1977) Journal of Polymer Science,
Polymer Symposia, 58, 297. 34. Nishio, Y., Yamane, T. and Takahashi, T.
(1984) Journal of Macromolecular Science, Physics, B23, 17. 35. Kojima, M.
and Satake, H. (1984) Journal of Polymer Science, Polymer Physics Edition, 22,
285. 36. Lotz, B. and Wittmann, J. C. (1984) Makromolekulare Chemie, 185,
2043. 37. Gross, B. and Petermann, J. (1984) Journal of Materials Science, 19,
105.

References 163 38. Broza, G., Rieck, U., Kawaguchi, A. and Petermann, J.
(1985) Journal of Polymer Science Polymer Physics Edition, 23, 2623. 39.
Petermann, J., Broza, G., Rieck, U. and Kawaguchi, A. (1987) Journal of
Materials Science, 22, 1477. 40. Jaballah, A., Rieck, U. and Petermann, J. (1990)
Journal of Materials Science, 25, 3105. 41. Xu, Y., Kawaguchi T., Asano, U. et
al. (1989) Journal of Materials Science Letters, 8, 675. 42. Petermann, J., Xu, Y.,
Loos, J. and Yang, D. (1992) Makromolekulare Chemie, 193, 611. 43. Lotz, B.
and Wittmann, J. C. (1986) Journal of Polymer Science, Polymer Physics
Edition, 24, 1559. 44. Xu, Y. and Petermann, J., unpublished results. 45. Shen,
Y., Yang, D. and Feng, Z. (1991) Journal of Materials Science, 26, 1941. 46.
Y., Yang, D. and Feng, Z. (1991) Journal of Materials Science, 26, 1941. 46.
Kawaguchi, A., Okihara, T. and Murakama, S. (1991) Journal of Polymer
Science, Polymer Physics Edition, 29, 683. 47. Petermann, J. and Gohil, R. M.
(1979) Journal of Materials Science, 14, 2260. 48. Petermann, J. and Xu, Y.
(1991) Journal of Materials Science, 26, 1211. 49. Lotz, B., Wittmann, J.C. and
Stocker, W. (1991) Polymer Bulletin, 26, 209. 50. Fuchs, H., Eng, L. M. and
Sander, R. (1991) Polymer Bulletin, 26, 95. 51. Chatani, Y., Fujii, Y., Shimane,
Y. and Ijitsu, T. (1988) Polymer Preprints, Japan, 37, E428 (English edition). 52.
Immirzi, A., De Candia, F., Iannelli, P. et al. (1988) Makromolekulare Chemie
Rapid Communications, 9, 761. 53. Greis, O., Xu, Y., Asano, T. and Petermann,
J. Polymer, 30 1590. 54. Guerra, G., Vitagliano, V. M., De Rosa, C. et al (1990)
Macromolecules, 23, 1539. 55. Corradini, P., Natta, G., Ganis, P. and Temussi,
P. A. (1967) Journal of Polymer Science, Part C, 16, 2477. 56. Lotz, B.,
Lovinger, A. J. and Cais, R. E. (1988) Macromolecules, 21, 2375. 57. Lovinger,
A., Lotz, B. and Davies, D. D. (1990) Polymer, 31, 2253. 58. Xu, Y. and
Petermann, J. (1990) Polymer Communications, 31, 428. 59. Petermann, J., Xu,
Y., Loos, J. and Yang, D. (1992) Polymer, 33, 1096. 60. Wittman, J. C. and
Lotz, B. private communication. 61. Lovinger, A., Davies, D. D. and Lotz, B.
(1991) Macromolecules, 24, 552. 62. Kawaguchi, A., Okihara, T., Ohara, M. et
al. (1989) Journal of Crystal Growth, 94, 857. 63. Bassett, G. A. (1958)
Philosophical Magazine, 3, 1042. 64. Bassett, D. C. (1968) Philosophical
Magazine, 17, 37. 65. Schultz, J. M. and Petermann, J. (1984) Colloid and
Polymer Science, 262, 294. 66. Schultz, J. M. and Peneva, S. K. (1987) Journal
of Polymer Science, Polymer Physics Edition, 25, 185. 67. Broza, G. and
Petermann, J. (1986) Praktische Metallographie, 17, 175. 68. Petermann, J. and
Broza, G. (1987) Journal of Materials Science, 22, 1108. 69. Petermann, J. and
Hoffmann, T. (1990) Epitaxy of metals on polymer substrates, in Advanced
Materials and Processes, Vol. 2 (eds H. F. Exner and V. Schuh- macher), DGM-
Verlag. 70. Wittmann, J. C. and Lotz, B. (1985) Journal of Polymer Science,
Polymer Physics Edition, 23, 205. 71. Greco, R., Mucciariello, G., Ragosta, G.
and Martuscelli, E. (1980) Journal of Materials Science, 15, 845.

164 Epitaxial growth 72. Robertson, R. E. and Paul, D. R. (1979) Journal of


Applied Polymer Science, 17, 2579. 73. Deanin, R. D. and Sansone, M. F.
(1978) Journal of Polymer Science, Polymer Symposia, 19, 211. 74. Greco, R.,
Mucciariello, G., Ragosta, G. and Martuscelli, E. (1981) Journal of Materials
Science, 16, 1001. 75. Lovinger, A. J. and Williams, M. L. (1980) Journal of
Applied Polymer Science, 25, 1703. 76. Lee, I. H. and Schultz, J. M. (1988)
Journal of Materials Science, 23, 4237. 77. Petermann, J. and Xu, Y. (1991)
Colloid and Polymer Science, 269, 455. 78. Xu, Y., Asano, T. and Petermann, J.
Colloid and Polymer Science, 269, 455. 78. Xu, Y., Asano, T. and Petermann, J.
(1990) Journal of Materials Science, 25, 311.

Part Two Processing-induced Structure

6 Higher order structure of injection-molded polypropylene M. Fujiyama 6.1


INTRODUCTION When determining the quality of a polymer, it is necessary to
consider its processability, namely, its processing properties and the quality of
the final product (product properties). As shown in Figure 6.1, the processing
properties are controlled by the primary structures of the raw material and by
processing conditions, while the product properties depend on both the primary
and higher order structures. The higher order structures are formed during
processing and are dependent on both the primary structures and the processing
conditions. Therefore, the ‘quality’ of a polymer is decided Property (Quality)
Figure -6.1 Causality of main factors in polymer processing. Polypropylene:
Structure, blends and composites. Edited by J. Karger-Kocsis. Published in 1995
by Chapman & Hall, London. ISBN 0 412 58430 1

168 Higher order structure of injection-molded polypropylene basically by the


material and processing conditions. As far as product properties are concerned,
one has only to see how the primary structures of the material and processing
conditions affect the various product properties, but this alone does not explain
why product properties change when a material and/or processing conditions are
changed. The quality (i.e. proper¬ ties) of the product is derived directly from
the primary structures of the material and the higher order structures resulting
from the material and processing conditions. Therefore, it is necessary to clarify,
by analysing the higher order structures, how the material and processing
conditions affect the product properties. From the viewpoint of processing
properties, it is necessary to consider how primary structures such as molecular
mass, molecular mass distribution and branching, and processing conditions
such as temperature and pressure affect the rheological and thermal properties,
which influence the flow, fusion and solidification behaviors, which are crucial
processing factors. These factors, acting during processing, also have an effect
on the higher order structures, such as the orientation of molecular chains and
the size of spherulites. Thus it is necessary to consider both the behavior at the
time of processing and its effects on the higher order structures of the product.
This Chapter describes our studies on the effects of molding conditions and the
characteristics of raw resin on the higher order structures and structure-property
relationships of injection molded polypropylenes (PP). 6.2 CRYSTALLINE
STRUCTURE OF INJECTION-MOLDED PP Figure 6.2 shows a polarized
micrograph of a thin section cut perpendicular to the flow direction (MD) from
micrograph of a thin section cut perpendicular to the flow direction (MD) from
an injection-molded PP ASTM flexural test specimen (Figure 6.3) [1]. A clear
skin-core structure composed of a surface skin layer with a high molecular
orientation and an inner core layer composed of spherulites with a low molecular
orientation is observed. 2 mm i i Figure 6.2 Polarized micrograph, under crossed
polars, of thin section cut perpen¬ dicular to flow direction of injection-molded
PP.

Crystalline structure of injection-molded PP 169 Figure 6.3 ASTM flexural test


specimen. Figure 6.4 shows wide-angle X-ray diffraction patterns taken from
various directions of the skin and core layers [1]. In the skin layer, the through
view (THRU) and edge view (EDGE) show fiber patterns, and the c-axes and the
a*-axes are highly oriented to MD. The end view (END) shows Debye rings, and
the crystalline molecular chains are almost unoriented when viewed from MD.
As for the core layer, the THRU and EDGE show a weak c-axis and a*-axis
orientation to MD and the END shows an unoriented state viewed from MD.
Figure 6.5 shows the crystalline orientation states in the THRU EDGE END
THRU Skin Layer Core Layer MD ND MD TD MD Figure 6.4 Wide-angle X-
ray diffraction patterns taken from various directions of skin and core layers of
injection-molded PP.

170 Higher order structure of injection-molded polypropylene Figure 6.5


Orientation states of crystals in (a) skin layer and (b) core layer of injection-
molded PP. skin and core layers assumed from the above results. In the skin
layer, the c-axes and the a*-axes are bimodally oriented to MD and the b-axes
rotate around MD. In the core layer, the c-axis and a*-axis orientation to MD is
weak and the b-axes rotate around MD. Figure 6.6 shows small-angle X-ray
scattering patterns of the skin and core layers taken form the THRU view [1].
The skin layer shows a clear two-spot pattern in the meridional direction, which
means that lamellae whose plate planes are perpendicular to MD exist. The
thickness of the lamellae calculated from the scattering angle is about 16 nm.
Although the wide-angle X-ray diffraction patterns show that a considerable
amount of the a*-axis-oriented component exists, no scattering is observed on
the equator in the small-angle X-ray scattering. This means that the a*-axis-
oriented lamellae are small and imperfect. The core layer shows scatterings in all
azimuthal directions, which means that the orientation of lamellae is random.
Figure 6.7 shows differential scanning calorimetric (DSC) thermograms of the
skin and core layers of the injection-molded PP and a compression- molded PP
[1]. All thermograms are normalized on a same sample mass. In the skin layer, a
high temperature melting component exists whose
high temperature melting component exists whose

Crystalline structure of injection-molded PP 171 Figure 6.6 Small-angle X-ray


scattering patterns taken from through view of skin and core layers of injection-
molded PP. Figure 6.7 DSC thermograms of skin and core layers of injection-
molded PP and compression-molded PP. melting temperature extends up to 182
°C. The fraction of this component which melts above 170 °C is about 5.3%.
The densities of the skin and core layers at 23 °C are 0.9073 and 0.9067 g/cm3,
respectively: the skin layer has slightly higher density than the core layer [1].
The crystallinity of the skin layer measured with X-ray diffraction is lower than
that of the core layer. The fact that the skin layer has slightly higher density than
the core layer despite its low crystallinity means that the amorphous chains are
strained in the skin layer.

172 Higher order structure of injection-molded polypropylene Figure 6.8a shows


the temperature changes of the storage modulus E' and loss modulus E" of the
skin and core layers. Although the skin and core layers show similar E' at low
temperatures, the skin layer shows Figure 6.8 Temperature changes of (a)
storage modulus E and loss modulus E" and (b) loss tangent tan <5 of skin and
core layers of injection-molded PP (107dyn/cm2 = 1 MPa). (

Crystalline structure of injection-molded PP 173 higher E' than the core layer at
high temperatures. While the sudden drop of E' caused by melting occurs at
about 150°C for the core layer, it occurs at about 180 °C for the skin layer.
Although, for the skin layer, the final temperature at which E' suddenly drops
corresponds well to the melting point measured thermally with DSC, it happens
at a much earlier stage for the core layer. Comparison of E' in MD and that in the
transverse direction (TD) of the skin layer shows that, although the former is
higher in the temperature range below 33 °C (which is slightly higher than the
glass transition temperature, Tg»20°C), the order reverses above 33 °C. The final
sudden drop of E' occurs at the same temperature of about 180 °C in both MD
and TD. The fact that the E' values in MD and TD cross each other at a
temperature slightly higher than Tg has been reported for cold-drawn and
annealed semicrystalline polymers [2]. It has been shown that the fact that the E'
values in MD and TD cross at a temperature slightly higher than the primary
relaxation transition tempera¬ ture can be interpreted when the serial connection
of crystalline (C) and amorphous (A) is strong in a mechanical model with
crystalline and amorphous regions. Figure 6.8b shows the temperature changes
of loss tangent tan <5 of the skin and core layers [1]. Comparing tan δ in MD of
the skin layer and that of the core layer, the Tg of the skin layer is about 2°C
higher than that of the core, and this peak in the case of the skin layer is smaller
higher than that of the core, and this peak in the case of the skin layer is smaller
than that of the core. This means that the amorphous molecular chains in the skin
layer are more extended and strained than those in the core layer. The
crystalline-phase related relaxation of the skin layer is weaker than that of the
core layer due to the lower crystallinity of the skin layer. While the core layer
shows a final rise of tan δ by melting at about 150 °C, the skin layer melts from
about 160 °C with a final rise in tan δ at about 180 °C. As an indication for the
anisotropy in the skin layer, the Tg in TD is about 7°C higher than that in MD
and the Tg peak in TD is smaller than in MD. This may be because in MD the
serial connection of C and A is stronger than the parallel connection of C and A
in the mechanical model of Takayanagi, Imada and Kajiyama [2], so the
character of the A region appears stronger in MD than in TD. Considering the
crystalline phase related mechanical relaxation peak, both its temperature and its
intensity are higher in TD than in MD. This may be because the parallel
connection of C and A in TD is stronger than the serial connection of C and A in
the mechanical model [2], so the character of the C region appears stronger in
TD than in MD. The final rise of tan δ by melting occurs at about 180 °C in both
MD and TD. Figure 6.9 shows the stress-strain curves of the skin and core layers
[1]. The skin layer in MD shows a very high yield stress with a dull yielding
peak, ruptures just after yielding and does not show necking. The core layer in
MD shows a low yield stress with a sharp yielding peak and a

174 Higher order structure of injection-molded polypropylene Figure 6.9 Stress-


strain curves of skin and core layers of injection-molded PP (1 kg/cm2 = 0.1
MPa). long necking region before rupture. The stress-strain curve of the core
layer is similar to that of the compression-molded specimen. The skin layer in
TD shows a further lower yield stress than the core layer in MD and shows
necking before rupture. The necking stress of the skin layer in TD is lower than
that of the core layer in MD. The yielding peak of the skin layer in TD is less
sharp than that of the core layer in MD. The yield stress of the skin layer in TD
is less than half that in MD. As mentioned above, in the skin layer, lamellae'are
perpendicular and parallel to the flow direction (MD) and there are crystallites
with high melting temperature and high strength. In general, a PP melt
crystallizes under a high shear stress in injection molding. This is particularly
notable near the surface. Therefore, it is considered appropriate to apply a shish-
kebab structure, proposed by Keller and Machin [3], to the structure of the skin
layer. Figure 6.10 shows a modified shish-kebab structure drawn to fit the actual
case on consideration of the above results. Crystalline lamellae - ‘kebabs’ - fill
the spaces, and fibrous crystals - ‘shishes’ - penetrate them in MD. Some kebabs
are linked up with neighboring kebabs. Although only shishes are drawn in
are linked up with neighboring kebabs. Although only shishes are drawn in
Figure 6.10 as linkages in MD, it is considered that, in reality, there also exist tie
molecules which link the kebabs in MD. Small and non-uniform lamellae whose
a*-axes are parallel to MD pile epitaxially on the c-axis-oriented lamellae which
are a component (kebabs) of the shish-kebab main skeleton structure. The a*-
axis-oriented imperfect lamellae are rather larger in quantity than th£ c-axis-
oriented perfect lamellae. The amount of shishes is about 5.3%

Effect of molding conditions 175 c-axis MD a*-axis T"ffl t; 160A ...I in Figure
6.10 Crystalline structure of skin layer in injection-molded PP. the total
crystalline structure. Amorphous chains fill the spaces of the crystalline
structure. The amount of the amorphous chains is about half of the total volume.
The 0*-axis-oriented component acts as a morphological plasticizer at the time
of bending and is the origin of good hinge of injection-molded PP. It is
considered that the skin layers of injection moldings of semicrystalline polymers
other than PP are composed of a pure shish-kebab structure without the a*-axis-
oriented component. The core layer of injection-molded PP is composed of
spherulites whose crystallites have weak c-axis and a*-axis orientations to MD.
6.3 EFFECT OF MOLDING CONDITIONS ON HIGHER ORDER
STRUCTURES OF INJECTION-MOLDED PP Figure 6.11 shows the effects of
injection molding conditions on the thickness of the skin layer [4]. Cylinder
temperature has the greatest effect, followed by injection speed, injection
pressure and mold temperature. These last two conditions have practically no
influence on the thickness of the skin layer in the range tested. The lower the
cylinder temperature and injection speed, the thicker the skin layer. Accordingly,
the following experiments were carried out, changing only the cylinder
temperature and keeping all other conditions constant.

176 Higher order structure of injection-molded polypropylene (a) < b ) 0.4. 0.2
0) c o -0 O- Sample-M I I I I I 200 400 600 Injection pressure (kg/cm2) (c) id )
Figure 6.11 Dependence of thickness of skin layer on molding conditions: (a)
cylinder temperature, (b) injection speed, (c) injection pressure and (d) mold
tem¬ perature. 6.4 EFFECT OF MOLECULAR MASS ON HIGHER ORDER
STRUCTURES Figure 6.12 shows changes in the crystalline texture observed
with a polarizing microscope, and wide-angle X-ray diffraction pattern of a
flexural test specimen injection-molded from a PP sample [5] at different
cylinder temperatures. The polarizing micrographs show clear skin-core
structures. The thickness of the skin layer decreases with increasing cylin¬ der
temperature. Also, c-axis and a*-axis mixed orientations are seen in the wide-
angle X-ray diffraction patterns, and the degree of orien^ tion of crystalline
angle X-ray diffraction patterns, and the degree of orien^ tion of crystalline
molecular chains decreases with increasing cylind^t; temperature.

Effect of molecular mass 177 Sample - B 1 mm Surface ·§ £ 3 ND T O-TD — a)


a> MD 55 Center MD T O- ND ►TD Cylinder Temp. 200°C 240°C 280°C
320°C Figure 6.12 Changes with cylinder temperature of polarized micrograph
ar wide-angle X-ray diffraction pattern of injection-molded PP. Figure 6.13a and
b show the thickness of the skin layer and the crystalline orientation functions as
a function of cylinder temperature [5]. The mole¬ cular mass is highest for
sample A and lowest for sample F. The thickness of the skin layer decreases
with increasing cylinder temperature, and comparison at the same cylinder
temperature indicates that the thickness decreases with decreasing molecular
mass. The absolute values of the c-axis orientation function fc and ft-axis
orientation function fb decrease ( a ) ( b ) Figure 6.13 Dependences of (a)
thickness of skin layer and (b) crystalline orienta¬ tion functions of injection-
molded homo PPs with various molecular masses on cylinder temperature.

178 Higher order structure of injection-molded polypropylene with increasing


cylinder temperature, and comparison at the same cylinder temperature indicates
that these values decrease with decreasing molecular mass. The a*-axis
orientation function fa* increases somewhat with increas¬ ing cylinder
temperature, and does not depend on molecular mass. Several researchers have
analyzed molecular orientation process in injec¬ tion molding from the
viewpoint of growth of a melt orientation at the gate and its relaxation in the
cavity [5] (see Section 6.9). According to this theory, the higher degree of
molecular orientation seen at lower cylinder tempera¬ tures is interpreted as
follows: the lower the cylinder temperature, the lower the injected resin
temperature, which causes a greater degree of melt orientation, resulting in a
longer relaxation time. In addition, crystallization takes place in a shorter time,
which ‘freezes’ the melt orientation in the molding. The higher degree of
molecular orientation associated with higher molecular masses is because the
higher the molecular mass, the more pronounced is the degree of melt
orientation and the relaxation time be¬ comes longer. Figure 6.14 shows the
changes of the c-axis orientation function fc in the flow direction [5]. fc
decreases on going away from the gate. This is because the flowing time of an
injected molten resin increases on going away from | Distance from Gate (cm ) j
Entrance End Figure 6.14 Distribution in flow direction, of crystalline c-axis
orientation function fc of injection-molded homo PPs with various molecular
masses.

Effect of molecular mass 179 the gate, which promotes relaxation of the melt
Effect of molecular mass 179 the gate, which promotes relaxation of the melt
and thus disorientation in the cavity. Figure 6.15 shows the changes of fc in the
thickness direction [5]. Here, H is half of the thickness of the specimen and y is
the distance from the center. fc is higher nearer the surface and decreases toward
the interior. This is because the relaxation of melt orientation is less nearer the
surface because the molten resin solidifies sooner. A peak and a shoulder are
observed in the middle region for samples E and D, respectively. These are
assumed to be caused by secondary flow during the cooling and pressure holding
process. When injection molding is carried out under a high holding pressure,
the molten resin in the cavity solidifies progressively from the surface to the
interior and the volume of the resin is reduced; excess molten resin flows into
the still molten inner region under the action of holding pressure. This secondary
flow is finished when the gate is cooled and sealed. Although this secondary
flow is slow, it occurs at low temperatures just above the solidification
temperature. This process results in a high melt orientation, long relaxation time
of the melt and short solidification time. Therefore, this ‘slight’ flow process is
considered to cause high molecular orientation. Figure 6.16 shows the changes
of crystallinity Xc in the thickness direction [5]. Xc is low at the surface region
where the cooling rate is high and increases toward the interior where the
cooling rate is low. Xc is higher for lower molecular masses and when the
cylinder temperature during process¬ ing is higher. Figure 6.17 shows the
changes of /^-crystal content, the K value [6], in the thickness direction [5]. /J-
crystals exist at the surface region of about twice | y/H f Surface Center Figure
6.15 Distribution in thickness direction of crystalline c-axis orientation function
fe of injection-molded homo PPs with various molecular masses.

180 Higher order structure of injection-molded polypropylene Surface Center


Figure 6.16 Distribution in thickness direction of crystallinity Xc of injection-
molded homo PPs with various molecular masses. the thickness of the skin layer
and do not appear in the inner region. The β-crystal content is higher as the
molecular mass is increased and the cylinder temperature is decreased. An
injection-molded PP shows the mixed c-axis and a*-axis orientation shown in
Figure 6.12. Here, we will evaluate the proportions of the c-axis- oriented
component and a*-axis-oriented component. Figure 6.18 shows the 20 scan
curve and the (110) plane and (0 4 0) plane reflection azimuthal scan curves of
wide-angle X-ray diffraction of a flexural test specimen molded from sample B
at a cylinder temperature of 240 °C [5]. A baseline (BL2) is drawn horizontally
at the bottom of the azimuthal scan curve of the (1 1 0) reflection. The area
around an azimuthal angle of 0° above the baseline is taken as C and the area
around an azimuthal angle of 90° above the baseline is taken as A*. The c-axis-
around an azimuthal angle of 90° above the baseline is taken as A*. The c-axis-
oriented component fraction [C] and the fl*-axis-oriented component fraction
[A*] are defined as follows: [C]-C+CA< (1) A* = C +A* (2)

Effect of copolymerization 181 Although there is no assurance that the absolute


amounts of the c-axis- oriented and a*-axis-oriented components can be
rigorously evaluated by [C] and [A*], a relative comparison may be made when
raw resin and/or molding conditions are changed. Figure 6.19 shows the
dependence of the a*-axis-oriented component fraction [A*] on cylinder
temperature [5]. [A*] is higher than 0.7 for all samples, which means that the a*-
axis-oriented component fraction is overwhelmingly greater than the c-axis-
oriented component fraction in injection-molded PPs. [A*] is higher as
molecular mass is lower and cylinder temperature is higher. That is, the 0*-axis-
oriented component compared to the c-axis-oriented component increases with
decreasing mole¬ cular mass and with increasing cylinder temperature. 6.5
EFFECT OF COPOLYMERIZATION Figure 6.20 shows the dependence of the
thickness of the skin layer of injection-molded PP copolymers with ethylene on
the cylinder temperature [7]. Each sample has a similar molecular mass. H, R,
and B indicate homo, random, and block copolymers, respectively, and the
numbers indicate the ethylene content. For example, R21 indicates a random PP
copolymer with

182 Higher order structure of injection-molded polypropylene (040) 90(TD)


Azimuthal Angle (°) BL2 BL1 0(MD) Figure 6.18 20 scan curve, (110) plane
reflection azimuthal scan curve and (040) plane reflection azimuthal scan curve
of wide-angle X-ray diffraction of injection- molded PP. an ethylene content of
2.1 wt%. The thickness of the skin layer is decreased by copolymerization with
ethylene. Figure 6.21 shows the crystalline orientation functions as a function of
the cylinder temperature [7]. The absolute values of fc and fb decrease with
increasing cylinder temperature and are decreased b> copolymerization with
ethylene. The a*-axis orientation func- tions fa* of the homo PP and block
copolymers increase with increasing cylinder tempera¬ ture, while those of the
random copolymers are not influenced by cylinder temperature. The random
copolymers show higher fa+ than the homo PP, anc the block copolymers
conversely show lower fa+ than the homo PP. The fac that the degrees of
molecular orientation, such as the thickness of the skir layer and /c, are
decreased by copolymerization with ethylene is due to the narrowing in the
molecular mass distribution, which decreases tho

Addition of crystallization nucleators 183 200 320 240 280 Cylinder


Temperature ( °c ) Figure 6.19 Dependence of a*-axis-oriented component
Temperature ( °c ) Figure 6.19 Dependence of a*-axis-oriented component
fraction [A*] of injection molded homo PPs with various molecular masses on
cylinder temperature. Figure 6.20 Dependence of thickness of skin layer of
injection-molded PP copolymers with ethylene on cylinder temperature. melt
orientation and its relaxation time, and due to the reduction in the crystallization
temperature, particularly for random copolymers [8]. The crystallinity Xc
decreases with increasing ethylene content [7]. The degree of the decrease is
higher for the random copolymers than for the block copolymers. Block
copolymerization with ethylene considerably de¬ creases the β-crystal content.
No β-crystals were found in the random

184 Higher order structure of injection-molded polypropylene 0.2L^- . ■ 1_ 200


240 280 320 Cyl inder Temperature ( °C ) Figure 6.21 Dependence of crystalline
orientation functions of injection-molded PP copolymers with ethylene on
cylinder temperature. copolymer moldings [9]. The a*-axis-oriented component
fraction [A*] is increased by copolymerization with ethylene. This increment is
higher for the random copolymers than for the block copolymers [7]. 6.6
EFFECT OF ADDITION OF CRYSTALLIZATION NUCLEATORS Figure
6.22 shows the dependences of the crystalline orientation functions on cylinder
temperature in injection moldings of PPs containing crystalliza¬ tion nucleators
[10]. CC, TC, BA and GA are PPs containing 0.5 wt% calcium carbonate, talc,
aluminum salt of benzoic acid and p-dimethyl- benzylidene sorbitol (Gelall
MD), respectively. The order of crystalliza¬ tion temperature is
GA>BA>TC>CC>PP. fc increases with increasing crystallization temperature.
This is because the higher the crystallization

Effect of particulate filling 185 Figure 6.22 Dependence of crystalline


orientation functions of injection moldings of PPs containing crystallization
nucleators on cylinder temperature. temperature, the shorter the time until
crystallization, which allow less relaxation in melt orientation [10]. The
crystallinity Xc is slightly increased by the addition of crystallization nucleators,
and the /^-crystal content and a*-axis-oriented component frac¬ tion [A*] are
decreased [10,11]. 6.7 EFFECT OF PARTICULATE FILLING Figure 6.23
shows the dependence of the thickness of the skin layer of injection moldings of
particulate-filled PPs on cylinder temperature [12]. TC and CC indicate talc and
calcium carbonate, respectively, and the numbers refer to the content (wt%). The
thickness of the skin layer is increased by the filling. A similar tendency is
observed for fc. The fact that the degree of molecular orientation is increased by
particulate filling is due
186 Higher order structure of injection-molded polypropylene Figure 6.23
Dependence of thickness of skin layer of injection moldings of particu¬ late-
filled PPs on cylinder temperature. to increase in thermal diffusivity and
crystallization temperature, which lead to faster cooling and more rapid
solidification. Consequently, the relaxation of the melt orientation is hindered
[12]. The crystallinity Xc is unaffected by CC and considerably increased by TC
filling [12]. The /^-crystal content is increased by CC and decreased by TC
filling [12]. The fl*-axis-oriented component fraction [A*] is not influenced by
CC, whereas it is decreased by TC filling [12]. Figure 6.24 shows wide-angle X-
ray diffraction patterns taken from various directions of flexural test specimens
injection-molded from unfilled PP and PP filled with 10wt% talc (TC-10) at a
cylinder temperature of 240 °C [13]. Although the TC-10 specimen shows the
same mixed c-axis and a*-axis orientation as the PP specimen in THRU, its
(040) plane reflection is very weak. Contrary to this, in EDGE and END, the
TC-10 specimen shows very strong (04 0) plane reflection on the equator
(thickness direc¬ tion). Figure 6.25 shows the crystal orientation state of
injection molding of talc-filled PP, which is assumed from the above results. It
shows a peculiar

Effect of glass fiber filling 187 THRU EDGE END TC-10 240° C THRU PP
240° C END MD M D TD L N D Figure 6.24 Wide-angle X-ray diffraction
patterns taken from various directions of injection moldings of PP filled with 10
wt% talc (TC-10) and of PP (PP). Figure 6.25 Crystal orientation state in
injection molding of talc-filled PP. crystal orientation in which the planes of the
talc particle plates are aligned parallel to the molding surface, the c- and a*-axes
of the PP crystals are bimodally oriented in the flow direction, and the b-axes are
oriented in the thickness direction. The degree of the bimodal orientation of the
c- and a*-axes decreases toward the interior of the molding. The h-axis
orientation is strong throughout the thickness direction, although it is a little
weaker at the surface skin and central regions. Since this peculiar crystal
orientation already appears with a very small amount of TC filling (0.5 wt%)
[13] and TD ND b-axis b -axis

188 Higher order structure of injection-molded polypropylene the b-axis


orientation to the thickness direction is restored even when melt- recrystallized
[14,15], it is assumed that talc has a remote influence on the crystallization of PP
so that the ft-axes are oriented perpendicular to the planes of the talc particle
plates [14,16]. Injection moldings of PPs filled with glass flake or mica, which
are platy particles like talc, do not show this peculiar crystal orientation [16].
Likewise, injection moldings of talc-filled high density polyethylene and
Likewise, injection moldings of talc-filled high density polyethylene and
poly(butene-l) do not show any peculiar crystal orientation [17]. Only talc-filled
PP injection moldings show the peculiar crystal orientation. 6.8 EFFECT OF
GLASS FIBER FILLING Since injection-molded PPs show mixed c-axis and
a*-axis orientations, the c-axis-oriented component and the a*-axis-oriented
component cancel each other in the calculation of the crystalline orientation
functions, and hence it is not very appropriate to evaluate the degree of
crystalline orientation from the orientation functions. Furthermore, in the case of
/J-crystal-rich injection moldings, since the (3 0 0) plane reflection of the jS-
crystals appears near the (0 4 0) plane reflection of the α-crystals, the crystalline
orientation functions cannot be calculated accurately. In such cases, it is
convenient to use the crystalline orientation fraction (OF) as a measure of the
degree of crystalline orientation [10,18]. OF is calculated by: OF,_g±y_ C +A*
+ N (3) where N (non-oriented) is the area enveloped by BL2 and BL1, which is
the baseline of the 20 scan curve, in the azimuthal scan curve of the (1 1 0) plane
reflection in Figure 6.18. Figure 6.26 shows the dependences of the crystalline
OF of glass fiber- reinforced PP injection moldings (FRPP) on cylinder
temperature [19]. GF indicates glass fiber and the number the content of GF
(wt%). GF-20M is a practical compound FRPP which contains silane-treated GF
and maleic anhydride grafted PP in order to improve the adhesion between GF
and PP. OF is increased by GF filling, because the melt orientation and thermal
diffusivity are increased by the filling. While the crystallinities Xc of the
untreated FRPPs do not differ from that of the unfilled PP, Xc of the practical
compound FRPP is considerably higher than that of the unfilled PP. While the ^-
crystal content of the practical compound FRPP agrees with that of the unfilled
PP, those of the untreated FRPPs are much higher than that of the unfilled PP.
While the a*-axis-oriented component fraction [A*] of the practical FRPP
changes only slightly from that of the unfilled PP, it decreases with increasing
GF content for the untreated FRPP injection moldings.

Analysis of molecular orientation process 189 Cylinder Temperat ure (°C )


Figure 6.26 Dependence of crystalline orientation fraction OF of injection mol¬
dings of glass fiber-reinforced PPs on cylinder temperature. 6.9 ANALYSIS OF
MOLECULAR ORIENTATION PROCESS IN INJECTION MOLDING The
injection molding process is an exceedingly complex phenomenon involving an
unsteady flow and heat conduction, and, in the case of semi¬ crystalline
polymers, a non-isothermal crystallization. But essentially, the orientation of
molecular chains in injection moldings can be regarded as flow-induced
orientation which has subsequently been frozen-in by cooling and solidification,
and the rheological and thermal properties of polymers play important roles in
and the rheological and thermal properties of polymers play important roles in
the orientation process. Analysis of the orientation process of molecular chains
in injection molding has been the subject of research for many years, and
recently the analysis has reached the stage of quantitative treatments [20].
Analyses of the orientation process of mole¬ cular chains in injection molding
are mostly undertaken for amorphous polymers not involving the complex
process of crystallization, and these analyses are made from the viewpoint of
stress growth during flow in the cavity and stress relaxation after cessation of
flow. The authors regarded recoverable shear strain yc as a measure of melt
orientation and analyzed the molecular orientation process in injection molding
of PPs from the viewpoint of growth of a recoverable shear strain at the gate and
its relaxation in the cavity [4, 5,8,10-12,21-23]. In injection molding, a molten
resin passes through the sprue, runner and gate, fills the cavity and becomes a
shaped article through cooling and solidification. In the meantime, the extent of
melt orientation changes at the different parts. In general molds, the cross-
section of the flow path at the gate is extremely small, compared with the sprue,
runner and cavity, and the

190 Higher order structure of injection-molded polypropylene flow rate is


constant through the flow path. Therefore, the shear rate at the gate is extremely
high and most of the melt orientation takes place during passage through the
gate. As shown in Figure 6.27, the recoverable shear strain having grown during
passage through the gate, ye0, undergoes a Maxwell-type relaxation with a
relaxation time λx while flowing through the cavity and becomes yel. At this
time, cooling is ignored until the resin completely fills the cavity, and the resin
temperature is regarded as identical to the injected resin temperature ^melf
Therefore, it is regarded that in the meantime a relaxation occurs with the
relaxation time t) at the injected resin temperature Tmelf Then, the resin at a
certain position xx from the gate is cooled by the mold at the temperature Tmold
and its temperature becomes T(t). The recoverable shear strain yel undergoes a
temperature- dependent Maxwell-type relaxation with a relaxation time λ2(Τ(ί))
and becomes a residual recoverable shear strain yer when the resin temperature
reaches the crystallization temperature Tc. yer is assumed to be proportional to
the degree of orientation of an injection-molded article. y t |< *1 * ^rnold !
TmeltJ—' \ T r] —>x 2H K i ΐ Tmold \ Entrance Flow Front VeO > Ve1 λ! (Tm
e 11 ) \ X2(T(t)) ^er Figure 6.27 Coordinates and movement of recoverable shear
strain ye in mold cavity. Here, the injection molding of an ASTM flexural test
specimen shown in Figure 6.3 will be considered. With injection rate V9 cavity
width B, and cavity thickness 2H, yel is expressed as follows: [-2 BHXl] (4)
Here, the Carslaw-Jaeger equation for one-dimensional nonsteady-state thermal
Here, the Carslaw-Jaeger equation for one-dimensional nonsteady-state thermal
conduction of infinite solid [24] is applied, ignoring the latent heat

Analysis of molecular orientation process 191 of crystallization. Then, the time


change of the resin temperature T(t) at y(0^y^H) from the center in the thickness
direction is expressed as fol¬ lows: T—Tn Tme\t — Tn mold mold = 2 Σ (-1)M
cosi(„+l/2)|^x „=o π(η+1/2) χβχρ|-(η+1/2)2π2-^ (5) where a is the thermal
diffusivity. If Equation 5 is rewritten in the form of t=f(T\ the residual
recoverable shear strain yer is calculated as follows: In —= Vel ic J T melt 1
df(T) λ2(Τ) d T (6) Half of the cavity thickness, H, is divided into 10 equal parts,
the residual recoverable shear strain at the center of each division, yerti, is
calculated by Equation 6, and the mean residual recoverable shear strain yer is
calculated as follows: 1 10 ^er=To ^er,i ^ The material characteristics needed to
calculate yer are the initial recover¬ able shear strain yeo> its relaxation times
λΐ9λ2, the thermal diffusivity a and the crystallization temperature Tc. yer is
higher as ye0 is higher, λχ and λ2 are longer, a is higher and Tc is higher. ye0
can be obtained from the end correction coefficient in capillary flow properties.
λγ and λ2 can be obtained from the characteristic relaxation time λ0 of the
absolute value of complex viscosity \η*(ω)\. Data are available in the literature
for a. Tc can be measured with DSC. Figure 6.28 shows how the c-axis
orientation function /c, the thickness of the skin layer, ST, and the mean residual
recoverable shear strain yer calculated from Equation 7 change in the flow
direction for a flexural test specimen molded from sample B at a cylinder
temperature of 240 °C [5]. fc and ST decrease on going away from the gate and
the degrees of these decreases are nearly the same as that of yer> thus
underscoring the validity of this analysis. Similar results were obtained for
specimens molded from other samples at other cylinder temperatures. Figure
6.29 illustrates the changes in the thickness direction of fc and the relative
residual recoverable shear strain yeT/yei calculated from Equation 6 for a
specimen molded from sample B at a cylinder temperature of 240 °C [5]. The
two values agree well, thus attesting the propriety of this analysis.

192 Higher order structure of injection-molded polypropylene Figure 6.28


Variations in flow direction of crystalline c-axis orientation function /c,
thickness of skin layer, ST, and calculated mean residual recoverable shear strain
yer of injection-molded PP. Figure 6.29 Variations in thickness direction of
crystalline c-axis orientation func¬ tion fc and calculated relative residual
recoverable shear strain yer/ye\ of injection molded PP.

Properties of injection-molded PPs 193 Figure 6.30 Relation between crystalline


c-axis orientation function fc and cal¬ culated residual recoverable shear strain
c-axis orientation function fc and cal¬ culated residual recoverable shear strain
yer at various positions in the thickness direction of flexural specimens
injection-molded from homo PPs with various molecular masses at various
cylinder temperatures. Figure 6.30 shows the relation between /c and calculated
residual recoverable shear strain yer at various positions in the thickness direc¬
tion of specimens molded from various samples at various cylinder tem¬
peratures [5], There are 65 data points. A fairly high correlation with a
correlation coefficient r = 0.871 exists between the two approaches. Figures 6.31
and 6.32 show the relation between, respectively, the thick¬ ness of the skin
layer and fc and calculated mean residual recoverable shear strain yer, at the
centers of flexural test specimens molded from various samples at various
cylinder temperatures [5]. Again, fairly high correlations exist regardless of the
kind of resin (molecular mass) and molding conditions (cylinder temperature).
As mentioned above, this theory can considerably well describe not only the
mean molecular orientation but also its changes in the flow and thick¬ ness
directions. This theory is applicable not only to homo PPs [5] but also to PP
copolymers with ethylene [8], PPs with nucleating additives [10,11], particulate-
filled PPs [12,23], and glass fiber-reinforced PPs [19].

194 Higher order structure of injection-molded polypropylene Figure 6.31


Relation between thickness of skin layer and calculated mean residual
recoverable sheat strain yer of flexural specimens injection-molded from homo
PPs with various molecular masses at various cylinder temperatures. 6.10
PROPERTIES OF INJECTION-MOLDED PPS Properties such as flexural
modulus, flexural strength, Izod impact strength, heat distortion temperature and
mold shrinkage of injection-molded homo PPs increase as the molecular mass
increases and the cylinder temperature decreases since the degree of molecular
orientation is higher [22]. The flexural modulus, flexural strength and heat
distortion temperature of injection molded PP copolymers with ethylene
decrease with increas¬ ing ethylene content, and compared at the same ethylene
content, the decrement is more pronounced for random copolymers than for
block copolymers [7]. The Izod impact strength increases with increasing
ethylene content, and the improvement is stronger for block copoly¬ mers than
for random copolymers. The mold shrinkage is decreased by random
copolymerization and slightly increased by block copoly¬ merization.

Properties of injection-molded PPs 195 Figure 6.32 Relation between crystalline


c-axis orientation function fc and calculated mean residual recoverable shear
strain yer of flexural specimens injection-molded from homo PPs with various
molecular masses at various cylinder temperatures. By the addition of
molecular masses at various cylinder temperatures. By the addition of
crystallization nucleators, the flexural modulus, flexural strength, heat distortion
temperature and mold shrinkage are increased and the Izod impact strength is
decreased. Figure 6.33 shows the effects of filler content and cylinder
temperature on the flexural modulus, flexural strength and mold shrinkage of
particulate- filled PP injection moldings [18]. TC and CC indicate talc and
calcium carbonate, respectively. The flexural modulus increases and the mold
shrink¬ age decreases with increasing filler content. The flexural strength shows
a maximum at a filler content of about 20 wt%. This is assumed to be caused by
a balance between an increase in rigidity and a decrease in elongation with
increasing filler content. In the case of TC filling, an abrupt increase in flexural
strength is seen at a TC content of 0.5 wt% owing to the crystalliza¬ tion
nucleating effect. Comparing the results at the same filler content level, the
above properties decrease with increasing cylinder temperature. This is because
the degree of molecular orientation decreases with increasing cylinder
temperature, as shown later. The flexural modulus, Izod impact strength and heat
distortion tempera¬ ture are largely increased, the flexural strength is slightly
increased, and the

Flexural Modulus (1o4/cm2) Figure 6.33 Dependences of (a) flexural modulus,


(b) flexural strength and (c) mold shrinkage of injection moldings of particulate-
filled PPs on filler content (104kg/cm2= 1 GPa; 1 kg/cm2 = 0.1 MPa). Filler
Content (wt%) (a) Filler Content (wt7.) (b) Filler Content (wt%) (c) Flexural
Strength (kg/cm2) Mold Shrinkage ( % ) 200°C 240°C 280°C 320°C Cylinder
Temperature 200 C 240eC 280°C 320°C Cylinder TemDerature

Structure-property relationships 197 mold shrinkage is strongly decreased by


glass fiber filling [19]. While the flexural strength, Izod impact strength and heat
distortion temperature of practical compound FRPP are much higher than those
of untreated FRPP, the flexural modulus and mold shrinkage of the practical
compound FRPP scarcely vary from those of the untreated FRPP. Injection
moldings show an anisotropy of properties owing to the orien¬ tations of
molecular chains or filled fibers. Figure 6.34 shows the anisotropy ( a ) (b)
Figure 6.34 Variations with angle to flow direction (MD) of flexural modulus
and flexural strength of injection moldings of (a) PP and (b) glass fiber-
reinforced PP (104 kg/cm2 = 1 GPa; 1 kg/cm2 = 0.1 MPa) of flexural modulus
and flexural strength of square plates molded from a PP and an FRPP [25]. The
solid lines show the calculation results by Hearmon’s equation [26]: _1 E{ cos4
Θ sin4 Θ / 4 1 Eo Ego \F4 5 E0 Egot sin2 Θ cos2 Θ (8) where Εθ is the flexural
modulus in the direction of 0° from the flow direction (MD) and Ee, E45 and
E90 are flexural moduli in the directions of 0°, 45° and 90° from MD,
E90 are flexural moduli in the directions of 0°, 45° and 90° from MD,
respectively. The calculated values fit extremely well with experimental ones.
The flexural modulus and strength of injection-molded PP are lowest in the 45°-
direction, which is assumed to be due to the shear deformation between lamellae
or crystalline mole¬ cular chains since the shear stress is the highest in the 45°-
direction. The flexural modulus and strength of injection-molded FRPP
gradually decrease from MD to TD.

198 Higher order structure of injection-molded polypropylene 6.11


STRUCTURE-PROPERTY RELATIONSHIPS Figure 6.35 shows the relations
between the flexural modulus, flexural strength and mold shrinkage and the
thickness of the skin layer of injec¬ tion moldings of particulate-filled PPs [18].
TC and CC indicate talc- and calcium carbonate-filled PPs, respectively, and the
numbers show the filler content (wt%). Since injection-molded PPs show skin-
core two-phase structures, as shown in Figure 6.2, the following relation holds
for the flexural modulus [4]: E = Es-{Es~fc\b-2t)(h-2t)3 (9) where E is the
flexural modulus of the whole, Es and Ec are the flexural moduli of,
respectively, the skin and core layers, b and h are respectively the width and
thickness of the specimen, and t is the thickness of the skin layer. A linear
relationship can be obtained if E is plotted against (b — 2t)(h — 2t)3 with
respect to specimens with different values of t prepared at various cylinder
temperatures, Es can be obtained from the E-axis intercept and Ec can be
obtained from the slope. Figure 6.35a shows such plots. Es increases with
increasing filler content, and compared at the same filler content, Es of injection
molding of flaky talc-filled PP is higher than that of spherical calcium carbonate-
filled PP. As shown in Figure 6.35b, there exist different rectilinear relationships
with positive slopes between the flexural modulus and the thickness of the skin
layer at each filler content level, and comparing at the same thickness of the skin
layer, the flexural strength of talc-filled PP injection molding is higher than that
of calcium carbonate-filled one. As shown in Figure 6.35c, there also exist
different rectilinear relationships with positive slopes between the mold
shrinkage and the thickness of the skin layer at each filler content level, and
comparing at the same thickness of the skin layer, the mold shrinkage of talc-
filled PP injection molding is lower than that of calcium carbonate-filled PP.
From these results, it can be stated that the thicker the skin layer, the higher the
flexural modulus, flexural strength and mold shrinkage. Figure 6.36 shows the
relations between the flexural modulus and strength and /c [22]. Between them,
there exist different downward-curved relation¬ ships with positive slopes at
each sample. This means that the flexural modulus and strength cannot be
treated only as a function of the molecular orientation, i.e. without considering
treated only as a function of the molecular orientation, i.e. without considering
the sample characteristics. Comparing the moduli of the samples at the same /c,
the sample with the lower molecular mass exhibits higher flexural modulus and
strength. This is assumed to be because the sample with the lower molecular
mass shows higher crystallinity, as shown in Figure 6.16. Figures 6.37 and 6.38
show the relations between the tensile yield strength and necking stress in each
direction of injection-molded PP rectangular

(b-2t)(h-2t) (cm ) Thickness of Skin Layer (mm) Thickness of Skin Layer (mm)

200 Higher order structure of injection-molded polypropylene Figure 6.36


Relations between (a) flexural modulus and (b) flexural strength, and crystalline
c-axis orientation function fc of injection-molded homo PPs with various
molecular masses (104 kg/cm2 = 1 GPa; 1 kg/cm2 = 0.1 MPa). Figure 6.37
Relation between yield strength in various directions and thick¬ ness of skin
layer of rectangular plates injection-molded from homo PPs (1 kg/cm2 = 0.1
MPa).

Appendix 201 Thickness of Skin Layer (mm) Figure 6.38 Relation between
necking stress in various directions and thickness of skin layer of rectangular
plates injection-molded from homo PPs (1 kg/cm2 = 0.1 MPa). plates, and the
thickness of the skin layer [25]. The yield strength is in the order of
MD>TD>45° and the necking stress is in the order of MD>45°>TD. The yield
strength and necking stress in each direction correlate linearly with the thickness
of the skin layer, and the molding with the thicker skin layer shows higher
anisotropy. When the thickness of the skin layer is extrapolated to zero, the yield
strength and necking stress in each direction meet at nearly the same points,
which means that an injection molding without the skin layer shows little
anisotropy. 6.12 CONCLUSIONS The effects of molding conditions and
characteristics of the raw polymer on the structures and properties of injection-
molded PPs were described. The structure-property relationships may be used as
a data base for the quality control and design of injection-molded PPs items.
Although these results were obtained on simple shape specimens, the approach
adopted may also be applied to complex shape moldings. The results
demonstrated for PPs may also be applicable, to some extent, to other
semicrystalline polymers. 6.13 APPENDIX: SYMBOLS AND
ABBREVIATIONS A Amorphous region A* Reflection area caused by a*-axis-
oriented component in (1 1 0) plane azimuthal scan

ligher order structure of injection-molded polypropylene a*-axis-oriented


component fraction American Society for Testing and Materials Block PP
component fraction American Society for Testing and Materials Block PP
copolymer with ethylene Width of cavity Width of specimen Aluminum salt of
benzoic acid Baseline of 20-scan curve Baseline of (1 10) plane azimuthal scan
curve Crystalline region Reflection area caused by c-axis-oriented component in
(1 1 plane azimuthal scan c-axis-oriented component fraction Specific heat
Calcium carbonate Differential scanning calorimeter Flexural modulus Flexural
modulus of core layer Flexural modulus of skin layerk Flexural modulus in 0°-
direction from MD Flexural modulus in 45 °-direction from MD Flexural
modulus in 90°-direction from MD Flexural modulus in 0°-direction from MD
Tensile storage modulus Tensile loss modulus End view Edge view Crystalline
A*-axis orientation function Crystalline b-axis orientation function Crystalline c-
axis orientation function Flexural modulus Flexural strength Glass fiber-
reinforced PP p-dimethyl-benzylidene sorbitol (Gelall MD) Glass fiber Homo
PP Half of cavity (specimen) thickness Thickness of specimen Heat of fusion /^-
crystal content Machine direction (flow direction) Molecular weight Molecular
weight distribution Reflection area caused by non-oriented component in (11(
plane azimuthal scan

References 203 n Summand ND Normal direction (thickness direction) OF


Crystalline orientation fraction PP Polypropylene R Random PP copolymer with
ethylene r Correlation coefficient ST Thickness of skin layer T Resin
temperature Tc Crystallization temperature Tg Glass transition temperature Tm
Melting point Tmeit Injected resin temperature Tmoid Mold temperature t Time
t Thickness of skin layer TC Talc TD Transverse direction THRU Through view
tan δ Loss tangent V Injection rate Xc Crystallinity Xi Distance from gate in
flow direction y Distance from center of cavity (specimen) a Thermal diffusivity
ye Recoverable shear strain ye0 Initial recoverable shear strain yel Recoverable
shear strain at position χγ from gate yer Residual recoverable shear strain yeTti
Residual recoverable shear strain of i-th division in thickness direction e
Dielectric constant \η*\ Absolute value of complex viscosity Θ Bragg angle Θ
Angle to MD λ0 Characteristic relaxation time λ i Relaxation time of recoverable
shear strain while flowing through cavity λ2 Relaxation time of recoverable
shear strain after cessation of flow p Density 6.14 REFERENCES 1. Fujiyama,
M. and Wakino, T. (1988) Journal of Applied Polymer Science, 35, 29-49.

204 Higher order structure of injection-molded polypropylene 2. Takayanagi,


M., Imada, K. and Kajiyama, T. (1966) Journal of Polymer Science, Part C, 15,
263-81. 3. Keller, A. and Machin, M. J. (1967) Journal of Macromolecular
Science, Bl, 41-91. 4. Fujiyama, M. and Kimura, S. (1975) Kobunshi
Ronbunshu, 32, 581-90. 5. Fujiyama, M. and Wakino, T. (1991) Journal of
Ronbunshu, 32, 581-90. 5. Fujiyama, M. and Wakino, T. (1991) Journal of
Applied Polymer Science, 43, 57-81. 6. Turner-Jones, A., Aizlewood, J. M., and
Beckett, D. R. (1964) Makromolekulare Chemie., 75, 134-58. 7. Fujiyama, M.
and Wakino, T. (1991) Seikei-Kakou, 3, 217-24. 8. Fujiyama, M. and Wakino,
T. (1991) Seikei-Kakou, 3, 301-7. 9. Fujiyama, M. and Wakino, T. (1991)
Seikei-Kakou, 3, 225-32. 10. Fujiyama, M. and Wakino, T. (1991) Journal of
Applied Polymer Science, 42, 2739-47. 11. Fujiyama, M. and Wakino, T. (1991)
Journal of Applied Polymer Science, 42, 2749-60. 12. Fujiyama, M. and
Wakino, T. (1991) Journal of Applied Polymer Science, 43, 97-128. 13.
Fujiyama, M. and Wakino, T. (1991) Journal of Applied Polymer Science, 42, 9-
20. 14. Rybnikaf, F. (1989) Journal of Applied Polymer Science, 38, 1479-90.
15. Fujiyama, M. (1992) International Polymer Processing, 7, 165-71. 16.
Fujiyama, M. (1992) International Polymer Processing, 7, 358-73. 17. Fujiyama,
M., unpublished data. 18. Fujiyama, M., Kawasaki, Y., and Wakino, T. (1987)
Nihon Reoroji Gakkaishi, 15, 191-202. 19. Fujiyama, M., (1993) International
Polymer Processing, 8, 245-54. 20. Fujiyama, M. (1989) Nihon Reoroji
Gakkaishi, 17, 5-12. 21. Fujiyama, M. and Kimura, S. (1978) Journal of Applied
Polymer Science, 22, 1225-41. 22. Fujiyama, M. (1986) Nihon Reoroji
Gakkaishi, 14, 152-66. 23. Fujiyama, M., Kawasaki, Y., and Wakino, T. (1987)
Nihon Reoroji Gakkaishi, 15, 203-9. 24. Carslaw, H. S. and Jaeger, J. C. (1957)
Conduction of Heat in Solids, Oxford University Press, London. 25. Fujiyama,
M., Awaya, H., and Kimura, S. (1977) Journal of Applied Polymer Science, 21,
3291-309. 26. Hearmon, R. F. S. (1961) An Introduction to Applied Anisotropic
Elasticity, Oxford University Press, London.

7 Knit-line behaviour of polypropylene and polypropylene-blends* G. Mennig


7.1 INTRODUCTION Knit-lines, sometimes also called weld-lines, occur in
primary processes whenever separated streams of the polymer melt are reunited.
This is true for a continuous process like extrusion where the spider arms of the
mandril of the tube die cause a separation into different melt streams, as well as
for discontinuous processes like compression and injection moulding where
there is a flow obstruction in the cavity of the mould. Such a separation of the
polymer melt into individual streams is not an academic problem, but is very uch
linked to a number of practical cases in processing. For example, Figure 7.1
shows the front bumper, injection moulded from a polypropy- lene/ethylene-
propylene diene monomer (PP/EPDM)-blend, for a mid-size European passenger
car [1]. Also shown is the runner and gate system because such a large technical
part cannot be injected through one gate only. From the number of not less than
20 gates it can be detected that there should be at least 19 knit-lines, excluding
additional knit-lines from other sources. In general, the presence of a knit-line
additional knit-lines from other sources. In general, the presence of a knit-line
has an effect on the surface appearance and the mechanical properties of the
moulded part. Knit-line weakness is attributed to the incomplete fusion of the
reunited flow fronts. In all practical cases the problem is increased by the need
for efficiency in terms of short cooling cycles. The overall result is a region with
different * Dedicated to Gerhard Schenkel on the occasion of his 80th birthday.
Polypropylene: Structure, blends and composites. Edited by J. Karger-Kocsis.
Published in 1995 by Chapman & Hall, London. ISBN 0 412 58430 1

206 Knit-line behaviour of PP and PP-blends Figure 7.1 Injection moulded front
bumper for a passenger car from a PP/EPDM blend. internal stresses and
morphology compared to the regions that are not knit-line infested. With
growing demand for high quality plastic parts the importance of knit-lines is
being increasingly acknowledged. 7.2 ORIGIN OF KNIT-LINES IN PLASTICS
PROCESSING The major sources for the creation of knit-lines are: • multiple
gating • flow obstructions • differences in cross-section of flow channel. Given
the choice between filling a large mould cavity by a single sprue which, after
ejection, has to be removed separately by machining, or filling the same cavity
by a number of pinpoint gates, the latter option is usually selected for economic
reasons. Multiple gating is therefore very common nowadays in injection
moulding. In addition, there may be cases where the quality of a large moulding
may increase by filling through more than one gate because the mean length of
flow within the cavity decreases with the number of gates. There are also cases
of very thin walled moulded parts where the advancing melt front would simply
freeze-in if a single gate were used.

207 Origin of knit-lines in plastics processing tasseOla.mnr FILL TIME CsecD


3.050 0.0 0.187 0.375 0.56? 0.75 0. 937 1. 1?5 1 . 3 1 r’ 1.5 1.687 1.875 ? . 06?
?.?5 ? . 4 3 7 ? . 6?5 ? . 81? 3.0 tasseGi Figure 7.2 Mould filling simulation
(isochrones) for a cup of PP (courtesy of Hoechst AG, Frankfurt). Any hole or
open gap in an injection moulded part has to be created by a flow obstruction
within the cavity. In Figure 7.2 a mould filling simulation is given for a cup. It is
quite obvious from the position of the central sprue underneath that the flow
obstruction forming the handle will create a knit-line exactly where the
maximum load will be in normal use and where, from touching, environmental
stress cracking is likely to occur. In this case, the advancing melt fronts are
meeting head-on, developing into a stagnant situation, whereas in the case of a
flow obstruction, as shown in Figure 7.3, it is believed that the reunited flow
fronts move together for quite some time, thus improving the bond. Multiple
gating and flow obstructions are well known sources of knit- lines in practice,
and a great deal of know-how and experience has been developed to deal with
and a great deal of know-how and experience has been developed to deal with
the resulting knit-lines, or rather the knit-line weakness. This is not the case with
the third major source, i.e. differences in cross-section, an elementary example
of which is shown in Figure 7.4 (see also [2]). As a result of the high flow
resistance in the middle portion of the part, two streams are advancing on both
sides, which will later form a knit-line depending on the ratio of cross-sections
and processing conditions. This is shown by the mould filling simulation given
in Figure 7.5. The

208 Knit-line behaviour of PP and PP-blends Figure 7.3 Injection moulded test
platen with flow obstruction. Figure 7.4 Knit-line as a result of different cross-
sections in the mould cavity (isochrones). pressure depression near the end of the
part clearly indicates a radial flow, which may result not only in a knit-line but
also in entrapped air. In addition, knit-line weakness is enhanced by foreign
matter encased in between the two advancing melt fronts. Quite often this is
simply air or some other volatile material, but it could also be an accumulation
of mould releasing agent. Another phenomena would be jetting. These sources
can be accounted for as failures in mould design or process control, and can
therefore be avoided, which is not the case with the three major sources
discussed above. Furthermore, rapid cooling, as observed in all practical
processes, plays a crucial role in knit-line formation and subsequent weakness. It
is obvious that in many cases knit-lines cannot be avoided. However, if the
position of a knit-line can be predicted, it is possible to place it

209 Principles of knit-line formation in a region of the moulded item where it


does minimum harm. Whereas this can be achieved easily in simple cases using
existing experience, for more complicated parts a mould filling study has to be
made. Such a study can be made by computer simulation as shown in Figures 7.2
and 7.5. By varying the number and location of the gates the position of the
resulting knit-line(s) can be influenced. This is a very elegant way to deal with
the problem of knit-line position which saves both time and cost, because it can
be done prior to mould making. However, once a mould exists knit-lines can
also be found experimentally by a series of short mouldings, as shown in Figure
7.6 for the bumper shown in Figure 7.1. For a succession of short mouldings
with increasing injection pressure and subsequently increasing volume of
injected material, the position of knit-lines in the final part can be traced easily.
Figure 7.5 Mould filling simulation (isobars) for the cavity in Figure 7.4
(courtesy of General Electric Plastics Europe, Russelsheim). 7.3 PRINCIPLES
OF KNIT-LINE FORMATION A typical flow situation which leads to a knit-
line is shown in a schematic way in Figure 7.7, where two advancing melt
line is shown in a schematic way in Figure 7.7, where two advancing melt
streams are meeting each other head-on. Generally, i.e. for all thermoplastics, the
resulting knit-line

210 Knit-line behaviour of PP and PP-blends weakness is considered to have


three sources [3-6]: • insufficient interdiffusion of macromolecules of the
separated melt streams across the interfacial layer; • flow of melt orthogonal to
the main direction of flow and subsequently resulting in orthogonal orientation
(Figure 7.7, middle part); • creation of a V-shaped notch at the surface of the
plastic item (Figure 7.7, lower part). The macromolecules in a melt front are in
free motion without major acting external forces (most probable
thermodynamical situation) and therefore show entanglement, both with
themselves and with neighbouring macromolecules. When touching each other,
there is no mechanism by which enough energy could be brought into the
interface (i.e. the knit-line) to force entanglement between neighbouring
macromolecules of the two melt streams A and B in Figure 7.7. This is
particularly true in the case of a stagnating flow situation, as shown in Figure
7.7. The only mechanism Figure 7.6 Mouldings for stepwise filling of the cavity
for the car bumper in Figure 7.1.

Principles of knit-line formation 211 Fountain Flow //////////Λ^/////////_///////// ■—


77777777777777777777/'777777777 V-Notch ////////////////////////////// Figure 7.7
Schematic view of formation of a knit-line. creating entanglement is
interdiffusion, or rather self-diffusion, which is time dependent and would
therefore be subsequently hindered in practical cases by the usually rapid
cooling, i.e. rapid transition of the melt into a solid body. Depending on the
amount of resulting self-diffusion the knit-line will show a behaviour in between
the two extreme cases of pure adhesion (no entanglement) and pure cohesion
(sufficient penetration depth of macro¬ molecules), as shown schematically in
Figure 7.8. Plastic parts usually exhibit a certain anisotropy of mechanical
properties which is based on the orientation of macromolecules created by shear
flow during shaping. Particularly in the case of thin walled items, an increase of
mechanical strength in the direction of flow is observed. It is quite obvious that
for an orthogonal flow the resulting orientation will create a weak spot in terms
of mechanical properties. As before, the rapid cooling will exacer¬ bate the
problem because the unwanted orientations are quickly frozen in. In the case of
fibre filled materials (or any other filler with an aspect ratio) the problem will be
enlarged by the additional fibre orientation. Contrary to the orientation of the
macromolecules alone, the fibre orientation cannot be influenced by the cooling
rate or subsequent annealing. The V-notch is observed only in injection
rate or subsequent annealing. The V-notch is observed only in injection
moulding. It is a result of the combined effect of rapid cooling and the inability
of entrapped air between the melt fronts to be blown out quickly enough.
Particularly near the metal

212 Knit-line behaviour of PP and PP-blends I ////////////A//////////ZZJ'. Figure 7.8


Extreme cases of self-diffusion of macromolecules of two melt streams: D, depth
of penetration; rk diameter of entangled molecules. surfaces of the mould, the
viscosity increases rapidly due to the rapid temperature decrease of the melt.
After the transition from melt to solid body a further welding of the two melt
streams is no longer possible. The result is a circumferential optical mark which
often contributes to the mechanical weakness. 7.4 MORPHOLOGY OF KNIT-
LINES Almost all the experimental investigations on knit-lines were done with
test specimens that were injection moulded. Here, two basically simple but
different methods are known. In the more widely used of these the runner system
of an injection mould is arranged in such a way that a cavity for a test specimen
is filled simultaneously from both the far ends, thus creating a knit-line in the
middle portion. Quite often the mould contains a second cavity of the same
dimensions which is correspondingly filled from one end only in the same shot.
Figure 7.9 shows examples for an injection mould for dumb-bell shaped test
specimens and test specimens for impact and bending tests [1]. Using this
method, a test specimen with knit-line and one without knit-line are injection
moulded simultaneously under exactly the same processing conditions. Owing to
the flow observed in the cavity of a cooled

Morphology of knit-lines 213 Figure 7.9 Injection moulds for creating test
specimen with and without knit-line. injection mould the flow pattern of the
advancing fronts is of fountain type (occasionally also called ‘volcano’ or ‘cold
weld’) and leads to a stagnant flow situation after unification. A lesser number of
investigations were made with injection moulded plaques (see Figure 7.3). The
flow obstruction creates a knit-line which is of different type (sometimes also
called ‘warm weld’). The test specimens are then taken out orthogonal to the
main direction of flow and to the knit-line by milling. In addition to the fact that
this kind of knit-line formation allows for longer fusion time and therefore
usually creates lesser weakness com¬ pared to the fountain type knit-line, the
main orientation due to flow is parallel to the knit-line itself. Naturally there is a
strong influence of the injection moulding process on the morphology of knit-
lines, which is true not only for microscopic but also for macroscopic
appearance. A typical example for the macroscopic appear¬ ance is shown in
Figure 7.10, where the shape of the resulting knit-line across the wall thickness
of a double gated injection moulded test specimen can be seen [7]. Due to the
of a double gated injection moulded test specimen can be seen [7]. Due to the
differences in thermorheological history of the two melt streams prior to
unification as well as the differences in packing pressure effects, the resulting
knit-line may be orthogonal to the surface but may also be distorted into a
parabolic or even ox-bow shape, both of which can be explained by the
combined effect of packing pressure and cooling rate. Since it may be assumed
that differences in the shape of knit-lines may

214 Knit-line behaviour of PP and PP-blends Figure 7.10 Influence of melt


temperature on shape of knit-line across tl ness of a tensile test specimen of
PP/EPDM: (A) schematic view; (B) d micrograph (243 °C, reflected light), d is
the thickness. influence at least the mechanical properties, caution should be e
when comparing results from the literature: such an observation is r in very few
papers, an exception being [8]. One of the earliest results on the morphology of
knit-lines in reported by Hobbs [9]. Test specimens were moulded from an i

215 Morphology of knit-lines polypropylene following the technique of Figure


7.3, from which micro- tomed sections were taken across the knit-line at
different distances behind the flow obstruction. Clear evidence of V-notches
approximately 20 μηι in depth was found for low mould temperatures. With
increasing mould temperature the V-notch decreased in depth, and at mould
temperatures of 180 °C and above no visible mark could be seen on the surface
of the plaque. However, when viewed by transmitted light, areas of light
scattering could be observed running along the weld. Regarding the crystalline
morphology, it was found that the primary morphological feature was one or
more cylindritic bands nucleated on highly ordered material running along and
besides the actual knit area. In fact, the morphology at the weld is a true image
of the local processing conditions. When applying stress the failure, initiated by
the V-notches, travelled through the regions of orientation rather then through
the centre weld areas itself. The spatial distribution of the a- and β-phases of a
commercial grade PP homopolymer without nucleating agent was studied in
[10]. From injection moulded tensile test specimens it was found that apart from
the well known influence of processing conditions on molecular orientation and
thus on the crystallization process the most important factor is the injection
mode, which determines the presence or absence of ‘welding’ zones in the
sample. These zones are easily recognizable by radiocrystallography, where the
jS-phase in particular gives an interesting insight into the phenomenon. The
morphology of tensile test specimens with a knit-line injection moulded from a
commercial polypropylene was studied by X-ray wide- and small-angle
scattering, light scattering and polarization microscopy in [11, 12]. While the
scattering, light scattering and polarization microscopy in [11, 12]. While the
lamellar morphology is of negligible influence, it was found that the mechanical
properties of the samples are strongly influenced by the spherulitic structure.
Clustering of jS-phase spherulites and strong mor¬ phological inhomogeneities
in the vicinity of the knit-line are the primary reasons for mechanical weakness.
As expected, with increasing mould temperature the degree of crystallinity
increases in general, and the spherulite size in particular. A correlation between
the degree of crystallinity and the yield stress is given in Figure 7.11. Knit-lines
in PP-blends are as visible as those in the homopolymer. This is demonstrated in
Figure 7.12 for a commercial PP/EPDM blend injection moulded at 200 °C. The
morphology of weld lines in a blend of PP/EPDM was investigated in [13, 14]. It
was found that the elastomeric phase in injection moulded samples of these
blends was dispersed at the knit-line region in the form of thin sheets or flattened
domains which were oriented perpendicular to the (main) flow direction, i.e.
parallel to the knit-line. It was concluded that they prevent adequate matrix
union and probably act as stress concentrators which can provide additional loss
of knit-line strength. There appeared to be no particles at all at the centre of the
knit-line area, thus there is no adequate bonding across the knit-line region. It
was

216 Knit-line behaviour of PP and PP-blends Figure 7.11 Influence of degree of


crystallinity in a knit-line on yield stress [11]. Figure 7.12 Knit-line in a
PP/EPDM blend in polarized light (courtesy of Hoechst AG, Frankfurt).

Morphology of knit-lines 217 suggested that less easily dispersed higher


viscosity EPDM compositions would improve the knit-line strength.
Deformation of the second phase is only possible for non-rigid (not cross-linked)
particles and a flow situation with high shear stresses. In Figure 7.13 the
deformation of EPDM particles in the knit-line region of PP/EPDM is shown
[1]. The same kind of deformation of the rubber phase was found for PP-blends
modified with EPDM and thermoplastic poly- olefinic (TPO) rubber [15].
However the results did not match exactly the findings of [13, 14]. Crazing is
initiated both at the edge of the frozen flow front in the case of the cold welded
specimens and in the shear zone parallel to the mould flow direction [8, 9].
Incompatible polymer blends will exhibit very poor knit-line strength. However,
better knit-line performance is expected for the blends modified with high
viscosity second phase and high melt elasticity [13, 14]. Figure 7.13
Deformation of EPDM particles in the knit-line region for a PP/EPDM blend.
The layered matrix structure and the deformation of the particles in blends show
analogies with the injection moulded composites reinforced with rigid fibres.
Indeed, orientation of the fibrous reinforcement parallel to the knit-line causes a
Indeed, orientation of the fibrous reinforcement parallel to the knit-line causes a
significant reduction of the tensile strength compared to the knit-free product
[16]. The strength ratio of specimens with and without knits was found to
decrease with increasing fibre concentration. In general, these results are
confirmed by the investigation presented in [17, 18], where again for injection
moulded test specimens from glass fibre- reinforced polypropylene a decreasing
knit-line strength was found for increasing fibre volume fraction. In Figure 7.14
the fibre orientation in the knit-line is shown from two different sources [16, 17].
An even worse

218 Knit-line behaviour of PP and PP-blends Figure 7.14 Scanning electron


micrographs of tensile fracture surfaces of welded (a, c) and unwelded (b, d)
injection moulded sample of glass fibre-reinforced PP. (a, b) 20 wt% [16]; (c, d)
35 wt% [18]. situation is given for long-fibre-reinforced polypropylene, which
shows poorer knit-line strength compared to short-fibre compounds because of
voids in and near the knit-line [19]. In general, these observations hold true for
any other filler with an aspect ratio. For polypropylene reinforced with mica
flakes it was found that the knit-line weakness is again due to orientation and
that the knit-line strength decreases significantly with increased volume fraction.
The smaller the flakes, the stronger the knit-line [20]. 7.5 MECHANICAL
PROPERTIES OF KNIT-LINES Before discussing the properties of knit-line
infested PP it should be kept in mind that the properties of any plastics part are
strongly dependent on the moulding process itself. This is particularly true for
injection moulded test specimens where, independent of the presence of a knit-
line, a very complex pattern of orientations and internal stresses is created which
influences the mechanical testing results. It is therefore very difficult to
distinguish between the process-induced and the knit-line-induced effects. Even
when the pro¬ cessing conditions are mentioned, comparison of results from
different

Mechanical properties 219 papers is almost impossible. Moreover, the grades of


PP investigated differ widely, and sometimes the results are presented as the so-
called knit-line factor (KLF), which is of relative nature: KLF ProPerty ^nit-line
infested test specimen property of test specimen without knit-line One of the
earliest papers can be taken as a very good example for the influence of knit-
lines on the mechanical properties of PP [21]. As described earlier, four test
specimen were injection moulded in one shot for tensile and impact testing with
and without knit-lines. For an unspecified isotactic PP it was found that KLF for
the tensile strength was around 1 (slightly decreasing with increasing melt
temperature), for elongation at break between 0.4 and 1 and for impact strength
temperature), for elongation at break between 0.4 and 1 and for impact strength
between 0.33 and 0.47, as shown in Figure 7.15. The result is interesting in the
sense that it clearly shows a higher sensitivity of elongation and impact to the
presence of a knit-line compared to the tensile strength, and also contradicts the
general belief that the knit-lines would always improve with increasing
temperature. It is somewhat surprising to note that there is such a pronounced
influence on elongation, which itself is related to the length of the test specimen,
of which the knit-line infested region is only a very small portion. As far as
strength obtained from tensile tests is concerned, it is confirmed that in some
cases PP does not indeed show any influence of the knit¬ line (e.g. [22]).
However, there are many results that refer to a certain knit-line weakness. A
collection of data from various sources for PP is f--1 j | Scatter ■ 1 ■ 1 t-l i Γ"! I
220 °C 1 1 250°C ^ 280°C i — j 1 1 1 l r "l ■ 1 LU1 ύ a b c Figure 7.15 Knit-
line factors for (a) tensile strength, (b) elongation at break and (c) impact
strength of isotactic PP for different melt temperatures [21].

220 Knit-line behaviour of PP and PP-blends given in Table 7.1 [8, 16, 18, 22-
27]. It shows a wide scatter of results. Differences are even observed when the
same material is tested by two independent institutions. For example, from the
exhaustive work by Buck- nail [27, 28], it is reported that a test specimen from a
PP copolymer processed by injection moulding by a major raw material
producer showed a KLF>1 in falling weight tests when tested in the laboratories
of the supplier, compared to a value of only 0.2 when tested elsewhere. Many
results have also been reported in the literature for a range of PP blends,
including with EPDM [7], with high density polyethylene (HDPE) [22], with
polycarbonate [23], and with low density polyethylene (LDPE) [29]. In view of
the vast differences in material properties and blending techniques it does not
seem feasible to give a comparison of results like that given for pure PP in Table
7.1. However, it should be noted that the presence of a second phase in general
increases knit-line weakness. A good example is given in [23], where it was
found that with increasing content of polycarbonate the strength of the PP/PC-
blends may increase by as much as 30%. The presence of a knit-line, however,
brings the strength values sharply down to 25% in case of stress at yield or even
less than 10% in case of elongation at yield. Similarly to the effect of knit-lines
for blends, the addition of fillers will increase the problem of knit-line strength.
A drastic reduction of mechanical properties may be observed, particularly in the
case of fibres. For example, whereas the polypropylene investigated in [16]
shows very little effect of the presence of a knit-line (Table 7.1), when the same
material is reinforced by 20% glass fibres the knit-line factor is only around 0.5.
The influence of different fillers is demonstrated in Table 7.2 for poly¬
The influence of different fillers is demonstrated in Table 7.2 for poly¬
propylene filled with talc, glass fibres (GF) and calcium carbonate (CC) [18]. As
expected, it can be clearly seen that the higher the aspect ratio, the more
pronounced the effect on knit-line strength. As already mentioned, filler content
also has a pronounced influence. It is reported [17] that with an increase of
volume fraction of filler content from 0 to 20% the tensile strength of knit-line
infested test specimens decreases by about 35%. In Figure 7.16 the influence of
filler content on the knit-line factor for tensile strength is given for different
grades of fillers [20]. Annealing may help to improve the mechanical properties
of knit-line infested regions, not only in pure PP or PP blends but also in filled
PP [20]. This is somewhat surprising because, unlike the macromolecules,
annealing cannot alter the position of filler particles, and cannot therefore alter
the overall orientation of fillers. It also should be reported that the addition of
fillers will not always have a drastic effect on knit-line strength compared to
unfilled material. In [26] the impact behaviour of injection moulded short-fibre
composites was investigated by impact testing circular discs. Whereas in the
case of pure PP the knit-line factor for maximum force is

Table 7.1 Compilation of results of knit-line strength from various literature


sources Knit-line factor Material Property a b Remarks Reference PP copolymer
Fatigue bending test 0.95 Constant load, [8] four-point bending PP Tensile
strength 0.95 0.94 a 220°C [16] Tensile modulus 1 1.18 b 250°C Elongation 0.05
0.06 PP Tensile strength 1 [18] PP Tensile strength 1 [22] PP homopolymer
Stress at yield 0.95 0.82 a: test specimen machined from injection moulded
plaque b: test specimen directly injection [23] Elongation at yield 0.99 0.97
moulded Tensile modulus 0.99 1 PP homopolymer Impact strength 0.55 1 a:
orthogonal knit-line [24] PP block copolymer 0.19 0.54 PP block copolymer
0.24 0.89 b: parallel knit-line PP homopolymer Elongation at break 1.29 PP
block copolymer 0.20 PP block copolymer 0.14 PP Tensile strength 0.92 [25] PP
Impact strength: [26] Total energy 0.48 Maximum force 0.63 PP homopolymer
Dart impact energy 0.55 Test performed with discs of [27] PP copolymer 0.18
special shape

222 Knit-line behaviour of PP and PP-blends Table 7.2 KLFs for various filled
PPs KLF for tensile strength Material Knit-line behind flow obstruction Knit-
line in double gated specimen PP 1 0.90 PP + 20% talc 0.82 0.67 PP + 30% GF
0.69 0.44 PP + 20% CC 0.97 0.91 Figure 7.16 Influence of volume and grade on
tensile strength of mica filled PP [20]. 1-4, different grades of filler. only 0.63,
the decrease of the same value for knit-line infested PP with 30% by weight of
glass fibres would be 15% only, i.e. KLF = 0.85. However, the KLFs, for total
energy are around 50% for both the materials. As expected from the results
energy are around 50% for both the materials. As expected from the results
presented so far, a certain non-conformity is also observed with the influence of
processing conditions on knit-line strength. Though a number of papers report an
increase of knit-line strength with increasing melt temperature and/or increasing
mould temperature [3, 30, 31], there are also observations like those shown in
Figure 7.17 [7]. It is interesting to note that for test specimens both with and
without knit-line, the absolute values of tensile strength decreased with
increasing melt temperature. However, since the decrease of the knit-line
strength is com¬ paratively smaller, the KLF would show an increase, which can
be mislead¬ ing when assessing of the situation. From the same investigation, it
isi reported that neither mould temperature nor packing pressure shows a
pronounced effect on knit-line strength from tensile tests and impact tests.
However, again, it was found that elongation is more sensitive to the. presence
of a knit-line as tensile strength, as shown in Figure 7.18 for the full set of results
obtained over a range of processing temperatures from 213 to

Concluding remarks 223 Melt Temperature (°C) Figure 7.17 Influence of melt
temperature on injection moulded PP/EPDM test specimens with and without
knit-lines. Figure 7.18 Knit-line factors for mechanical properties of a PP/EPDM
blend over a range of processing parameters: a, yield stress; b, elongation at
yield; c, tensile modulus; d, elongation at break; e, impact strength. 258 °C,
mould temperatures from 25 to 55 °C and packing pressures from 35 to 55 MPa
[1]. 7.6 CONCLUDING REMARKS In many practical cases knit-lines cannot
be avoided and can at best be placed in a region with minimum load. It is
therefore of importance to know

224 Knit-line behaviour of PP and PP-blends the decrease of mechanical


properties created by a knit-line beyond the purely qualitative observation that
given sufficient distance to a flow obstacle the knit-line region may recover to
ultimately regain the strength of the material itself. Although there has been a
great deal of practical experience in addition to the results from a number of
investigations on thermoplastics in general and on PP in particular, the overall
findings show that it is not possible to predict the influence of a knit-line. For
quantitative assessment experimental investigations seem inevitable.
Unfortunately this is also true for aspects which seem to be confirmed by
practical experience, such as improving the knit-line weakness by increasing the
melt temperature or the mould temperature. Moreover, almost all the results
reported on mechanical properties are from laboratory tests only, yet knit-lines
are certainly not an academic phenomenon, as shown in Figure 7.19, where the
knit-line strength is given for samples taken from the bumper Cross Section of
knit-line strength is given for samples taken from the bumper Cross Section of
Car Bumper Figure 7.19 Tensile strength of test specimen (a) without and (b)
with knit-lines taken from an injection moulded car bumper. 1, test specimen. 2,
sprue. shown in Figure 7.1. Most of the results were reported from tensile tests,
which seem to be less sensitive to the presence of a knit-line (with the exception
of elongation) than dynamic tests (impact or fatigue, the latter being reported
only once [8]). For practical applications dynamic testing is far more important,
as is creep, for which no data are available. As expected, the situation is even
more complex when dealing with PP blends or filled PP. Although the basic
mechanism of knit-line formation is known, the reasons for the resulting knit-
line weakness are not yet fully understood.

References 225 This is mainly because in all practical cases the given three
sources (insufficient interdiffusion, orientation, V-notch) are interrelated in such
a way that it is usually impossible to correlate experimental findings with one of
the sources only. Furthermore, the complex flow situation is overshadow¬ ing in
a very severe way, so that it is impossible to make statements with regard to the
speed of entanglement or the degree of orientation (or probably the effect of
orientation on speed of entanglement). Overall, the results leave no doubt that
the knit-line strength is not a material property but the property of a system
depending as much on processing conditions and, subsequently, on the local
flow and cooling situation as on the physical properties of the material itself. 7.7
ACKNOWLEDGEMENTS Financial support from the Deutsche
Forschungsgemeinschaft (DFG) and the Bundesministerium fur Wirtschaft
through the Arbeitsgemeinschaft Industrieller Forschungsvereinigungen (AIF) is
gratefully acknowledged. 7.8 APPENDIX: Abbreviations CC Calcium carbonate
EPDM Ethylene-propylene diene monomer GF Glass fibre HDPE High density
polyethylene KLF Knit-line factor LDPE Low density polyethylene PC
Polycarbonate PP Polypropylene TPO Thermoplastic polyolefinic 7.9
REFERENCES 1. Mennig, G. (1988) Mat.-Wiss. u. Werkstofftech, 19 (11), 383-
90. 2. Morwald, K. and Reitmann, P. (1984) Plastverarbeiter, 35 (11), 68-72. 3.
Malguarnera, S. C. (1982) Polymer-Plastics Technology and Engineering, 18
(1), 1-45. 4. Mosle, H. G., Criens, R. M. and Dirk. H. (1984) SPE ANTEC
Technical Papers, 30, 772. 5. Bataille, F., Vu-Khanh, T. and Fisa, B. (1985) SPE
ANTEC Technical Papers, 31, 1174. 6. Kim, S.-G. and Suh, N. P. (1984) SPE
ANTEC Technical Papers, 30, 111. 7. Mennig, G. (1992) Kunststoffe, 82, 235-8.
8. Watkinson, K., Thomas, A. and Bevis, M. (1982) Journal of Materials
Science, 17, 347-58.

226 Knit-line behaviour of PP and PP-blends 9. Hobbs, S. Y. (1974) Polymer


226 Knit-line behaviour of PP and PP-blends 9. Hobbs, S. Y. (1974) Polymer
Engineering and Science, 14, 621-6. 10. Trotignon, J. P., Lebrun, J. L. and
Verdu, J. (1982) Plastics and Rubbe) Processing and Applications, 2, 247-51. 11.
Singh, D. and Mosle, H.-G. (1988) Makromolekulare Chemie, Macromoleculai
Symposia, 20/21, 489-500. 12. Wenig, W., Singh, D., Botzen, G. and Mosle, H.-
G. (1990) Angewandte Mak romolekulare Chemie, 179, 35-56. 13. Thamm, R.
C. (1977) Rubber Chemistry and Technology, 50, 24-34. 14. Thamm, R. C.
(1989) Polymer Engineering and Science, 29, 209-13. 15. Karger-Kocsis, J. and
Csikai, I. (1987) Polymer Engineering and Science, 27 241-53. 16. Vaxman, A.,
Narkis, M., Siegmann, A. and Kenig, S. (1991) Polymer Composites 12, 161-8.
17. Fisa, B. and Rahmani, M. (1991) Polymer Engineering and Science, 31,
1330-6. 18. Sanschagrin, B., Gauvin, R., Fisa, B. and Vu-Khanh, T. (1990)
Journal o> Reinforced Plastic Composites, 8, 194-208. 19. Bouti, A. and Fisa, B.
(1991) SPE ANT EC, 49, 2112-16. 20. Svazic, M. (1986) SPE ANTEC
Technical Papers, 32, 1136-40. 21. Ehms, E. and Bussian, M. (1972) Plaste und
Kautschuk, 19, 214-17. 22. Bartlett, D. W., Barlow, J. W. and Paul, D. R. (1982)
Journal of Applied Polymei Science, 27, 2351-60. 23. Bourgeois, S., Fisa, B. and
Favis, B. D. (1989) SPE ANTEC, 47 323-6. 24. Kloos, F., Strouk, H., Uebe, R.
and Heufer, G. (1991) Angewandte Mak¬ romolekulare Chemie, 185/186, 97-
108. 25. Menges, G., Schacht, Th. and Ott, St. (1988) Kunststoffberater, 33 (4),
54-7. 26. Hogg, P. J. (1987) Composites Science and Technology, 29 (2), 89-
102. 27. Bucknall, C. B. (1986) Pure and Applied Chemistry, 58 (7), 999-1014.
28. Bucknall, C. B. (1986) Pure and Applied Chemistry, 58 (7), 985-98. 29.
Nolley, E., Barlow, J. W. and Paul, D. R. (1980) Polymer Engineering and
Science 20, 364-9. 30. Malguarnera, S. C. and Manisali, A. (1981) Polymer
Engineering and Science, 21 586-93. 31. Malguarnera, S. C., Manisali, A. I. and
Riggs, D. C. (1981) Polymer Engineering and Science, 21, 1149-55.

8 Welding and fracture of polypropylene interfaces R. P. Wool 8.1


INTRODUCTION The joining or welding of polymers can be accomplished by
several techniques, including thermal welding, vibrational welding, friction
welding, solvent welding, surface chemical modification, ion beam surface
modifica¬ tion, resonance heating and other more exotic but less common
techniques. In this chapter we focus on polymer welding, in which the molecules
near the surfaces become mobile and the weld strength develops by a combina¬
tion of surface approach, surface rearrangement, wetting and diffusion. The
molecular mobility can be induced by heat from friction, (ultrasonic) vibration,
impact, hot plates, solvents, microwave radiation, etc. We consider here the
problem of strength development at polymer- polymer interfaces in terms of the
static and dynamic properties of random- coil chains [1-4]. When two pieces of
molten polymer are brought into contact, wetting or close molecular contact (van
molten polymer are brought into contact, wetting or close molecular contact (van
der Waals) first occurs followed by interdiffusion of chain segments back and
forth across the wetted interface. After a contact time, t, we determine the
mechanical energy, G, required to separate the two pieces as a function of time,
temperature T contact pressure, P, and molecular weight M of the linear random-
coil chains, i.e.: G=W(t9T9P9M) (1) where W is the welding function to be
determined. Solutions to this problem can be applied to many similar situations
involving polymer-polymer interfaces, such as processing of powder and pellet
resin, internal weld lines Polypropylene: Structure, blends and composites.
Edited by J. Karger-Kocsis. Published in 1995 by Chapman & Hall, London.
ISBN 0 412 58430 1

228 Welding and fracture of polypropylene interfaces of polymer melts during


extrusion and injection molding, welding of surfaces, lamination of composites,
coextrusion, and autohesion of uncured linear elastomers. 8.2 POLYMER
INTERFACES 8.2.1 Polymer welding At a polymer-polymer interface, the
chains have been physically separated at i = 0 such that diffusion to a distance x
equal to the radius of gyration Rg, will be necessary to re-establish the
interpenetrated structure of the virgin state of the melt shown in Figure 8.1. The
time to achieve this level of interpenetration can be designated as the weld time,
iw. The dynamic properties of polymer chains in dilute and concentrated states
have been analyzed in detail by Rouse [5], de Gennes [6, 7], de Gennes and
Leger [8], Doi and Edwards [9] and many others. We could estimate iw from the
diffusion behavior in the bulk or in solution. For example, one-dimensional
Fickian diffusion of polymer chains gives the center of mass motion >2, as: (X}2
= 2Dt (2) (a,b) (c) Η H Figure 8.1 The interdiffusion of random coil chains at a
polymer-polymer interface (—) is shown schematically in terms of (a, b) surface
rearrangement and approach, (c) wetting, (d) diffusion and (e) virgin state.

Polymer interfaces 229 where D is the self-diffusion coefficient. If we assume


for polymer welding that <X)2^Kg at i = iw, then it follows that: t^R2JD (3)
From the classical work of Rouse [5], we find that for both dilute chains (non-
entangled) and for melt chains with molecular weights less than the critical
entanglement molecular weight, Mc, we have the molecular weight dependence
of the self-diffusion coefficient DRO, as: Pro -1 (4) Chains with Rouse dynamics
would therefore have a weld time given by iw(Rouse)^M2 (5) For highly
entangled polymer melts with M>MC, the diffusion coefficient is [6-9]: Dr*M~2
(6) such that the weld time could be estimated as tw(reptation)« M3 (7) Thus,
small differences in molecular weight could result in large differences in the
time to achieve optimal weld conditions. 8.2.2 Strength development at polymer
time to achieve optimal weld conditions. 8.2.2 Strength development at polymer
interfaces The problem of evaluating the strength of polymer interfaces is sum¬
marized in Figure 8.2. Material A is brought into contact with material B to form
an A/B interface. The weld is fractured and the strength is related to the structure
of the interface through microscopic deformation mechan¬ isms. Typically, a
crack propagates through the interface region preceded by a deformation zone at
the crack tip. The development of the relation between interface structure and
strength proceeds as described below. Step 1 The time-dependent structure of the
interface is determined. Relevant properties may be characterized by a function
H(t), which for the case of polymer melts can usually be described in terms of
the static and dynamic properties of the polymer chains. For example, with
symmetric (A = B) amorphous melt interfaces, H{t) describes the average
molecular properties developed at the interface by the interdiffusion of random
coil chains, as [4]: H(t) = Hx(t/T)^ (8)

230 Welding and fracture of polypropylene interfaces p d i T a Interface -—2h


Figure 8.2 Fracture (wedge cleavage) of an A/B interface occurs by propagation
of a crack through a deformation zone at the crack tip. The fracture energy is
related to the structure of the interface in terms of the deformation mechanisms
occurring in the deformation zone. in which τ is a characteristic relaxation time
and His the equilibrium value of the property H(t) at t = z. H(t) can be measured
from the concentration depth profile C(x, t) for symmetric interfaces. This
function could be a measure of the extent of diffusion across the interface, e.g.
the average monomer interpenetration distance, the number of chains crossing
the interface plane, etc. With incompatible amorphous interfaces, limited inter¬
diffusion occurs up to an equilibrium depth d<*>, which becomes the important
descriptor of the interface structure. The interface properties can usually be
independently measured by a number of spectroscopic and surface analysis
techniques, such as secondary ion mass spectroscopy (SIMS), X-ray
photoelectron spectroscopy (XPS), specular neutron reflection (SNR), forward
recoil spectroscopy (FRES), scanning electron microscopy (SEM), transmission
electron microscopy (TEM), infrared (IR) and several other methods.
Theoretical and computer simulation methods can also be used to evaluate H(t).
Thus, we assume for each interface that we have the ability to measure H(t) at
different welding times and that the function is well defined. where
f/00«M(3r_s)/4 r, s= 1, 2, 3 (9)

Polymer interfaces 231 Step 2 After a contact time £, the material is fractured or
fatigued and the mechanical properties G(£), determined. The measured
properties will be a function of the test configuration, rate of testing,
temperature, etc. These include the critical strain energy release rate G1C, the
temperature, etc. These include the critical strain energy release rate G1C, the
critical stress intensity factor KlCi the critical crack opening displacement <5C,
the critical fracture stress σ0 the fatigue crack propagation rate da/dN, and other
properties. Related properties can be measured when other modes of fracture
(torsion and shear) are used. Step 3 The fracture properties G(t) are related to the
interface structure H(t) through suitable deformation mechanisms deduced from
the microm¬ echanics of fracture. This is the most difficult part of the problem
but the analysis of the fracture process in situ can lead to valuable information
on the microscopic deformation mechanisms. SEM, optical and XPS analysis of
the fractured interface usually determine the mode of fracture (cohesive,
adhesive or mixed) and details of the fracture micromechanics. However,
considerable modelling may be required with entanglement and chain fracture
mechanisms to realize useful solutions. We then obtain a solution to the
problem: G(t)=f{H(t)} (10) where/ is a known function of H(t) at constant
temperature and pressure. In the simplest case, G(t) is proportional to H(t), as
found for many polymer welding problems [4]. 8.2.3 Structure and strength of
interfaces When the interrelation between the structure and strength of the
interface is determined, several unique predictions for the mechanical properties
can usually be made. In the case of polymer melts, the strength predictions are
made in terms of the static and dynamic scaling laws of the polymer chains.
Thus, we can investigate scaling laws for fracture energy G1C, fatigue crack
propagation rate da/diV, etc., as a function of time and molecular weight. For
example, if G1C &H(t)9 where H(t) is given by Equation 8, then we expect
solutions of the form: Gic(i)*tr/4M"S/4 (11) Glc(T)»M(3r"s)/4 (12) The fracture
energy is then predicted to have a precise time dependence, a molecular weight
dependence of the rate of welding and a molecular

232 Welding and fracture of polypropylene interfaces weight dependence of the


virgin or fully welded state. Experimental determi¬ nation of the constants r and
s should also indicate which of the molecular properties is important in
controlling the fracture energy. We have argued [4] for interface fracture
dominated by chain disentanglement that r = s = 2. The relation between G1C
and many other property H(t) is Gic/Goo =(H/Hao)2/r (13) This suggests that the
average interpenetration chain length </> (with r = 2) plays a major role in
controlling the strength of polymer melt interfaces. Also, we expect that Glc&n2
(with r=l), which is important for compatibilizers that can segregate to an
interface, and G1C&X2, which is important for incompatible polymer interfaces.
The relation between struc¬ ture and strength of interfaces is a very interesting
problem, and has also been discussed by Kausch [10], Prager and colleagues
[11-13] and de Gennes [14]. 8.2.4 Classification of polymer interfaces Polymer
[11-13] and de Gennes [14]. 8.2.4 Classification of polymer interfaces Polymer
interfaces can be categorized in four broad groups: 1. Symmetric polymer-
polymer interfaces 2. Asymmetric polymer-polymer interfaces 3. Polymer-non-
polymer interfaces 4. Multicomponent polymer interfaces The different classes
within each group are listed in Table 8.1. The symmetric interfaces are most
commonly encountered in high volume applications with melt processing of
plastics. Manufacturing methods in¬ volving extrusion, injection molding,
compression molding, powder and pellet sintering, welding and lamination all
involve amorphous polymer- polymer interfaces. When the melt is cooled, the
interface may remain in the amorphous state when polymer glasses (polystyrene
(PS), polymethylmethacrylate (PMMA), polyvinyl chloride (PVC),
polystyreneacrylonitrile (PSAN), polycarbonate (PC)) and rubbers
(polybutadiene (PB), polyimides (Pis)) are being made, or crystallize when
semicrystalline materials (polyethylene (PE), polypropylene (PP), polyethylene
terephthalate (PET), polyamide (PA)) are used. The crystallization process can
significantly change the structure that existed in the melt. Symmetric polymer-
polymer interfaces The symmetric amorphous interface has been studied by
Voyutskii, [15], Kausch [10], Prager and Tirrell [13], de Gennes [14], and Wool
and colleagues [1-4]. In these studies, the structure of the interface was deter-

Polymer interfaces Table 8.1 Classification of polymer interfaces 233 1.


Symmetric A/A polymer-polymer Amorphous Crystal Liquid crystal 2.
Asymmetric polymer-polymer Amorphous structural asymmetry Amorphous
dissimilar compatible Amorphous dissimilar incompatible Crystalline similar
Crystalline dissimilar compatible Crystalline dissimilar incompatible 3.
Asymmetric polymer-non polymer Polymer-metal Polymer-ceramic Polymer-
biological 4. Multicomponent polymer interfaces Blends Compatibilizers
Interactive Non-interactive mined from interdiffusion of amorphous chains and
related to the strength of the interface by several microscopic deformation
mechanisms. Reviews of ‘healing’ and welding in symmetric amorphous
interfaces are given by Kausch [10] and Wool [4]. When a polymer interface is
cooled to temperatures Tc, below the melting point Tm, crystallization
commences, as with PP, PE, PET, etc. Crystalliza¬ tion occurs via homogeneous
or heterogeneous nucleation, followed by growth of crystalline lamella, which
involves a secondary nucleation process. We may inquire how the crystallization
processes in the vicinity of the interface affect its structure and strength. Several
points are of interest, especially for partially diffused interfaces, i.e. interfaces
for which the average monomer interdiffusion distance X(t)<Xo0, where X~Rg.
Consider two melt pieces, the surfaces of which have been equilibrated in air
(without oxidation), and then brought into good contact for a time ί<τ, and
(without oxidation), and then brought into good contact for a time ί<τ, and
rapidly cooled to Tc<Tm. The chains near the surface have a reduced entropy
and a more highly ordered state with potentially partially oriented regions.
Therefore, at times immediately after contact, before extensive diffusion has
occurred, the surface layer of chains should have a higher probability for the
formation of critical sized embryos. In effect, the surface layer acts as pre-
embryos for the normal bulk homogeneous nucleation mechanism. Also, the
orientation of the embryos on the surface is such that their principal director or
c-axis orientation direction lies parallel to the

234 Welding and fracture of polypropylene interfaces surface. Crystal growth on


both sides of the interface with this orientation would result in an interface
region with poor mechanical properties. Similar problems apply to liquid
crystals. The oriented embryo problem is further aggravated by orientation due
to melt flow in a mold. In the above problem, the impinging melt fronts forming
the weld line in a double gated mold (i.e. with two injection points) have the
chains in a biaxially oriented state. At a symmetric polymer-polymer interface,
the spatial distribution of nucleating agents is an interesting problem and
depends on the history of the melt. Before the interface was formed in the melt,
the polymer may have been crystallized several times as a normal part of the
processing and pelletizing operation. During crystallization, many of the foreign
particles that are not participating in nucleation can be zone refined, or exuded to
the boundaries of the local crystallization front. These impurities may form in
pockets, which could lie in the interstices of the impinged spherulitic structure.
Their redistribution in the melt will depend on the melt flow and mixing history.
Compression molding and sintering of powder and pellet resin should involve
more predictable distributions than injection molding or coextrusion, for
example. Asymmetric polymer-polymer interfaces While the number of
symmetric A/A interfaces can be counted on one hand, the number and
complexity of asymmetric A/B interfaces is far greater, as shown in Table 8.1.
The welding of asymmetric incompatible and compat¬ ible amorphous
interfaces has been studied by Helfand and colleagues [16-18], Wu, Chuang and
Han [19], Wool and colleagues [20-24] and Paul [25]. Even though incompatible
polymers are immiscible, a region of mixing, d<*,, develops at the interface
which is predicted to vary as the inverse square root of the Flory-Huggins
interaction parameter, χ [26]. On the basis of studies of symmetric interfaces in
which the welding fracture energy is found to be proportional to the average
interdiffused contour length, l(t) [1, 2, 4], we predict that the fracture energy of
an incompatible interface behaves as Glc&l(t)&d2& l/χ. Experimental support
for this prediction was found with incompatible interface pairs consisting of
for this prediction was found with incompatible interface pairs consisting of
PS/PMMA, PSAN/PC, and PSAN/PMMA. Amorphous structural asymmetry
This interface is essentially chemically symmetric but involves an A/B couple
whose architectural structures are different, as shown in Figure 8.3. Many
practical examples of interfaces with structural asymmetry exist in current
technologies. Fabrication of composites with epoxies often occurs in a sequential
process such that the first laminate A, is more cured (crosslinked) than the
second laminate B. Hundreds of laminates formed by tape winding

Polymer interfaces 235 Linear Branched Cyclics M, M2 Crosslinked Figure 8.3


Molecular architectures. may be involved, for example in the manufacture of 30-
cm-thick submarine hulls made with fiber reinforced epoxy. Similar technology
is used for rocket motors. If sufficient diffusion does not occur at the A/B
interface before the reaction takes place, then the composite can be weak. Thus,
the diffusion rate must exceed the chemical reaction rate. One can therefore use
additives which will promote the interdiffusion but not alter the reaction rate
significantly. A similar process applies to electronic material fabrication with Pis
and metallized layers where a fresh PI layer is placed on an older more cured PI
layer. An important example of structural asymmetry in commodity plastics is
that of the amorphous interface of linear polymer chains with different molecular
weights, Mj/Mj, and molecular weight distributions. Such interfa¬ ces are
commonly encountered in coextrusion of different grades of the same polymer,
lamination of composites with varying reactivity ratios in the matrices, coatings,
recycling of plastics and more generally in high shear rate extrusion processes
with homopolymers where the residence time affects the molecular weight
distributions. More exotic examples exist when one considers A/B combinations
of linear, branched, cyclic, stars, combs, crosslinked and other structures. For
example, the melt processing of the PE family (high density PE (HDPE), low
density PE (LDPE), linear low density PE (LLDPE) and UHMPE) with

236 Welding and fracture of polypropylene interfaces different molecular


weights involves a matrix of four (diagonal) M1/M2 interfaces and six (off
diagonal) asymmetric interfaces, as shown in Table 8.2. In general, if we have n
different structures within a given family of the same chemical composition, we
will have n Mj/Mj asymmetric, η A/A symmetric interfaces and a number of
asymmetric A/B interfaces NA/B given by ΝΑ/β=Φ—l)/2 (14) Table 8.2
Interface A/B pairs for polyethylene A/B HDPE LDPE LLDPE UHMPE HOPE
LDPE LLDPE UHMPE HDPE/HDPE LDPE/HDPE LDPE/LDPE
LLDPE/HDPE LLDPE/LDPE LLDPE/LLDPE UHMPE/HDPE UHMPE/LDPE
UHMPE/LLDPE UHMPE/UHMPE One might conjecture that interfaces formed
from the same chemical species but different structure represent a simple
from the same chemical species but different structure represent a simple
extension of the symmetric A/A interface. However, this is not the case and the
structure of these interfaces in terms of the static and dynamic properties of the
chains can be exceedingly complex. For example, with a high and low molecular
weight couple of linear chains, the flux of molecules across the interface plane is
uneven due to the more rapid motion of the low molecular weight species. As a
result, the interface plane moves (in the direction of the low molecular weight
side), the high molecular weight side swells with the low molecular weight
species and the motion of the chains alters with time and position during
diffusion. Incompatible amorphous interfaces The second most important
interface, besides the symmetric amorphous interface, is the incompatible
amorphous interface. Most polymer pairs are not miscible to the extent that they
can form homogeneous melts and solutions. Immiscible polymers can be forced
to form pseudo-homogeneous mixtures by rapid pressure-drop quenching from
supercritical solvents. However, annealing induces phase separation.
Incompatible polymer-poly¬ mer interfaces are commonly encountered in
semicrystalline polymer blends (PP particles dispersed in a PE matrix),
toughened glass (rubber particles in a PMMA matrix), recycling of plastics (co-
mingled plastics from municipal solid waste streams), coextrusion (PSAN in
PC), and many other ap¬ plications, which are discussed in [4]. The major
practical difficulty with incompatible interfaces is that they are very weak
compared to the virgin strength of either component. For example, the
PMMA/PS interface has a strength of about 45J/m2 compared to the virgin
strength of about 1000 J/m2 [23, 24]. The strength can be considerably improved
by using

Polymer interfaces 237 compatibilizers. These are typically diblock or triblock


copolymers which have miscible groups capable of diffusing into one side and
anchoring the other by straddling the interface. The cornerstone for the treatment
of miscibility in polymer blends and interfaces is the Flory-Huggins
thermodynamic theory [26]. For a binary A/B blend, the Helmholtz free energy
per monomer ΔΑ, of mixing chains with degree of polymerization NA and NB
and monomer volume fractions φ and (1 — φ) respectively, is determined by:
AA//cT= {χφ(1 — φ)}enthalpy + {Φ In φ/Ν A + (1 - φ) ln(l - </>)/iVB}entropy
(15) where T is the temperature, k is Boltzmann’s constant and χ is the Flory-
Huggins interaction parameter. The entropy term on the right- hand side, known
as the combinatorial entropy of mixing, is always negative but becomes smaller
and approaches zero with increasing NA and Nb. This theory was developed for
a lattice with a coordination number z. While the entropic factor favors mixing
as in ideal solutions, the monomer interactions expressed through nearest
as in ideal solutions, the monomer interactions expressed through nearest
neighbor attractive ener¬ gies £, may oppose mixing and this effect is quantified
through the χ parameter: where eAA, £BB and £AB are the interaction energies
between A A, BB and AB monomers, respectively. Thus, the χ parameter only
examines specific monomer interactions and examines the balance of energy
required to break the homopolymer bonds and form two new AB bonds. Three
situations are important: 1. When £AA+£BB =2£AB, χ = 0 and we have an
athermal blend which is similar to the symmetric A/A interface. Strength
through interdiffusion can occur. 2. When £AA +£BB <2£AB, the formation of
AB bonds is favored, χ is negative and the blend is compatible. Strength will
develop at a faster rate than in the athermal blend since the molecules will
experience an additional driving force to diffuse across the interface. 3. When
£AA +£bb >2£ab, χ is positive and mixing is not favored. Under these
conditions, A and B components will not mix, and mechanical mixtures will
attempt to phase separate. At an A/B interface, diffusion is limited to a narrow
region whose thickness is less than the radius of gyration of the chains. When
Na=Nb, the miscibility of A in B is determined only by the critical X value χ0,
given by: X — — (z~ 2)(£aa + £bb —2eAB)/2kT (16) Xc=VN (17)

238 Welding and fracture of polypropylene interfaces When χ>χ0, then the blend
is immiscible; when χ<χ0, mixing occurs, xc is independent of temperature, the
temperature dependence of χ is determined by: x{T) = a/T+b (18) where a and b
are constants and T has units of degrees Kelvin. For example, with the
PS/PMMA pair, the temperature dependence of χ was determined by Russell
(private communication) as: χ(Τ) = 3.902/T+0.0284 (19) As the temperature
increases, χ decreases and compatibilization is enhanced. The temperature at
which compatibility is obtained is known as the upper critical solution
temperature (UCST). While polymers such as PS/PMMA have great difficulty in
forming homogeneous blends, we may investigate the limited mixing that can
occur at the interface. This subject has received considerable attention from
Helfand and colleagues [16-18]. They examined the incompatible interface
between two immiscible polymers of infinite molecular weight and described the
diffusion of polymer A in polymer B by solving the diffusion equation of a
random walk in a potential field created by the incompatible monomer. The
extent of the interfacial mixing in determined by a balance of two forces in the
simple case. The compressed configurations of the chains in the surface layer
provide an entropic driving force to diffuse across the interface by a distance d,
and expand the random coil dimensions. This action is counter¬ balanced by the
unfavorable mixing of incompatible monomers. The dominant entropic force for
this limited diffusion is not the combina¬ torial entropy but rather the
this limited diffusion is not the combina¬ torial entropy but rather the
configurational entropy change at the surface. The entropy change for a random
coil chain of n steps which relaxes from a surface reflected configuration to a
completely random configuration with end-to-end vector dO09 as shown in
Figure 8.4, is: AS = klnn (20) Figure 8.4 Conformation of a polymer chain at an
interface before and after diffusion.

Polymer interfaces 239 However, the enthalpy change for mixing this chain of
length n with incompatible monomers is: Minimizing the free energy ΔΑ = AH
— TAS with respect to n and solving for the equilibrium value of d„ &n1,2y
Helfand’s solution for the equilibrium interface thickness d^ is: where b is the
statistical segment length, which is about 0.65 nm for PS and PMMA. For the
PS/PMMA interface at 140 °C, Equation 22 gives doo = 2.7 nm. The
significance of this layer thickness on mechanical proper¬ ties of the interface is
discussed later. Multicomponent polymer interfaces Several important
multicomponent interfaces are obtained with polymer blends and composites.
With rubber toughened epoxies and thermoplastic matrices, one has a mixture of
phase separated (typically spherical) small rubber particles in a higher modulus
matrix. For example, the tough ABS plastic consists of a copolymer matrix
(acrylonitrile-styrene) with a small volume fraction of rubber particles
(butadiene with styrene grafts for adhesion to matrix). The rubber particles can
have a complex microstruc¬ ture with inner cores of concentric layers of varying
moduli and an outer skin with a grafted molecular layer to enhance adhesion
with the PSAN matrix. In a uniaxial stress field, the lower modulus particles
cause a stress concentration in the brittle matrix at the equatorial poles
(perpendicular to the stress field) of the particle. The local high stress initiates a
craze and the rubber particle deforms to accommodate the craze opening. If the
adhesion of the rubber particle to the matrix is sufficient to prevent the craze
from reaching the critical crack opening displacement, which could result in
catastrophic failure, then the energy adsorption mechanism has been suc¬
cessfully accomplished. The craze propagates safely and usually terminates in
another particle. Other toughening mechanisms in these composites involve
cavitation, crack arrest or blunting and shear yielding of the matrix ligaments
between particles. The blending of plastics with rubber particles can result in a
substantial enhancement of properties, for example, the creation of bullet-proof
glass from brittle PMMA and PC. The process is delicate in that one is deliber¬
ately inducing extensive microscopic damage as the energy absorbing
mechanism while attempting to prevent the propagation of a single fatal crack.
Considerable attention is being paid to the design of these composites, with
emphasis on the particle microstructure and adhesion characteristics, and
emphasis on the particle microstructure and adhesion characteristics, and
ΑΗ&χη (21) <*« =2£>/(6χ)1/2 (22)

240 Welding and fracture of polypropylene interfaces the success of the


composite design is typically evaluated from the bulk properties. However, in
larger injection molded parts (such as automobile body parts) and compression
molded laminates, the interface properties may be radically different from the
bulk properties. In the following section, we discuss several cases involving
multicomponent interfaces. Interfaces of two-component blends with particles
Consider the two-component blend consisting of matrix A, with volume fraction
φΒ of spherical particles of radius R, as shown in Figure 8.5 where PS spheres
are dispersed in a PE matrix. The number of particles per unit volume Nw is
given by: Νν = 3φΒ/(4πΛ3) (23) Figure 8.5 The surface of PE blend containing
PS particles (courtesy of P. H. Geil). When the particles are randomly
distributed, the number touching unit area of free surface NB is determined from
the number of particles whose centers are within a depth of radius R from the
surface as: NB=3(/)B/(4nR2) (24) where 0<φΒ<ζ1. ‘Touching’, in this case,
allows protrusion of the spherical particle through the matrix surface (see Figure
8.5) up to a distance not

greater than 2R. The average surface area occupied by a sphere is 2/3 nR2 such
that the surface area fraction FB occupied by the particles is simply: ^b = 0b/2
(25) When two of these surfaces are brought into contact, the resulting interface
is comprised of three component interfaces namely, A/A, B/B and A/B. Thus,
we have two symmetric and one asymmetric interface such that areal fractions:
FaA + F BB + F AB — 1 (26) The area fraction of each component interface can
be estimated as follows. At the interface we now have 2NB particles which only
contribute to either A/B or B/B interfaces. Therefore, the fraction of A/A
interfaces is 1 —(Fab +^bb)> or: F AA — 1 — Φβ (27) The fraction of B/B
contacts will be proportional to φ2 with a propor¬ tionality constant of order
unity such that: FBB~ φ2 (28) and FAB ~ Φ φ2 (29) Thus, for a typical blend
containing 10% by volume of particles, the composite interface will consist of
90% pure matrix/matrix, 9.9% matrix/particle and 1% particle/partide interfaces.
To an excellent first approximation, one can ignore the φ2 term and consider the
influence of the particle/matrix interface strength on the pure matrix interface.
When the pure matrix is healed to a strength comparable to the virgin state, then
the healed composite strength should be identical to the bulk virgin strength.
However, two situations can arise which will adversely affect the interface
strength; these involve either a matrix rich skin on the surface or a particle rich
surface. A matrix rich skin typically develops during compression and injection
molding where the blend fluid comes in contact with a hard wall and the random
molding where the blend fluid comes in contact with a hard wall and the random
distribution of particles protruding from the surface is not allowed. The matrix
forms a skin when the particles bounce back from the hard wall. In that case, a
crack may propagate continuously through this particle depleted layer when the
interface is formed from two such surfaces. Particle rich surfaces can form at
internal weld lines due to the fountain effect inducing particle separation; at
semicrystalline surfaces by a type of zone refining when crystallization occurs in
a temperature gradient; and in single and twin screw extruders due to high shear
of the fluid between the screw and the barrel wall and for other reasons usually
associated with the Polymer interfaces 241

melt flow. When this situation develops, the resulting interface can be very
weak. In complex parts, this problem can be minimized to a certain extent. By
controlling the mold design via location of the injection gates, the weld lines can
usually be located at planes that are not orthogonal to anticipated service
stresses. 8.3 THEORY OF FRACTURE The Griffith approach to brittle tensile
fracture for a material with a crack of length a (Figure 8.6), is given in terms of
the stored elastic strain energy U, and the energy S to create new surface area.
This theory forms the basis of linear elastic fracture mechanics (LEFM). It states
that the incremental change in strain energy d U with crack length da, exceeds
the energy to create surface area dS. For a fracture specimen as shown in Figure
8.6 the critical stress at fracture ac, is obtained as: Figure 8.6 Compact tension
fracture mechanics specimen containing a crack of length a, width B and length
W. The quantities G1C and the modulus E are constants such that the fracture
variables of crack length a and critical stress ac can be separated out in the form:
alna = EGlc (31) Since the term EGiC is constant, then the left-hand side must
be also be constant and its square root is given in terms of the stress intensity
factor K1C: Klc = ac(na)112 (32) Welding and fracture of polypropylene
interfaces 242

Polymer entanglements 243 which has typical units of MPam1/2. K1C is also a
measure of the stress concentration factor at the crack tip when fracture occurs
and is commonly referred to as the fracture toughness. G1C and Klc are related
to each other by: G1C = Klc/E (plane stress) (33) and in plane strain, the right-
hand side is multiplied by (1 — t?2). Several important proportionalities can be
deduced from the Griffith theory. For example, the strain energy density U0 in
the material is given by the uniaxial approximation: ϋ0=σ2/2Ε (34) Thus, at the
critical stress, we have from the above relations: Glc=2naU0c (35) where U0c is
the critical strain energy density. The relation G1C ~ U0c will be used later,
when we introduce a microscopic mechanism whereby we compute the stored
when we introduce a microscopic mechanism whereby we compute the stored
strain energy required to pull a chain from the interface. In the Dugdale model of
fracture where a crack propagates through a thin deformation zone at the crack
tip, the fracture energy is determined by: Glc = a*d (36) where σ* is the average
stress creating the deformation zone at the crack tip and δ is the crack opening
displacement at fracture. From the Griffith theory we can argue that σ0πσ* such
that: δπσ*!Ε (37) For welding at interfaces, the modulus E is independent of
both time and molecular weight such that we expect δ and σ to have similar
scaling laws. 8.4 POLYMER ENTANGLEMENTS The Griffith idea has
considerable value if one can determine how stored energy is consumed in
forming the fracture surface. To address welding and healing problems, a stored
strain energy approach to fracture was adopted which considers both chain
pullout via disentanglement and chain fracture mechanisms at the interface. We
first consider an entanglement model, since entanglements are the most
important structural feature in polymer melts. 8.4.1 Entanglement model for
random coil polymers We now develop an entanglement model which can be
used to provide mechanical connectivity and relate the interface structure to the
breakdown

244 Welding and fracture of polypropylene interfaces process of the deformation


zone at a crack tip. Entanglements develop from the interpenetration of random
coil chains and are important in determining rheological, dynamic and fracture
properties. The subject has been viewed by Graessley [27] and Aharoni [28],
and, more recently, was investigated by Kavassalis and Noolandi [29]. Of
particular importance is the role of entanglements in controlling melt viscosity
where the zero shear viscosity, η0, at low molecular weights M, behaves as a
simple Rouse fluid with η0 «Μ, and at high molecular weight η0&Μ3Λ. The
cross-over to a high viscosity entangled fluid occurs at the critical entanglement
molecular weight Mc. A similar effect is seen for the fracture energy G1C of
glassy polymers such that when M<MC, G1C~M, and when M>MC, G1C«M*,
where x is of the order of 2-4 with a pronounced transition at Mc [4]. Both
viscosity and fracture processes are related to the transmission of forces through
unit area (stress) and in this chapter we examine the molecular structure capable
of transmitting such forces in an entangled polymer. An entangled amorphous
linear chain network is shown schematically in Figure 8.7. The bridge theory of
connectivity in an amorphous network [3,4] requires that the number of chain
segments p crossing any load- bearing plane exceeds the number of chains n by
3p>n. When p<3n, a Figure 8.7 Entanglements in a polymer melt. The bold
section represents the entanglement loop.

Polymer entanglements 245 network cannot form and the chains readily slip
Polymer entanglements 245 network cannot form and the chains readily slip
apart by Rouse motion. When p > 3n, the chains are sufficiently interpenetrated
to form an entangle¬ ment network and relaxation occurs by diffusion in the
presence of entangle¬ ment constraints. At p&3n, the polymer chains are
critically connected and the average bridge structure has three crossings. This
structure constitutes the unit ‘mesh’ in the network. The existence of an
entanglement network in an isotropic concentrated melt can be explored by
counting the number of bridges and chains intersecting an arbitrary plane, as
shown in Figure 8.7. A bridge is a segment of chain which crosses the plane
three times (bold section in Figure 8.7). It is sufficiently long to complete one
circular loop through the plane. The bridge is capable of transmitting forces
across the plane in the melt for a time dependent on the relaxation of this chain
segment. The number of chain segments crossings per unit area p is independent
of molecular weight in the virgin state, p »M°. However, the number of chains n
intersecting the plane decreases with increasing molecular weight as n«M_1/2.
Thus by varying the molecular weight, we can reach a state where the number of
bridges is comparable to the number of chains. We define the number of bridges
per chain pCi as: such that when pc = 1, p/n = 3, where the factor of three
considers the three crossings per bridge. When pc<l, each chain contributes less
than one bridge and the melt is not connected in a network. When pc > 1, an
entanglement network exists and we can determine Mc from the condition that
pc = l, or p = 3n. This argument can be readily tested because if the hypothesis is
valied, Mc can be determined without any fitting parameters for all random coil
polymers, simply by calculating p and n, as follows. The number of chains
intersecting an arbitrary plane through random coil chains is given in terms of
the molecular weight and random coil parameters as [3, 4]: where NA is
Avogadro’s number, p is the density, b is the bond length, M0 is the monomer
molecular weight, is the characteristic ratio and j is the number of bonds per
monomer. Here n = Nv<D), in which the number of chains per unit volume Nv
=pNJM, and </)) = 1.31 Ni,2b is the diameter of a random coil chain. The
number of chain segments crossing unit area is given by p = l/a, where a, is the
projection of the cross-sectional are of the chain segment on the plane. When p =
3n at Mc, then we obtain the critically connected network when the following
condition is fulfilled: Pc = 2(p/n-l) (38) n=l.3l[C'Oj/{M0M)']il2bpNA (39) 3an=
1 (40)

246 Welding and fracture of polypropylene interfaces The cross-sectional area of


the chain segment a could be determined from the monomer volume using a =
(M0/pNA)2/3. However, this approach tends to ignore aspect ratio and any chain
structure due to conformation, such as helical segments. A more useful approach
structure due to conformation, such as helical segments. A more useful approach
is to determine the molecular area from unit cell dimensions (when they exist)
via the volume/length ratio, a = 21,2V/L, such that: a = 21,2zM0/(cpNA) (41)
where the length L is the c-axis length (backbone direction) of the unit cell, the
unit cell volume V=zM0/(pNA), and z is the number of monomers per c-axis
length. The factor of 21/2 accounts for the random orientation of chains crossing
the plane, compared with their perfect alignment in a crystal. During a
mechanical deformation, the change in the orientation factor of the chains must
be considered as the melt becomes anisotropic and the entanglement density
changes in different directions [4]. Substituting for a, n and p, and solving for
Mc at pc = 1, we obtain the expression for the critical entanglement molecular
weight as: Mc = 30.89 {zb/C)2) M0 C*, (42) in which all quantities are known,
or can be estimated. Also, {zb/Q2j»1 for many polymers, and an excellent
approximation for Mc for vinyl (j = 2) polymers is obtained as: Me^OCeoMo
(43) Note that Mc^u2C00, which means that stiffer chains generally have higher
Mc values. Nc is related to Ne via Nc=27/4 Ne. If we identify Ne as the length
of chain to form one loop, then our theory provides: Nc=9/4Ne (44) Generally,
one finds from experiment that Nc &2Ne, with minor deviations. The relation
for the molecular weight dependence of the number of bridges per chain is
pc^M1/2. Since pc = l at Mc, it follows that the proportionality constant is
M“1/2 such that: Pc =(M/MC)1/2 (45) Thus, the total number of bridges per unit
area in the virgin state is given by Pao =Pcn, as: Poo = 1.31[C
J/(M0Mc)]1/2VJVa (46) The entanglement model presented above, unlike all
others, contains no adjustable parameters; it only contains known constants and,
therefore, its utility can be tested by comparing the theoretical Mc with
experimental Mc values. In the following sections, we examine the model’s
ability to predict

Polymer entanglements 247 Mc> its concentration dependence and other


properties influenced by en¬ tanglements. 8.4.2 Critical entanglement molecular
weight, Mc In this section we examine the ability of the entanglement model to
predict values for Mc using: Mc = 30.89(zb/C)2jM0Coo (47) and, where
appropriate, using the short formula Mc^30M0Coo (see Table 8.3). In the
calculations below, the crystal data were obtained from [30] and Mc data from
review tables in [28, 31]. Table 8.3 Theoretical and experimental critical
entanglement molecular weights, Mc Polymer M ^Experiment) Mc(theory)
Polyethylene 4 000 4 200 Polystyrene 31200 32 000 Polypropylene 7 000 7 600
Polyvinyl alcohol 7 500 7 000 Polyvinyl acetate 24 500 25 000 Polyvinyl
chloride 11000 10 700 Polymethylmethacrylate 18 400 18 000 Polyethylene
oxide) 4 400 5 000 Polypropylene oxide) 5 800 5 000 Polycarbonate 4 800 4 300
oxide) 4 400 5 000 Polypropylene oxide) 5 800 5 000 Polycarbonate 4 800 4 300
Polyethylene b = 0.154nm (C-C bond length); c = 0.255 nm (c-axis unit cell
length); z = 1 (number of monomers per c-axis length); M0 = 28 (monomer
molecular weight); = 6.7 (characteristic ratio); j = 2 (number of C-C bonds per
monomer); Mc »4000 (theory) and Mc = 3800 (experiment). The agreement
between theory and experiment is good. A planar zig-zag (all-trans)
conformation of the PE chain in an orthorhombic unit cell was used to describe
the cross-sectional area of the molecule. In the melt, we can expect many gauche
conformers to be populated. However, the theory is not very sensitive to the
exact average conformational details because of compensating effects in the
helix term a, given by: a = C/(zjb) (48) in Equation 47. Here, large excursions of
the internal rotation angles only result in small changes in the length c, and when
z increases, c also increases. However, small changes in a are amplified by the
a2 contribution to Mc. The

248 Welding and fracture of polypropylene interfaces equation for Mc may be


rewritten: Mc = 30.89 Coo M0/(a2j) (49) The factor a2j is the difference
between the exact and short formula for Mc. Note that a2j«l for a 3/1 helix with j
= 2. For a 2/1 all-trans PE helix, a = 0.827, but for a 3/1 trans-gauche PE helix, a
= 0.703. The latter conformation is improbable and would give Mc «5500. The
theory predicts that we get an increase in Mc with increasing cross- section or
helix of the chain. The short formula gives Mc = 30CooM0 = 5600. Thus, for
molecules related to the PE family, e.g. LDPE, HDPE, LLDPE, and polymers
containing large CH2 sequences, such as polybutadiene, nylon 6, nylon 6-6,
nylon 6-10, etc., we expect to have similar Mc values in the vicinity of Mc
«4000-5000, with some small variation due to differences in and M0. When the
polymer molecule becomes highly polar or rigid, the theory is not expected to
work since new mechanisms of transmitting forces are introduced.
Polypropylene b = 0.154nm (C-C bond length); c = 0.65nm (3/1 isotactic helix);
z = 3 (number of monomers per c-axis); M0 = 42 (monomer molecular weight);
Coo =5.8 (characteristic ratio); j = 2 (number of C-C backbone bonds per
monomer); Mc «7600 (theory) and Mc «7000 (experiment, isotactic PP).
Although the isotactic configuration can adopt many local conformations in the
melt, the average helix factor of a = 0.7 and a2j = 0.99 is not expected to be
changed. The short formula gives Mc =7300. 8.4.3 Deformation and
disentanglement In this section, we explore how entanglement networks can
strain harden and disentangle to produce fracture at interfaces. Strain hardening
When a polymer in a concentrated melt is subjected to a uniaxial extension, the
maximum draw ratio am which a chain of molecular weight M can experience is
determined by: am=L/R = N1,2otb (50) where L = Nboch is the extended chain
determined by: am=L/R = N1,2otb (50) where L = Nboch is the extended chain
contour length, ab is the draw ratio of the statistical segment repeat unit with
length b, and R = N1,2b is the end-to-end vector. At am, the chain is fully
extended and can only extend further by small amounts involving much higher
stresses along the back¬ bone bonds of the chain. This is called strain hardening
and the resulting melt is in an optimally oriented and completely disentangled
state.

Polymer entanglements 249 When a step strain is rapidly applied to an entangled


melt, strain hardening first occurs at the level of entanglements. As described by
Donald and Kramer [32], the strain is taken up by the ‘slack between entangle¬
ments’ such that ame»Mtli2. If Me «Mc/2, and using our value of Ne «30/2, we
obtain the universal result [4]: a me «40“ 1/2 (51) It is an interesting prediction
of the entanglement model that all polymer melts strain harden at about the same
draw ratio of 4 at polymer concentra¬ tion φ = 1. Donald and Kramer examined
the draw ratio of craze fibrils af in several glassy polymers using a densitometry
analysis of TEM images. They found fibrillar draw ratios in the range of 2-4.
Recently, Kunz (private communication) and Yang, Kunz and Logan [33], using
atomic force microscopy (AFM) study of crazing in polystyrene and othe glassy
poly¬ mers, found a maximum draw ratio of about 4, consistent with this model.
Disentanglement and fracture We apply the entanglement model to examine
disentanglement as a function of a uniaxial draw ratio, a, in polymer melts. The
number of chains crossing a plane normal to the applied strain increases linearly
with draw ratio as: n{oc) = an00 (52) where n^ is the number of chains per unit
area at a=l. The number of bridges increases more slowly with draw ratio since
p&l/a, where a is the cross-sectional area of a chain; a decreases in an affine
manner with increasing draw ratio to give: ρ(α)=ρ0Οα/(α2 + 1/α)1/2 (53) in
which p00/21/2=(M/Mc)1/2 is the number of bridges at a=l. At the critical point,
p(<x) = n(oi) and from Equations 52 and 53 the critical draw ratio is determined
by: ac2 + l/ac=2M/Mc (54) Solving this cubic equation, we obtain ac in terms of
M/Mc as: ac =(8/3 M/Mc)1/2 cos{l/3 cos _1[-i(3Mc/2M)3/2]} (55) where ac is
only dependent on M/Mc. This equation predicts that ac = l when M/Mc = 1, and
ac«4 when M/Mc« 8. The latter value corresponds to the maximum draw ratio
am, to cause strain hardening between en¬ tanglements (Equation 51). The
dependence of ac on M/Mc is shown in Figure 8.8. The significance of ac is that
it represents the minimum strain necessary to cause fracture by disentanglement
and we thus expect G1C» (<Xc-l)2.

250 Welding and fracture of polypropylene interfaces Figure 8.8 Critical draw
ratio for disentanglement of amorphous polymers versus molecular weight.
When M »MC, a useful approximate solution to Equation 55 gives the critical
When M »MC, a useful approximate solution to Equation 55 gives the critical
draw ratio for disentanglement as: ac*(2 M/Mc)1/2 (56) We see in Figure 8.8
that this approximation is reasonable where a slope of \ is obtained over most of
the data range, except near Af/Mc«l, where the latter equation incorrectly
predicts that ac»21/2. Disentanglement is considered to occur as shown in Figure
8.9, where the response of an entangled chain to a constant draw ratio a is
shown. The average entangled chain with M/Mc > 1 is uniaxially deformed to a
constant draw ratio a. The extension is accommodated by extending the random
walk (slack) between entanglements. The entanglement points deform affinely
and the chain stores elastic strain energy Uc &Ε(α— l)2. Rouse dynamics causes
a retraction of the extended chain and the stored strain energy begins to release.
As the chain shortens towards its equilibrium contour length, it begins to lose
entanglements. The time dependence of the retraction process can be
approximated as a simple exponential such that the length of the chain L(t) as a
function of time is: L(t) = L0 exp — t/zR0 (57) where L0=MRe/Me is the initial
unperturbed length and tro is the Rouse relaxation time of the chain. When the
chain retracts to a critical length Lc»2acRe, then each chain possesses one bridge
and the network becomes critically connected. The time to achieve the critically
connected state corresponds to the failure time of the entanglement network, rf,
and can be described with the approximate

Disentanglement and fracture 251 V V Figure 8.9 Disentanglement mechanism


in amorphous polymers, (a) Slack between entanglements is stretched out to a
draw ratio of 4; (b) Rouse relaxation of the chain occurs, resulting in chain
retraction from entanglements; and (c) chain retracts to a critically connected
state and the polymer fractures. form of Equation 56 by: τΓ«τΚ01η(α?/2) (58)
Thus, when M = 8MC, and a = 4, τΓ»τΚ0. When M<8MC, then τΓ<τΚ0. Using
the exact form of Equation 56, we obtain τ{ = 0 when M = MC for draw ratio α=
1. When M>8MC, the chains cannot disentangle at the Rouse time and begin to
re-entangle due to reptation. Thus, bond rupture would be necessary to complete
the fracture process and the value of M*»8MC sets an upper limit for
disentanglement. 8.5 DISENTANGLEMENT AND FRACTURE OF
INTERFACES 8.5.1 Welding at interfaces For welding amorphous polymer
interfaces, the average diffused contour length </> controls the disentanglement
process. The contour length has

contributions from reptation L{t/Tr)1/2 and Rouse dynamics Lc which combine


approximately via the weighted segmental dynamic contribution as: Figure 8.10
Predicted fracture energy versus welding time for polymer interfaces with the
molecular weights indicated. Fracture energy J/m2 Welding time t/Tr M/Mc
molecular weights indicated. Fracture energy J/m2 Welding time t/Tr M/Mc
Welding and fracture of polypropylene interfaces 252 Thus, when t~0, <l>~Lc.
Rapid segmental diffusion results in an initial contour length corresponding to
Mc at times of the order of the relaxation time τβ of an entanglement segment.
The healing time dependence of the critical draw ratio is obtained from ac*
[2</>/Lc]1/2 as, such that at t = Tr, occ is of order (M/Mc)1/2. At fracture, the
strain energy density U0 is related to G1C through Equations 31 and 34. Since
U0 =E(occ —1)2/2, then we have G1C(£)~ E(ac — l)2 and in the approximate
form: Figure 8.10 shows a log-log plot of the exact calculation for G1C versus
t/Tr. The data are not exactly linear but behave approximately as: With
increasing molecular weight, the data converges on the pure reptation prediction:
Note that as M approaches Afc, G1C goes to zero for all welding times. (59)
(60) (61) (62) (63)

Incompatible semicrystalline 253 8.5.2 Molecular weight dependence of fracture


As t/TT goes to unity in welding experiments, we recover the molecular weight
dependence of the fracture energy in the virgin state. From Equation 61 we
obtain the useful relation: G1C*M/MC[1-(MC/M)1/2]2 (64) in the range
MC<M<8MC. When M>8MC, we expect G1C~M°. It is interesting to note that
a plot of log G1C against log M will have an average slope between 2 and 3 in
this molecular weight range, as shown in Figure 8.11. When the fracture energy
reaches the plateau, complete disen¬ tanglement can be accommodated by bond
rupture. We know that bond rupture essentially adds nothing to the total fracture
energy. Figure 8.11 Fracture energy versus molecular weight. Finally, the critical
stress intensity factor K1C, fracture stress σ, and fracture strain e, are all
proportional to G^2 and hence from Equation 64: (Κ1€)σ,ε)^(Μ^2-Μΐ/2) (65)
which is in a convenient form for experimental investigation. 8.6
INCOMPATIBLE SEMICRYSTALLINE INTERFACES The object of this
section is to consider the mechanisms of strength development at incompatible
crystallizable interfaces formed with PE and PP. Factors involved in the strength
development of amorphous interfaces will be considered in addition to local
crystallization effects proposed by Galeski and Bartczak [34, 35] and Galeski
and Piorkowska [36]. In the latter process, volume contraction due to
crystallization of spherulites in the

254 Welding and fracture of polypropylene interfaces vicinity of the interface is


important and can result in a breakup of the original interface plane with a
subsequent influx of melt across the interface. Crystallization of the influxes can
produce a mechanically inter¬ locked interface with a fracture energy
considerably greater than that predicted by our model for incompatible
amorphous interfaces without crystallization. Isotactic PP and HDPE are
amorphous interfaces without crystallization. Isotactic PP and HDPE are
thermodynamically incompatible with one another. Both theory and experiment
show that local segment inter¬ diffusion can occur as long as the temperature is
high enough to permit a significant degree of segmental mobility. Theoretical
predictions of the interface thickness, d^^-lOnm, were made by Helfand and
colleagues [16-18], Kammar [37] and Shilov [38]. Segments crossing the inter¬
face due to limited interdiffusion and mechanical interlocking due to local
crystallization contribute to the strength of the interface. The purpose of the
present work is to observe how the interface forms and strength develops
between the two polymers after contact is made in the melt and crystallization
proceeds under various thermal histories. 8.6.1 Interface preparation and thermal
history The polymers studied and their properties are listed in Table 8.4. LLDPE
and HDPE were used to investigate structure development in the interface and
HDPE was used for the butt welding mechanical studies. The PP/PE interface
was prepared by laminating plates of PP and PE together in a Carver hot press
with varying conditions, including contact time and temperature. Further details
are given in [20-22]. Table 8.4 Test materials Material Code MI TJ°C) Mwx
10“5 Appearance PP Shell 5520 5.0 161 2.0 Pellet HDPE Chemplex 5853 1.3
131 1.3 Pellet LLDPE Grsin 7144 20.0 122 Powder LLDPE Grsin 7147 50.0 121
Powder Following melting at 195 °C for 30 min, the laminate was cooled down
in two ways. The first method involved isothermal crystallization at 136 °C (or
at 141 °C) for times of up to 1.5 h (the temperature fluctuation of the hot press
was ±3°C within 6 min). The other method involved rapid cooling from the melt
to room temperature at an average cooling rate of around 20°C/min. Further
details of the thermal histories are shown in Figure 8.12.

Incompatible semicrystalline 255 Figure 8.12 Thermal history for the PP/PE
laminate isothermally crystallized at 136 °C for (a) 0.5 hr, (b) 1.0 hr, (c) 1.5 hr
and (d) 2.5 hr, and (e) rapid cooling at 10 °C/min. 8.6.2 Structure development at
semicrystalline interfaces Based on the studies of amorphous polymer interfaces
by Wool and colleagues [1-4, 20-22] and on the crystallization behavior of
semicrystal¬ line polymers, we expect the interface formation of PP/PE to occur
in five steps as shown in Figure 8.13: (a) surface rearrangement, (b) wetting, (c)
interdiffusion, (d) crystallization and (e) solidification. The steps are discussed
individually below. Surface rearrangement Surface rearrangement phenomena
occur at the surfaces both before and after contact has been made to form the
polymer-polymer interface. We are primarily concerned with structure formation
or relaxation near the surface, chain end distribution functions, surface
segregation of impurities, low molecular weight species and nucleating agents,
and non-equilibrium chain configurations. With regard to crystal nuclei memory
and non-equilibrium chain configurations. With regard to crystal nuclei memory
effects, we know from Rault’s data [39] that the molten PP/PE laminate is in an
equilibrium state (with loss of previous crystallization history) after holding at
195 °C for 30 min. The relaxation time for PP with Mw = 105 is around 30 min
if the previous crystallization was done at a cooling rate of 20°C/min. The
conditions for making the plates and laminates in our experiments were similar
to these ones. Surface rearrangement effects also involve the non-equilibrium
configur¬ ations of the melt chains due to the reflecting boundary condition of
the surface ‘compressing’ the normal Gaussian coil shape of the amorphous
chains which lie within a distance of the radius of gyration of the surface. The
reduced entropy of the surface layer of chains acts as a driving force to permit
limited interdiffusion at incompatible interfaces. Rouse-like

256 Welding and fracture of polypropylene interfaces (a) (b) Rearrangement


Wetting and interdiffusion Crystallization Solidification Figure 8.13 Schematic
diagram of the stages of the interface formation of an A/B polymer pair during
crystallization. relaxation of the compressed configurations has been shown by
Zhang and Wool [40] to cause rapid interdiffusion up to distances of the order of
the entanglement spacing in polymer melts. Wetting This step involves the
establishment of contact of PE with PP at distances of the order of the van der
Waals radius. Surface roughness and impurities play an important role in
determining the rate of wetting and contact of the

Incompatible semicrystalline 257 surfaces to form the interface. Usually the


spreading coefficient, Fij9 is used to estimate the wettability of phase-i spreading
on phase-j as [41]: F^rj-r.-ry (66) where Γ is the surface tension. The following
values of the surface tension at 140 °C were used to calculate Equation 66: ΓΡΕ
=0.0288 J/m2; ΓΡΡ = 0.0231 J/m2; FPP/PE =0.011 J/m2, and thus FPP/PE=
0.046 J/m2 [41]. From a thermodynamic point of view, since F> 0, we infer that
PP can spread on the PE melt and that good contact can be achieved at the
interface. Therefore, interdiffusion can subsequently occur in the wetted areas. If
the time dependence of wetting is comparable to the time depend¬ ence of
interdiffusion then the extent of diffusion is determined by the convolution
product of wetting and diffusion functions [1]. Interdiffusion Using the Flory-
Huggins theory [26], we obtained χΟΓ = 5.6χ 10"4 for the PP/PE system studied
with Mw = 105. Polymer pairs with χ>χοχ should be incompatible. At long
contact times, the equilibrium interdiffused thickness of the interface, d<*>, was
derived as: doo = 2h/(6xab)1/2 = 5.4 nm (67) and Xah = V/RT(d&-db)2 (68)
where b is the effective bond length of a statistical segment in the melt, with
typical values around 1.0 nm, d is the solubility parameter expressed in units of
typical values around 1.0 nm, d is the solubility parameter expressed in units of
J1/2/cm2/3, V is the molar volume (cm3/mol), R is the gas constant (J/mol K)
and Fis the temperature (K). The data needed for the calculation of %ab in this
study are available at 140 °C from [32] as follows: dPP = 16, dpE = 17, K =
32.8, R = 8.314 and Γ=413 K, which gives xab =0.011 and doo = 5.4 nm. The
calculated values show that xab is much larger than the value of
xcr(0.011»0.00056), such that the PP/PE polymer pair used here should
represent an incompatible interface in the melt. The end-to-end distance R of PE
chains in the melt is obtained as R = 33.8nm and the radius of gyration as
<F2/6)1/2 = 13.8 nm. It is clear that the PE molecules which diffuse across the
interface to a 2.7 nm depth (depth = doo /2) achieve distances that are a small
part of the chain dimen¬ sions (13.8 nm). Thus, a thin layer of mixed PP and PE
forms in the melt prior to crystallization. When crystallization occurs, the fate of
the layer will be sensitive to the crystallization mechanism with opportunities for
either further phase separation induced by crystallization, or entrapment via
entanglements in the amorphous component of the semicrystalline structure.

258 Welding and fracture of polypropylene interfaces Crystallization Both PP


and PE are semicrystalline polymers with a spherulitic structure Voids and
defects are induced by the density change due to crystallizatioi occurring in the
melt. For example, the volume contraction for PP coul< reach 10%. With the
difference in melting points and crystallization rates, i is possible for one
polymer to crystallize while the other is still in the liqui< phase. The volume
contractions associated with the random nucleation o spherulites near the
interface have the effect of pulling the other polymer ii the liquid phase across
the interface and creating interspherulitic influxes This effect, known as local
crystallization, [34-36], often results in j mechanically interlocked interface.
Intraspherulitic influxes also develop which contribute to the strength of the
interface. The surface morphology of semicrystalline polymer has been shown to
b( dependent on the nature of the mold surface, the cooling rate, and process ing
conditions [34, 41-43]. Because of differences in both melting points anc
crystallization rates, the PP/PE laminate can be formed with either the PI
crystallized against the PE melt or the PE crystallized against the PP melt
Consequently, the interface strength may change with the thermal history Also,
it is well known [44, 45] that melt cooled PP has a three-phase systen -
amorphous, smectic, and crystalline - and these may play a role in the
macroscopic properties of the interface region. Solidification Partial
solidification of the PP/PE laminate first occurs with crystallization and the final
solid interface is obtained by further cooling of the sample. The interface
morphology, and in turn the interface strength, will depend on the two steps. If
the PE influxes have fully developed during slow cooling, further cooling will
the PE influxes have fully developed during slow cooling, further cooling will
not have a significant effect on the interface morphology. In contrast, when
cooling rapidly from the melt to room temperature, the interface morphology
will be determined by the relative crystallization rates on both sides of the
laminate. These two cases were studied in our experiments and are discussed in
the next section. 8.6.3 PP/PE interfaces Isothermally crystallized PP/HDPE and
PP/LLDPE The laminates crystallized at 136 °C were subjected to varying
crystalliza¬ tion times to produce specimens with different interface structure.
SEM micrographs of etched sections cut perpendicular to the interface showed
that the interface shape changes from being flat to an irregular wave shape with
increasing crystallization time. The PP surface of a fractured PP/HDPE interface
is shown in Figure 8.14. With increasing time, the size

Incompatible semicrystalline 259 Figure 8.14 SEM micrographs showing the PP


surface of a fractured PP/HDPE interface with crystallization times at 136 °C of
(a) 0.5 hr, showing one spherulite and fine fibers; (c) 1.0 hr, showing many more
spherulites growing near the interface; and (d) 1.5 hr, showing well-developed
spherulites with ‘clean’ shell debonded surfaces and interstitial areas due to
polymer segregation and volume contraction, (b) is a higher magnification of the
fibers shown in (a). of the spherulites increases and their number decreases
appreciably. Nu¬ cleating agents are an important variable in the determination
of the interface structure. A high nucleation density would result in structures
similar to that shown in Figure 8.14b, whereas stronger interfaces form with
fewer spherulites (Figure 8.14d). The material remaining in the interstices
between spherulites is clearly visible. Figure 8.14 shows that adhesive fracture
occurred in the interface plane with extensive pull-out and fracture of (PE)
fibrils. The fractured surfaces were well matched with each other and became
more undulating with increasing crystallization time. Well defined spheres, holes
and highly drawn (PE) fibers appear on the PP side. DSC measurements were
conduc¬ ted with the material from the fracture surface of the PP side. A 132 °C
melting peak was found in the DSC curve which did not exist in the pure PP and
which supports the hypothesis that fracture remnants of the HDPE influxes
resided in the PP side.

260 Welding and fracture of polypropylene interfaces The butt strength tests
indicate that the interface strength increases with increasing crystallization time,
as shown in Figure 8.15. The yield strength of HDPE is about 20 MPa, which is
about twice the strength of the interface formed at 1.5 h. Although the
comparison of fracture stresses is complicated by the fracture mechanism
differences, the results are highly supportive of a considerable enhancement of
differences, the results are highly supportive of a considerable enhancement of
the interface strength due to mechanical interlocking. The ratio of the fracture
stresses is much greater than the ratio of the influx material cross-section/area
that one could deduce from Figure 8.14. Thus, we expect contributions to
strength from intraspherulitic entrapment as well as the interspherulitic influxes.
' i 1 1 1 1 0.00 0.50 1.00 1.50 2.00 Isothermal crystallization time (hr) Figure
8.15 The butt tensile strength of the PP/HDPE laminate as a function of the
isothermal crystallization time at 136 °C. An estimate of the effect of influxes on
the interface strength σ and fracture energy G, can be approximated for the fully
crystallized interface using simple geometric arguments [29]. We let the
interface be composed of PP spherulites of radius r, in which the interstitial area
is completely occupied by PE influx fibrils. The centers of the circles
representing the contiguous spherulites are separated by a distance 2r. The
fracture energy of the interface is given by: GttN{G{ (69) where N{ is the
number of influxes per unit area and Gf is the energy to fracture the average
influx fibril. We assume that the energy to fracture the average fibril is simply
related to its cross-sectional area Af via: Gf = G0Af (70) where G0 is the
fracture energy of pure PE in the virgin state. We also assume here that the
influx fibrils are sufficiently interlocked that they

Incompatible semicrystalline 261 fracture rather than pull out. The area of an
average fibril is determined from the interstitial area between four contiguous
circles on a square lattice as: Af = r2(4 — π) (71) The number of fibrils per unit
area is determined from: Nf = l/(4r2) (72) Substituting for Nf and Af in the
above relations, we obtain the fracture energy due to influxes as: G = 0.22 G0
(73) which is independent of the spherulite radius. To compare with our fracture
experiments, we have the strength of the interface a&G1/2 such that: σ = 0.46 σ0
(74) where σ0 xG^2 is the tensile strength of the pure PE material. For HDPE,
σ0 values are in the range 22-31 MPa, giving interface strength of order σ= 10-
14 MPa. The maximum values reported in Figure 8.15 for the PP/HDPE
interface are consistent with this simple analysis. If the spherulite centers are
placed on a hexagonal lattice, then σ = 0.3 σ0. Heterogeneous nuclei can migrate
across the interface from the PP side to the PE side [34-36] during the wetting
step at 195 °C for 30 min. It is possible that when the PP crystallizes first, the PP
side could be ‘open’ in regions without solidifying for a certain time due to the
depletion of nuclei. This would facilitate the formation of the PE influxes when
the laminate crystallizes at 136 °C. When the crystallization time is around 0.5 h
(Figure 8.14), the PP spherulites initiated near the interface are still too small to
reach the interface. A planar appearance with a few sporadic disk-like
spherulites was seen by SEM in the fractured interface after further cooling for
spherulites was seen by SEM in the fractured interface after further cooling for
solidification. With a higher magnification of 1500 x, a number of very short and
fine fibers (3-5 pm) were revealed covering the entire surface. The limited
interdiffusion at the incompatible polymer melt interface resulting in the
entanglements between the two polymers is thought to be one of the mechanisms
in the development of interface strength and facilitating the influxes of fluid. In
contrast with the profuse fine scale fibers appearing everywhere on the fracture
surface of the short-time crystallized laminate, longer and ticker PE fibrillar
remnants are clearly seen in the interstices of the PP spherulites for the long-time
crystallized laminate. However, only a few fine fibers can be seen on the ‘shell’
of PP spherulites (Figure 8.14). The withdrawal or segregation of the PE chains
from the PP side during the slow crystalliza¬ tion process may have contributed
to the lack of fine fibers on the surface. This is consistent with rapid
crystallization causing entrapment of PE while slow crystallization results in
further phase separation from the interdiffused layer in the melt.

262 Welding and fracture of polypropylene interfaces The formation of the


influxes to depths of several micrometers should be considered when evaluating
concentration profiles in terms of diffusion mechanisms at crystallizing
interfaces. Obviously, the contribution of in¬ fluxes could give misleading
results when superimposed on diffusion con¬ trolled concentration profiles.
Many experiments of this type are reported in the literature without
consideration of the effective breakup of the interface. Summarizing the results
for isothermally crystallized PP/PE interfaces, the origin of the interface strength
may come from two sources. One is the entanglement of the PP and PE chains in
the melt to a depth of about 5 nm which results in a fine fibrillated fracture
morphology (Figure 8.16). The other is the mechanical interlocking caused by
the PE ‘flow’ resulting in influxes, which should be the most effective for the
long-time crystallization case. Figure 8.16 An optical micrograph of the
PP/LLDPE fractured interface obtained by rapid cooling from 195 °C in the hot
press. Note the thin fibrils on the fracture surfaces. Following impingement, the
size of the PP spherulites no longer increases with time and the interface strength
should level off. In our experiments, it seems that the interface strength tends to
level off at around 1.5 h at 136 °C. The maximum strength at this point is about
1/3 to 1/2 of the tensile strength of the virgin PE and considerably stronger than
the interface formed without the development of influxes. The force transferred
to the PE melt during the crystallization of PP is dependent on the location of the
PE fluid with respect to the occluded PP melt and the growing PP spherulites, as
shown in Figure 8.17. As the PP molecules crystallize, a driving force towards
the spherulites is created from the local volume contraction and is transferred to
the spherulites is created from the local volume contraction and is transferred to
the PE through melt

Incompatible semicrystalline 263 Figure 8.17 Schematic diagram of the effect of


the nucleation conditions on the force exerted on PE molecules for (a) nucleation
occurring at the interface and (b) nucleation occurring away from the interface.
entanglements. If the PP spherulites crystallize on the interface plane, the
direction of the resultant force exerted on the PE is largely parallel to the
interface (Figure 8.17a). Therefore, the PE molecules have difficulty with ‘flow’
into the PP matrix. The quenched and short-time crystallized laminate gives
many spherulitic nuclei appearing right on the border , which either lets the
interface ‘close’ to influxes, or minimizes the pulling force on PE as mentioned
above. The other situation (Figure 8.17b) is similar to the slow isothermal
crystallization case. The PP melt is surrounded by the PP spherulites nucleated
away from the interface and the PE influxes can be formed. With regard to
varying the kind of PE (HDPE or LLDPE) in these isothermal crystallization
experiments, no significant change in the structure of the interface was observed.
It can be expected that during isothermal crystallization, polyethylenes with
different viscosity (different melt index) could develop similar interface
structure provided they have sufficient time to flow. The butt tensile strength of
the interface obtained at 141 °C for 1 h is about 5.5 MPa and is lower than the
value obtained at 136 °C (10 MPa).

264 Welding and fracture of polypropylene interfaces Considering the


temperature dependence of the crystallization rate, the higher the crystallization
temperature, the smaller the spherulite size at constant time, and in turn the
smaller will be the dimension of PE influxes formed. Non-isothermally
crystallized PP/HDPE The structure and strength of quench cooled PP/PE
interfaces were studied. This process more closely resembles the thermal history
experienced by PP/PE blends during commercial melt processing. Although the
melting temperature Tm of PP is different from those of HDPE and LLDPE (160
°C, 130 °C, 124 °C for PP, HDPE, LLDPE, respectively) the crystallization
temperatures Tc, are very close and overlap. The Tc at a cooling rate below
10°C/min are 119°C, 113°C and 110°C for HDPE, PP and LLDPE, respectively.
The crystallization behavior of the laminates at a constant cooling rate was
studied using sections of the laminates on the Mettler 52 hot stage with a
polarized microscope. The results are described below. After a section of the
PP/HDPE laminate was held at 190 °C for 10 min and then cooled at 10 °C/min,
the nuclei of the PP occurred randomly at 126 °C in the entire area of the PP
side. The nuclei of HDPE were invisible at this point because of their small size.
As the temperature dropped to 121 °C, the nuclei of HDPE uniformly spread
As the temperature dropped to 121 °C, the nuclei of HDPE uniformly spread
rapidly over the HDPE side (Figure 8.18). In a short time, the crystallization of
the HDPE was completed, forming a solid boundary with PP which was not yet
fully crystallized. As soon as the spherulites in the HDPE side appeared, a
number of new PP nuclei occurred along the track of the interface; the latter can
be seen clearly (Figure 8.18) by remelting the section to 150 °C after the
crystallization had occurred at 121 °C. The nucleation of the new PP spherulites
near the interface may occur by one of the following possible means: (i) the
crystallized boundary of the HDPE behaves as a solid surface which acts to
nucleate the molten PP; epitaxial growth is possible; (ii) stresses developing at
the interface may come from the difference in thermal expansion coefficients,
6.8 x 10“5 °C_1 for PP and 3.3 x 10“4 °C-1 for PE, and volume contraction due
to the first order transition with crystallization. In the case where the HDPE side
crystallized first, the PP influx was still difficult to form since the HDPE
spherulites were usually very small (Figure 8.18). Also, new PP spherulites
which formed near the interface may have made PP influx formation too
difficult. Non-isothermally crystallized PP/LLDPE At a cooling rate of 10
°C/min the nuclei also appeared in the PP side first and fewer nuclei appeared in
the boundary. When the PP spherulite had

Incompatible semicrystalline re 8.18 Crystallization of a PP/HDPE interface, (a)


At a cooling rate 7min the HDPE solidified first (top right), shown at 121 °C;
some spherul shown on the PP side (bottom left) near the interface, (b) The PP
mi< rulitic boundary near the interface during remelting from 121 °C at 150 °C.
ad over the entire PP area, including the boundary region of the ] the temperature
of the section had reached 111 °C, the LLDPE had i crystallized (Figure 8.19).
The PP side of the interface crystallized fi irently without inducing nuclei in the
LLDPE side as was observed

i Welding and fracture of polypropylene interfaces ire 8.19 The PP/LLDPE


interface formed under a cooling rate of 10 °C/n 'ractured PP/LLDPE interface;
(b) heating of (a) to 150 °C, followed by cool 0 °C/min. The observation was
made at 111 °C while LLDPE had still i tallized. The black areas on the PP side
represent the PE influxes. PP/HDPE case. Then the LLDPE begins
crystallization, which lar in some ways to the process of the isothermal
crystallization of t HDPE laminate. The LLDPE influxes could also be formed,
but not 1 an extent as in the isothermal case. Since the temperature dropped

Incompatible semicrystalline pidly, the PP spherulites nucleated rapidly and


could not form the pro undary for accepting the LLDPE influxes. Several
could not form the pro undary for accepting the LLDPE influxes. Several
interesting phenomena were observed in the SEM micrograph 7PE interfaces.
With a long crystallization time we can clearly see gure 8.20 concentric rings on
the shell of PP spherulites. The rings fc ure 8.20 SEM micrographs of PP
spherulites showing rings and contour li the adjoining spherulites. The
spherulites were formed at a PP/HDPE interf ler long-time isothermal
crystallization conditions at 136 °C.

268 Welding and fracture of polypropylene interfaces continuous contour lines


through several adjoining spherulites. Padden and Keith [46] once showed the
same phenomenon using polarized optical microscopy and attributed the regular
banding pattern to rhythmic fluc¬ tuations in the mode of crystallization in
parallel with the thermostatic control of the heater, ± 2 °C. The rings are caused
by the lamellae twist and the temperature fluctuation during spherulite growth.
Four types of spherulites are formed during crystallization of polypropylene
[46,47]. They grow at different temperatures with different growth rates. The
temperature range for formation of these spherulites is approximately 128-138
°C. The temperature fluctuations of the laminate in the Carver press fell into this
range. The lines in Figure 8.20 therefore mirror the on-off cycle of the
thermostat on the hot press. Given that the hot press had a ±3°C fluctuation in a
6 min cycle, the black rings could be used to determine precisely local spherulite
growth rates by SEM analysis of solidified PP. Another interesting feature
obtained from the SEM micrographs is that from Figure 8.20 we may say that
the spherulites with clear black and white radial stripes on their tops are type III
spherulites. The type III spherulite has a concave boundary toward the type I
spherulite, which is consistent with the literature indicated by Padden and Keith
[46]. Since PP has so many structures in different levels the interface strength
could be affected in a complex manner by each of these structures. In addition to
the influxes formed in the interstices of spherulites, intraspherulitic entrapment
of melt is also possible. On several occasions, we noted that the PP spherulites
on the fracture surface would be ‘decorated’ with plastically deformed PE rem¬
nants. Thus, the polymer melt becomes entrapped in the growing lamellar
crystallites and provides another mechanism of mechanical interlocking of
incompatible crystalline interfaces. This mechanism deserves further study. 8.7
COMMENTS ON INCOMPATIBLE SEMICRYSTALLINE INTERFACES
The interface strength of laminated PP and PE is dependent on factors such as
the nature of the PP and PE, the crystallization temperature Tc, the
crystallization time, the nucleation control and the solidification conditions. The
interface strength is related to the structure, which is affected by the
crystallization conditions of the two contacting polymers. In conclusion, for the
isothermal crystallization case at 136 °C, the volume contraction asso¬ ciated
isothermal crystallization case at 136 °C, the volume contraction asso¬ ciated
with the nucleation of spherulites located away from the interface makes it
possible for the PE influxes to be formed between the PP spherulites in the
interface region. With increasing crystallization time, the size of the influxes
increases and some pear-like influxes appear, resulting in an increase in the
interface strength. In our experiments, the crystallization temperature and the
solidifi¬ cation conditions were effective at changing the size of the influxes.

Appendix 269 At slow crystallization conditions, the higher crystallization


temperature results in fewer and larger spherulites nucleating at random near the
interface. Influxes form slowly and eventually grow to promote mechanical
interlocking. With fast crystallization conditions, a large number of nuclei form
near the interface and tend to block off the influxes from the melt which is also
attempting to crystallize. Thus, rapid crystallization of PP/PE interfaces typically
results in a very weak interface. For the non-isothermal crystallization PP/HDPE
case at a cooling rate of 10°C/min, the HDPE side crystallizes first, then induced
crystals appear in the PP side near the interface. Both events are detrimental to
the formation of influxes, and the interface strength is weak. For the non-
isothermal PP/LLDPE crystallization case at 10°C/min cooling rate, since the Tc
of LLDPE is lower than that of PP, the interface morphology looks like that of
the isothermal crystallization case. However, the PP spherulites are smaller
owing to the high supercooling and the PE influxes are small. 8.8
ACKNOWLEDGEMENTS The author is grateful to NFS, AFOSR and FAMI
for recent financial support of this work. Particular appreciation is expressed to
Baoling Yuan for her assistance in the analysis of PP/PE interfaces. 8.9
APPENDIX: SYMBOLS AND ABBREVIATIONS A Helmholtz free energy Af
Fibril cross-sectional area AFM Atomic force microscopy C(x, t) Concentration
depth profile Cco Characteristic ratio D Self-diffusion coefficient da/dN Fatigue
crack propagation rate Dro Rouse self-diffusion coefficient DSC Differential
scanning calorimetry doQ Region of mixing F* Spreading coefficient G
Mechanical energy Gu Critical strain energy release rate Gf Energy to fracture
average influx fibril Go Fracture energy of pure PE HDPE High density
polyethylene IR Infrared

Welding and fracture of polypropylene Number of bonds per monomer


Boltzmann’s constant Critical stress intensity factor Interpenetration chain length
Chain contour length Length of entanglement Low density polyethylene Linear
elastic fracture mechanics Linear low density polyethylene Molecular weight
Critical entanglement molecular weight Avogadro’s number Number of influxes
per unit area Number of particles per unit volume Pressure Polyamide
per unit area Number of particles per unit volume Pressure Polyamide
Polybutadiene Number of bridges per chain Polycarbonate Polyethylene
Polyethylene terephthalate Polyimide Polymethylmethacrylate Polypropylene
Polystyrene Polystyreneacrylonitrile Polyvinyl chloride End-to-end distance of
entanglement Radius of gyration Scanning electron microscopy Secondary ion
mass spectroscopy Specular neutron reflection Time Temperature Crystallization
temperature Transmission electron microscopy Melting temperature Weld time
Stored elastic strain energy Upper critical solution temperature Ultrahigh
modulus polyethylene Strain energy density Critical strain energy density
Welding fraction

References 271 (X}2 Center of mass motion X(t) Monomer interdiffusion


distance XPS X-ray photoelectron spectroscopy z Number of monomers per
oaxis length Γ Surface tension δ Crack opening displacement <5C Critical crack
opening displacement 6 Fracture strain η0 Zero shear viscosity p Density σ
Fracture stress (Tc Critical fracture stress σ* Average stress creating
deformation zone at the crack tip τ Relaxation time Tf Failure time of
entanglement network re Relaxation time of entanglement segment χ Flory-
Huggins interaction parameter Xc Critical Flory-Huggins interaction parameter
8.10 REFERENCES 1. Wool, R. P. and O’Connor, K. M., (1981) Journal of
Applied Physics, 52 (10), 5953-63. 2. Kim, Y. H. and Wool, R. P (1983)
Macromolecules, 16, 1115-20. * 3. Wool, R. P. (1984) Rubber Chemistry and
Technology, 57 (2), 307-18. 4. Wool, R. P. (1994) Structure and Strength of
Polymer Interfaces, Hanser Press, New York. 5. Rouse, P. E. (1953) Journal of
Chemical Physics, 21, 1272. 6. de Gennes, P.-G. (1971) Journal of Chemical
Physics, 55, 572. 7. de Gennes, P.-G. (1980) Compte Rendu Hebdomadaire des
Seances de VAcademie des Sciences, Ser. B, 291, 219. 8. de Gennes, P.-G. and
Leger, L. (1982) Annual Review of Physical Chemistry, 33,49. 9. Doi, M. and
Edwards, S. F. (1978) Faraday Transactions, 2, 1789. 10. Kausch, Η. H. (1987)
Polymer Fracture, 2nd edn, Springer-Verlag, Heidelberg, Chapter 10. 11. Tirell,
M., Adolf, D. and Prager, S. (1986) IMA Volumes in Mathematics and its
Application, Vol. 5, Springer-Verlag, Heidelberg. 12. Adolf, D., Tirrell, M. and
Prager, S. (1985) Journal of Polymer Science, Polymer Physics Edition, 23, 413.
13. Prager, S. and Tirrell, M. (1981) Journal of Chemical Physics, 75 (10), 5194.
14. de Gennes, P.-G. (1981) Compte Rendu Hebdomadaire des Seances de
VAcademie des Sciences, Ser. B, 292 (2) 1505. 15. Voyutskii, S. (1963)
Autoadhesion and Adhesion of Polymers, Polymer Reviews Vol. 4, (eds H. F.
Mark and E. H. Immergut), Wiley Interscience, New York. 16. Helfand, E.
(1983) in Polymer Compatibility and Incompatibility, (ed. M. Sole), MMI Press,
Midland, MI. 17. Weber, T. A. and Helfand, E. (1976) Macromolecules, 9, 311.
Midland, MI. 17. Weber, T. A. and Helfand, E. (1976) Macromolecules, 9, 311.
18. Helfand, E. and Tagami, Y. (1972) Journal of Chemical Physics, 56, 3592.

272 Welding and fracture of polypropylene interfaces 19. Wu, S., Chuang, H.
and Han, C. D. (1986) Journal of Polymer Science, Polymer Physics Edition, 24,
143. 20. Yuan, B.-L. and Wool, R. P. (1990) Polymer Engineering and Science,
30, 1454. 21. Wool, R. P., Willett, J. L, McGarel, O. J. and Yuan, B.-L. (1987)
ACS Polymer Preprints, 28 (2), 38. 22. Wool, R. P., Yuan, B.-L. and McGarel,
O. J. (1989) Polymer Engineering and Science, 29 (19), 1340. 23. Foster, K. L.
and Wool, R. P. (1991) Macromolecules, 24, 1397. 24. Willett, J. L. and Wool,
R. P. (1993) Macromolecules, 26, 5336. 25. Paul, D. R. (1978) Polymer Blends
Vol. 1, Academic Press, New York, p. 25. 26. Flory, P. J. (1953) Principles of
Polymer Chemistry, Cornell University Press, Ithaca, New York. 27. Graessley,
W. W. (1982) Advances in Polymer Science, 16 (1), 47-67. 28. Aharoni, S. M.
(1983) Macromolecules, 16, 1722. 29. Kavassalis, T. A. and Noolandi, J. (1987)
Physical Review Letters, 59, 2674 and (1989) Macromolecules, 22, 2709. 30.
Tadokoro, H. (1979) Structure of Crystalline Polymers, Wiley, New York. 31.
Wu, S. (1990) Polymer Engineering and Science, 30 (13), 753. 32. Donald, A.
M. and Kramer, E. J. (1982) Journal of Polymer Science, Polymer Physics
Edition, 20, 899-909. 33. Yang, A. C., Kunz, M. S. and Logan, J. A. (1992) ACS
Polymer, Materials Science and Engineering Preprints, 67, 387. 34. Galeski, A.
and Bartczak, Z. (1986) Polymer, 27, 544. 35. Galeski, A. and Bartczak, Z.
(1984) Polymer, 25, 1323. 36. Galeski, A. and Piorkowska, E. (1983) Journal of
Polymer Science, Polymer Physics Edition, 21, 1313. 37. Kammar, H. W. (1979)
Zeitschrift fiir physikalische Chemie, 258 (6), 1149. 38. Shilov, V. V., Tsukruk,
V. V. and Lipatov, Y. S. (1984) Polymer Science USSR, 26 (7), 1503. 39. Rault,
J. (1986) CRC Critical Reviews in Solid State and Material Science, 13 (1), 57.
40. Zhang H. and Wool, R. P. Macromolecules, submitted. 41. Wu, S. (1982)
Polymer Interface and Adhesion, Marcel Dekker, New York, p. 98. 42. Gray, D.
G. (1974) Journal of Polymer Science, Polymer Letters Edition, 12, 509. 43.
Zupko, Η. M. (1974) Journal of'Applied Polymer Science, 18, 2195. 44. Hsu, C.
C., Geil, P. H., Miyaji, H. and Asai, K. (1986) Journal of Polymer Science,
Polymer Physics Edition, 24, 2379. 45. Natta, G. and Corradini, P. (1960) Nuovo
Cimento (Suppl), 15, 40. 46. Padden, Jr, F. J. and Keith, H. D. (1959) Journal of
Applied Physics, 30 (10), 1479. 47. Geil, P. H. (1973) PolyWfkSingle Crystal,
Robert E. Krieger, New York, p. 266.

9 Self-reinforcement of polypropylene J. Song, M. Prox, A. Weber and G. W.


Ehrenstein 9.1 INTRODUCTION Polypropylene (PP) is widely used because of
its low density, low cost, ease of processing and ability to be modified to meet a
variety of performance characteristics. Nevertheless, PP without fibre
variety of performance characteristics. Nevertheless, PP without fibre
reinforcement has difficulty penetrating the most significant areas of engineering
plastics because of its low strength and modulus. Fibre reinforced PP is
experiencing strong growth in this area, but with a corresponding sacrifice in
density, cost and fabricability. PP consists of carbon-carbon repeating units
along the backbone chain. Although this chemical bond is one of the strongest in
nature, that strength is not realized in polymer parts because the molecules exist
primarily as entangled random coils. Crystallization improves the strength and
modulus of polymers by reducing the degree of molecular randomiz¬ ation. The
basic crystal structure of a semicrystalline polymer consists of chain-folded
lamellae radiating outward from a nucleation site to form a spherulite. Orienting
semicrystalline polymers further improves the strength and modulus in the
orientation direction. Orientation is used to produce a high degree of crystal
alignment in high strength polyamide, polyethylene- terephthalate, polyethylene
and polypropylene textile fibres. The crystal structure still contains regions
between the chain folds that cannot carry a load, reducing the overall strength
and modulus of the fibre. When the crystal segments are lengthened there are
fewer chain fold defects, resulting in a stronger and stiffer fibre. Polypropylene:
Structure, blends and composites. Edited by J. Karger-Kocsis. Published in 1995
by Chapman & Hall, London. ISBN 0 412 58430 1

274 Self-reinforcement of polypropylene Because its backbone is the very strong


carbon-carbon bond, PP has a theoretical ultimate strength of 16GPa and an
ultimate modulus of 50GPa [1]. The primary reason why these high mechanical
properties are not achieved is that thermodynamic considerations favour the
formation of the spherulitic crystal structure as the polymer is quenched from the
melt. Orientation of a semicrystalline polymer below its melting point produces
extended chain crystals having fewer chain folds and defects. Because of the
high ultimate strength of the carbon-carbon bond, a fully aligned polymer has a
very high strength and modulus. There are several solid-state processing
techniques capable of produc¬ ing the extended chain crystals necessary for the
production of very high modulus polymers: drawing, extrusion, rolling and high
pressure crystalliza¬ tion. Melt processing is the most common polymer
processing technique, be¬ cause it allows high deformation rates and is capable
of producing a wide variety of product shapes. PP parts with high strength and
modulus can be produced in melt processing at relatively low temperatures and
under high pressures. The high performance parts are produced in melt
processing not by orienting the crystals, but by inducing oriented crystallization
in the polymer. Our experiments were carried out in injection moulding and melt
extrusion. 9.2 SOLID-STATE PROCESSES Many investigations have been
extrusion. 9.2 SOLID-STATE PROCESSES Many investigations have been
made by solid-state processing techniques. The processing temperatures of solid-
state processes are below the melting point. Remarkable self-reinforcement
effects are observed. Short descriptions of some solid-state processes follow [2].
9.2.1 Simple drawing Drawing is the simplest way to orient a semicrystalline
polymer. The process consists of applying a tensile force to the polymer at a
temperature below its melting point. Drawing is usually accomplished by
conveying the polymer through a series of rolls in which the downstream set of
rolls rotates at a faster rate than the upstream set. The processing temperature
was 100 °C, and the speed of drawing was 0.5 m/min for a drawing ratio of 27:1.
The modulus and tensile strength achieved were, respectively, 22 GPa and 930
MPa. 9.2.2 Die drawing The die drawing process consists of pulling the polymer
through a shaped die without the application of any pressure. An important
feature of this

Solid-state processes 275 process is that the polymer does not maintain a large
contact area with the die during the deformation, therefore, it is more efficient in
converting the applied forces to molecular orientation. Draw ratios as high as
20:1 have been achieved at output rates of 0.5 m/min. Compared with
conventional fibre and film formation, this is not a high production rate.
However, the products produced were highly oriented shapes, such as rods, I-
beam sec¬ tions, tubes and other shaped sections. The modulus achieved was 20
GPa. 9.2.3 Zone drawing In simple drawing and die drawing orientation is
conducted under isother¬ mal conditions. In the zone drawing process the
orientation is confined to a small volume of locally heated polymer. There have
been modifications of this process in which the temperature immediately past the
deformation zone is maintained at a level appropriate for annealing the polymer.
This process modification is known as zone annealing, and heat is applied to the
oriented polymer instead of cooling in the temperature zone control section. The
process used for drawing polypropylene consists of heating the poly¬ mer to 137
°C in the heating zone, drawing it at this temperature, and cooling it at —10 °C.
Annealing is not usually required, although it has been practised by imposing a
post-drawing zone temperature of 130 °C. The maximum draw ratio achieved
with polypropylene was 21:1, with a maxi¬ mum output rate of 0.25 m/min. The
modulus and tensile strength achieved were 18 GPa and 600 MPa, respectively.
9.2.4 Drawing of gels The polymer is dissolved in a suitable solvent above its
crystalline melting point, followed by cooling below the melting point, where
the semicrystal¬ line gel is formed. Because the polymer solution is very diluent,
it reduces the molecular entanglement in the condensed state. The concentration
of the polymer in the solution controls the degree of molecular entanglement,
of the polymer in the solution controls the degree of molecular entanglement,
and it must be reduced as the molecular weight of the polymer increases. The gel
is essentially a solid material with a very high microvoid content. Because it has
few entanglements per molecule it can be drawn to high degrees. With a draw
ratio of 50:1, a modulus of 36 GPa and a strength of 1030 MPa were obtained.
The necessity of handling large quantities of solvent is the major limitation of
the gel drawing approach, even though the solvents are relatively nonvolatile,
and can be recycled. The solvent handling techniques must be capable of
maintaining the polymer concentration in the solution within a narrow critical
range to prevent premature precipitation at flow orifices, and property variations
in the drawn fibre.

276 Self-reinforcement of polypropylene 9.2.5 Ram extrusion Extruding


semicrystalline polymer billets through a die at temperatures below the melting
point has also been used to produce highly oriented structures. The molecular
rearrangement that occurs in solid-state extrusion is similar to that obtained in
the drawing process; namely, elongational deformation of the polymer. The
reduction in cross-sectional area from the billet to extrudate is called the
extrusion ratio and is analogous to the draw ratio; Ram extrusion has been used
to produce fibres, films, tapes, sheets, tubes and a number of other shapes. The
process is very slow [3]. Attempts to extrude at higher rates results in process
instabilities. These instabilities consist of an extrudate that is highly distorted,
and emerges from the die in an erratic, ‘stick-slip’ fashion. 9.2.6 Hydrostatic
extrusion Slight improvements over the ram extrusion process can be obtained
from the hydrostatic extrusion process. The use of a hydrostatic fluid to provide
lubrication in the deformation zone reduces the pressure required to extrude the
polymer. Any reduction in the pressure on the polymer stabilizes the process.
Hydrostatic extrusion is still very slow, however. The modulus achieved was
16.7 GPa. 9.2.7 Rolling Rolling has been used primarily to produce highly
oriented films and tapes. The important process parameters are the roll and sheet
temperature, degree of deformation and molecular mass of the polymer. In the
rolling process, the primary measure of deformation is the amount of thickness
reduction achieved between the feedstock and product. This quantity, the
thickness reduction, is analogous to draw ratio and extrusion ratio. A modulus of
5 GPa and a tensile strength of 300 MPa at a production rate of 25 m/min has
been reported. The transverse properties of rolled films were similar to those of
the unoriented sheet, except for a strong tendency of the oriented sheet to
fibrillate, or split, along the orientation direction. 9.3 MELT PROCESSES The
precondition for the oriented crystallization is a high degree of orienta¬ tion of
the molecules in the melt. McHugh has calculated the amount of extension as a
the molecules in the melt. McHugh has calculated the amount of extension as a
function of deformation rate and molecular mass in polyethy¬ lene solutions [3].
His results are presented in Figure 9.1, which illustrates the enormous difference
between shear and elongational flow. Extensional forces are far more conducive
to chain extension and hence to nucleation of

Melt processes 277 Figure 9.1 Percentage extension of molecular chain in


extension and shear flow. fibrils than shear forces. Therefore, a convergent die
geometry and a high flow velocity are used to increase the strain rate [4], as
shown in Figure 9.2. With increasing pressure, the crystallization rate vk
increases. According to van Krevelen [5], i;k«exp( — 1/ΔΓ), where ΔT=Tm—
Tx. In order to promote the growth of shish-kebab layers, it is necessary to
obtain the greatest possible difference between the crystallization temperature
and the melting point. The crystallization temperature Tx during the process can
hardly be influenced, the crystallite melting point Tm is proportional to the
pressure, as shown in Figure 9.3. Figure 9.2 Schematic diagram of a die to
produce a longitudinal flow gradient.

278 Self-reinforcement of polypropylene Figure 9.3 Crystalline melting point as


a function of pressure for PP homopolymer (1 bar = 0.1 MPa). Abramov et al [1]
increased the strength of PP by the KWART tech¬ nology, which has not been
described in more detail. Strength values of 60 MPa were obtained using this
method. The effects of self-reinforcement on mechanical and thermal properties
of crystalline structures are analysed using mechanical tests, thermal analysis
and microscopical tests. 9.3.1 Experimental Process description of injection
moulding Studies were conducted on the PP homopolymers listed in Table 9.1.
PP grades with a low melt flow index (MFI) (PP4, PP5) and a high molecular
mass are used in extrusion and blow moulding. PP6 exhibits high molecular
mass and high isotacticity. The rest of the products are injection-moulding
grades, used for engineering components (PP1) and high wall thicknesses (PP2,
PP3). The weight-average molecular mass Mw, polydispersity Mw/Mn, density
and yield stress values in Table 9.1 are data from the manufacturers. Tensile test
specimens having dimensions of 4 mm x 5 mm x 45 mm within the measuring
range were used for the studies (Figure 9.4). The specimens were moulded on an
conventional injection moulding machine at a maxi-

Melt processes 279 Table 9.1 PP grades used Material Mw (g/mol) MJMn MFI
190 °C/5 kg g/(l 0 min) Density P (g/cm3) Yield strength (N/mm2) Young's
modulus (N/mm2) PP1 240000 6.2 12.0 0.907 30 1600 PP2 355 000 11.0 4.0
0.905 30 1400 PP3 380000 7.0 4.0 0.906 30 1500 PP4 470000 6.4 0.5 0.904 31
1400 PP5 650000 9.5 0.6 0.902 31 1300 PP6 653 000 20.0 0.3 0.886 32 1300 45
1400 PP5 650000 9.5 0.6 0.902 31 1300 PP6 653 000 20.0 0.3 0.886 32 1300 45
4 thickness o 45 2 thickne! 3S I „ 50 _1 4 thickness Figure 9.4 Geometry of the
test specimen. mum injection pressure of 175 MPa and a clamping force of 500
kN and subsequently tested. The moulding process was monitored by measuring
the hydraulic pressure, injection pressure, injection rate, injection and follow-up
pressure times, melt temperature and mould wall temperature. Process
description of extrusion The experimental setup consists of a conventional
pressure regulated single screw extruder, a convergent extrusion die and two or
more calibration zones connected directly with the die. The calibration zones are

280 Self-reinforcement of polypropylene oil-temperatured and allow a


longitudinal temperature gradient of more than 40 K within a length of 100 mm.
For the high extrusion pressures up to 100 MPa a 30 mm diameter extruder with
a length to diameter (L/D) ratio of 30 is used. For the flat strip profiles a ribbon
haul-off is used to increase extrusion velocity. Haul-off forces up to 500 N were
measured. The high-strength round-profiles with a diameter of 6 mm were
produced continuously with an extrusion velocity range from 20 to 50 mm/min.
The extrusion die has an inlet diameter of 30 mm. The outlet diameter was, like
the calibration diameter, 6 mm. The angle of convergence was changeable. Best
results were achieved with an angle of 45°. With the 6 mm round- profile die,
pressures up to 100 MPa were possible [6,7]. Besides rotational-symmetric
profiles, flat strip profiles with a cross- section of 70 mm x 2 mm were extruded.
Figure 9.5 shows the simplified die geometry. The extrusion pressure could be
varied in a range from 30 to 70 MPa and extrusion velocity from 40 to 200
mm/min. The reference material for all profiles was PP4 (Table 9.1). For all
tensile tests, specimen of dimen¬ sions 45 mm x 5 mm x 2 mm were milled from
the flat strip profile (Figure 9.4). Mechanical tests Besides the determination of
tensile strength and modulus, tensile tests are also used to compare the toughness
of self-reinforced and normally pro¬ cessed polypropylene and to measure the
anisotropic mechanical properties in and perpendicular to the flow direction.
Tensile tests were conducted at different temperatures. The toughness is
determined by the area below the stress-strain curve up to the yield point, and for
brittle material behaviour, up to the tensile strength. Long-term tensile tests were
conducted at various temperatures to deter¬ mine the creep properties of self-
reinforced PP in comparison to conven¬ tionally processed PP. The load was
varied to obtain strain values not more than 5% after 1 hour. An instrumented
Charpy impact test was conducted acceleration zone simplified die geometric
Figure 9.5 Simplified flat strip die geometry.

Melt processes 281 at — 40 °C for characterizing material toughness at low


Melt processes 281 at — 40 °C for characterizing material toughness at low
temperatures. The size of the specimens was 4 mm x 5 mm x 50 mm (Figure
9.4). The heat deflection temperature was determined according to Vicat (DIN
53460) using test specimens with dimensions lOmmx 10 mm x 4 mm. At a
heating rate of 50 K/h the test bar was loaded with a mass of 5 kg. The
coefficient of thermal expansion was determined at a constant heating rate of 5
K/min using specimens with dimensions 10 mm x 5 mm x 4 mm and a thermal
mechanical analyser (TMA). The change in length of the specimen under a
quartz glass ram with an additional load of 2 g was measured at increasing
temperature. The contact area was 5 mm x 4 mm when measuring the coefficient
of expansion parallel with the flow direction and lOmmx 4 mm when measuring
normal to this direction. Microscopic tests A polarization microscope was used
to measure the retardation, which characterizes the orientation of materials. Thin
cut specimens with a thick¬ ness of 10 pm were used for the investigations. A
high degree of orientation of crystal leads to a high retardation. Dynamic
mechanical analyser The dynamic mechanical analyser (DMA) imposes a
sinusoidal stress on a sample in the torsion or bending mode and determines the
sample modulus and mechanical loss factor as a function of temperature and
frequency. Information about transition temperatures and thermal stability of
mechan¬ ical properties can be obtained easily using DMA. Structure Analysis
In order to characterize the resultant structure, the specimens were studied using
wide-angle X-ray scattering (WAXS) and differential scanning colorimetry
(DSC). The heating rate in DSC was 10 K/min. DSC measure¬ ments were
evaluated by dividing the peak curve at peak temperature into two parts. The
ratio of melting enthalpy of these two parts was calculated to compare low
temperature and high temperature resistant crystal parts. 9.3,2 Properties
Injection moulding The parameters for moulding the self-reinforced test
specimens were determined on the basis of the experiments so far made. As
previous studies show [7], the most important parameters for self-reinforcement

282 Self-reinforcement of polypropylene in injection moulding are a high


injection rate for orientation of the chains and flow-induced formation of nuclei,
and a high injection and follow-up pressure for pressure-induced crystallization.
The optimum values for the other parameters were found by variation of the melt
and mould temperatures and kept constant for the ensuing studies. The most
important processing parameters are listed in Table 9.2. The mechanical
properties of self-reinforced PP as functions of the melt temperatures are shown
in Figure 9.6. When plotting the maximum values for strength and modulus at
various mould temperatures, an optimum mould temperature of 60 °C is
obtained (Figure 9.7). Table 9.2 Processing parameters for the preparation of the
PP specimens Normally processed Self-reinforced Injection rate Injection
PP specimens Normally processed Self-reinforced Injection rate Injection
pressure Follow-up pressure Melt temperature Mould temperature 30 mm/s 180
mm/s 87.5 MPa 171.5 MPa 87.5 MPa 171.5 MPa 260 °C 160-190 °C 25 °C 25-
80°C 100 5 75 I I ' ^ O) c * 50 25 t*. l^· 1 » *-* strength X-X modulus r— Mol
Mat< 1 ild tempe erial: PP- 1 1 rature: 60 homopoh 1 >3C ^mer 1 E ,E z UJ * ‘o
jo 0) jjj D Ό O S 165 170 175 180 185 190 Melt temperature [°C] Figure 9.6
Strength and modulus as a function of melt temperature. The molecular mass
being the decisive criterion for the self-reinforcibility of polymers, as noted by
Maertin [8] for high density polyethylene (HDPE), it is of fundamental
importance for self-reinforcement to use a suitable material. The relationship
between the mean molecular mass of the poly¬ propylenes and their mechanical
properties is shown in Figure 9.8. The mechanical properties of material
processed under normal conditions show only slight dependence on molecular
mass. When processing material with

Melt processes 283 Mould temperature [°C] Figure 9.7 Strength and modulus as
a function of mould wall temperature. 100 Ίο Q_ 5 75 1—1 s b ? 50 0) w Φ 1 25
£ Λ a-a strength normal | | A"* strength self-reinforced t> o modulus normal | | 1
ro ω ^ D Modulus of elasticity [GPa] ■ ο·α moouius seiweiniorc \ X o-- „
""•■■O' Materi al: PP-ho 1 I mopolym I . J er 1 u ( ) 100 200 300 400 500 600
700 Molecular weight Mw [103 g/mol] Figure 9.8 Dependence of mechanical
properties on molecular mass. self-reinforcing injection moulding, however, a
maximum property improve¬ ment is found for PP4 with a mean molecular mass
Mw of 4.7 x 105 g/mol and a narrow distribution, Mw/Mn of 6.4. PP4 was
therefore used as standard material for the following studies. The modulus and
strength determined in the tensile test at room tempera¬ ture result in a more
than 2.5-fold increase for the self-reinforced material compared with the
normally moulded one. The self-reinforced specimens form only minor necking
but have a marked yield point shifted toward clearly higher elongations (Figure
9.9). In the studies determining the temperature behaviour of these mechanical
properties, a clearly better behaviour over a wide range of temperatures was
found for the self-reinforced material than for the normally processed PP4
(Figure 9.10). A high pronounced toughness is observed as well. Whereas at

284 Self-reinforcement of polypropylene 100 80 Ίο Q_ 5 60 1—1 b to jjj 40 00


20 0 0 10 20 30 40 50 strain € [%] Figure 9.9 Stress-strain diagram of self-
reinforced and normal PP. CL o L_J * o '5 to J2 a> "Ό O a temperature of 23 °C
normally processed PP exhibits a toughness of 700mJ/mm2, a 7-fold increase in
toughness is observed by means of the self-reinforcement (Figure 9.9). The
creep of the self-reinforced PP is much lower than that of the normal PP, as
creep of the self-reinforced PP is much lower than that of the normal PP, as
shown in Figure 9.11. In contrast to conventionally processed material,
mechanical properties of self-reinforced PP by injection moulding survive at
high temperatures for long times. Figure 9.12 shows creep behaviour of self-
reinforced and normally PP at 120 °C. The increased toughness at low
temperatures was confirmed by Charpy impact test at —40 °C. Both total energy
and energy to fracture are doubled. Table 9.3 shows the friction coefficients and
wear factor of the normal and

Melt processes 285 Time [h] Figure 9.11 Creep behaviour of normal and self-
reinforced PP at 23 °C. Time [h] Figure 9.12 Creep behaviour of normal and
self-reinforced PP at 120 °C. self-reinforced PP. Both the wear process and the
friction are reduced by self-reinforcement. The results confirm the theory that
improved mechanical properties lead to improved tribological behaviours [9,
10]. Extrusion Effect of process parameters The pressure for a normal extrusion
process is not more than 10 MPa, while the die temperature is more than 250 °C
and the extrusion rate is more than lOOOmm/min. For self-reinforcement to
occur, it is necessary to reduce the

Self-reinforcement of polypropylene Table 9.3 Effects of the self-reinforcement


on the tribological behaviours of PP in a pin-on-steel disk wear configuration
with conditions P = 4N/mm2, u = 0.5m/s and T= 23 °C Wear factor (10-
6mm3/Nm) Friction coefficient Rz = 0.4 pm Normal PP 100 0.41 Self-
reinforced PP 26 0.38 Rz = 2.8 pm Normal PP 210 0.46 Self-reinforced PP 100
0.32 die outlet temperature to approximately 160 °C so that the crystallization
process can start directly after the acceleration zone. As a result of solid friction
in the calibration zone, the pressure increases and the extrusion rate decreases.
Self-reinforcement is first observed at an extrusion pressure of approximately 30
MPa. With increasing pressure, the extrusion rate increases at a constant die
temperature and reaches its maximum value of 200 mm/min at an extrusion
pressure of 70 MPa. The dependence of the modulus of elasticity on the
extrusion velocity during the continuous, self-reinforcing extrusion process is
shown in Figure 9.13. With higher extrusion velocity the extensional flow
increases and the time in which the molecules can relax is reduced. Higher
extrusion pressure makes it possible to increase the extrusion velocity. For the
flat strip tool the extrusion pressure is limited at 70 MPa. Extrusion velocity
[mm/min] Figure 9.13 Modulus in extrusion and perpendicular direction of PP-
high-strength flat strip profiles (1 bar = 0.1 MPa). Modulus of elasticity [MPa]
286
287 Melt processes With the flat strip profile it is possible to measure the
modulus in the perpendicular direction too. In this direction the profiles show a
Young’s modulus of about 2300 MPa. The modulus increases with increasing
pressure and flow rate. The Young’s modulus could also be increased in the
perpendicular direction by a factor of two through self-reinforcement. In
conjunction with a large longitudinal flow gradient in the die, the high extrusion
pressure favours crystallization conditions. At the same time, rapid solidification
over the entire cross-section is achieved by means of supercooling of the melt in
the die. Mechanical properties and structure High-strength PP extrudates are
smooth, and exhibit, in addition to higher strength and stiffness, high surface
quality and dimensional stability. The interior is found to have a very
transparent, quite homogeneous core with a high degree of stiffness and low
degree of deformability. The highest tensile strength measured at 6 mm rod
profiles exceeded 60 MPa, while the Young’s modulus was about 3.3 GPa. The
flat strip profiles have a strength of up to 50 MPa and a Young’s modulus of
about 2.5 GPa. Figure 9.14 shows a typical stress-strain diagram for an extruded
self-reinforced specimen. In the extru¬ sion direction, both the tensile strength
and the strain at break are greater. 0 5 10 15 20 25 30 35 40 Strain e [%] Figure
9.14 Stress-strain diagram for self-reinforced PP specimen in and across
extrusion direction. With lower calibration temperature we can obtain a faster
crystallization. This leads to a fixing of a more highly oriented structure, which
has a higher tensile strength in both the extrusion direction and the perpendicular
direction (Figure 9.15). If the calibration zone temperature is too low,
crystallization starts in the convergent die. As a result the extrusion Stress σ
[MPa]

288 Self-reinforcement of polypropylene 50 V 40 0. S I—I £ 30 09 c £ to 20 10 -


a 1 1 1— PP-H self-r ·-· extrusic einforced at 5< >n direction DO bar V- -*7
perpendicular direction V 1 1 110 120 130 140 150 Calibration zone temperature
[CC] Figure 9.15 Tensile strength as function of calibration temperature. 160
resistance increases enormously, causing the process to stop. In contrast, normal
PP homopolymer extrudates have a tensile strength of 32 MPa and a Young’s
modulus of 1.2 GPa in both directions. 9.4 THERMAL AND
MICROSCOPICAL ANALYSIS 9.4.1 Extrusion In the outer layers of the edge
region a retardation up to 0.02 is measured. This correlates with higher
crystallization velocity in this region. In this way it is possible to freeze the
molecules in a more highly oriented form (Figure 9.16). The shear modulus and
the damping factor obtained from the DMA in torsion mode show that with
controlled self-reinforcement during extrusion the shear modulus at high
temperatures also increases (Figure 9.17) and therefore a higher dimension
temperatures also increases (Figure 9.17) and therefore a higher dimension
stability at high temperature is expected. 9.4.2 Injection moulding The high
dimension stability at high temperature and the resulting mechan¬ ical strength
at high temperatures indicate that the increase in strength is the result of the
formation of special morphological structures. Shear modulus and damping
obtained from DMA in bending mode show a significant thermal stability and
mechanical improvement of the high-strength material at all temperatures
(Figure 9.18). The coefficient of thermal expansion of normally processed PP is
subjec¬ ted to a slight continous increase in the direction of flow from room

Thermal and microscopial analysis 289 Distance from surface [mm] Figure 9.16
Retardation as function of distance from surface (1 bar = 0.1 MPa). 0.25 0.2 «'a
0.15 % Oi C ‘5. 0.1 § o 0.05 0 — 60 -40 -20 0 20 40 60 80 100 120 Temperature
[CC] Figure 9.17 Shear modulus and logarithmic decrement of damping as
functions of the temperature of PP extrudates. temperature up to the softening
temperature of 90 °C. Only from this point onwards does expansion grow
substantially. In contrast, self-reinforced PP not only has a substantially smaller
thermal expansion but this value also remains constant up to a higher softening
temperature (Figures 9.19 and 9.20). There is only an insignificant difference
between the two materials when the thermal expansion in the value parallel
coefficients, measured in and normal to the direction of flow, are considered.

290 Self-reinforcement of polypropylene Figure 9.18 Dynamic bending modulus


and damping of PP as a function of temperature. Figure 9.19 Temperature
dependence of the coefficient of thermal expansion in the flow direction.
Conventionally processed PP has a spherulitic structure with fine spherulites at
the edge region and bigger ones in the core region. Micro¬ graphs of high-
strength PP do not show any spherulitic structure. In the edge region of the self-
reinforced specimen there is a high retardation, as shown in Figure 9.21, because
of the high orientation. DSC studies (Figure 9.22) show a very narrow melting
peak for conven¬ tionally processed PP. This suggests a relatively uniform
crystallite size. The energy required for melting is 97.9 J/g. The self-reinforced
PP, however, Damping

Thermal and microscopial analysis 291 Temperature [°C] Figure 9.20


Temperature dependence of the coefficient of thermal expansion per¬ pendicular
to the flow direction. 0.02 c 0.015 < c o | o.oi o> oi 0.005 m—■ self-reii t><3
normall forced y processed ■ ΉΙ Material: 1 PP-H 1 1 00 1 0 0.4 0.8 1.2 1.6 2
Edge distance [mm] Figure 9.21 Retardation as a function of distance from
surface. shows a broader melt peak with a clear tendency toward a double peak.
surface. shows a broader melt peak with a clear tendency toward a double peak.
Here, a changed morphology, which partly results in a higher thermal stability
than in the case of normally processed PP, is found. The melting enthalpy of
100.8 J/g is also higher for the self-reinforced PP. Evaluation of melting
enthalpy in a low temperature and a high temperature stable part show an
increase of the melt enthalpy ratio of about 60% (Figure 9.23). This
demonstrates that the higher thermal stability of the specimen is well- founded in
a high-temperature resistant crystalline structure of the self- reinforced
specimen. The assumption of the highly oriented molecules in the self-reinforced
PP is confirmed by the result of the wide-angle X-ray deflection. The wide-angle

292 Self-reinforcement of polypropylene Edge distance s [mm] Figure 9.23 DSC


analysis of melting behaviour of PP. X-ray scattering (WAXS) micrographs
(Figure 9.24) reveal that the struc¬ ture formed in the conventionally processed
specimen has no preferred orientation whereas that in the self-reinforced
specimen shows a strong anisotropy. 9.5 CONCLUSION High processing
pressure and controlled flow conditions (velocity, melt and tool temperature) in
conjunction with a longitudinal flow gradient enable the production of high
performance PP with conventional melt processing machines. In comparison
with the normally processed PP, the self-rein- forced PP exhibits a high
molecular chain orientation largely parallel to the

Appendix ϊ 9.24 WAXS micrographs of (a) normally processed and (b) self-reinf
direction and improved mechanical behaviour, frictional propertiei resistance, as
well as improved thermal stability. ;ause of the low degree of molecular
randomization, a high trans >th is also obtained. A striking phenomenon is the
increase in tc particularly at low temperatures. lile high performance PP parts
without fibre reinforcement ca iced with injection moulding, high-strength
profiles in a varie s can be produced with continous extrusion. The technology
desc: s chapter presents new possibilities in PP processing.

294 Self-reinforcement of polypropylene 9.6 APPENDIX: SYMBOLS AND


ABBREVIATIONS DMA Dynamic mechanical analyser DSC Differential
scanning calorimetry HDPE High density polyethylene L/D ratio Ratio between
flow length and wall thickness MFI Melt flow index [g/10min)] according to
DIN 53735 Mw Weight-average molecular mass Mw/Mn Polydispersity PP
Polypropylene Tm Melting temperature TMA Thermal mechanical analyser Tx
Crystallization temperature uk Rate of crystallization WAXS Wide-angle X-ray
scattering 9.7 REFERENCES 1. Abramov, V. V., Kuznecov, V. V., Veselow, A.
V. et al (1984) Plaste und Kautschuk, 31, 220-3. 2. Bigg, D. M. (1988) Polymer
Engineering and Science, 28, 830-41. 3. McHugh, A. J. (1975) Journal of
Engineering and Science, 28, 830-41. 3. McHugh, A. J. (1975) Journal of
Applied Polymer Science, 19, 125. 4. Petermann, J. (1982) Eigenverstarkung
von Kunststoffen, Verbundwerkstoffe und Werkstojfverbunde in der
Kunststofftechnik, VDI-Verlag, Dusseldorf. 5. van Krevelen, D. W. (1978)
Chimia, 32, 279. 6. Pornnimit, B. (1990) Eigenverstarkung von Polyethylen in
der Extrusion, PhD Thesis, University of Kassel. 7. Pornnimit, B. and
Ehrenstein, G. W. (1991) Extrusion of self-reinforced poly¬ ethylene, in
Proceedings of the 7th Annual Meeting of the Polymer processing Society,
Hamilton, Ontario, April. 8. Maertin, Cl. E. (1988) Zur Eigenverstarkung von
Polyethylen im SpritzguB, PhD Thesis, University of Kassel. 9. Song, J.,
Ehrenstein, G. W. (1993) Friction and wear of self-reinforced thermoplastics, in
Advances in Composite Tribology (ed. K. Friedrich), Elsevier, Amsterdam, 19-
63. 10. Song, J., Maertin, C. and Ehrenstein, G. W. (1988) SPE ANTEC, 46,
587.

10 Processing-induced structure formation* H. JaneschitZrKriegl, E.


Fleischmann and W. Geymayer 10.1 INTRODUCTION As is well known, the
interior structure (texture) of extruded or injection moulded plastic articles is far
from being homogeneous or isotropic. This is true for amorphous polymers as
well as for semicrystalline ones. In amor¬ phous polymers the most common
way of investigating these shortcomings is the observation of birefringence
patterns, as found in prepared cross- sections in injection moulded items [1].
Also, stress and density distribution can be observed by more sophisticated
methods [2]. In crystalline polymers, which are considered in this chapter, one
can simply observe a spatial distribution of grain sizes (spherulites) or highly
oriented and birefringent surface layers with the aid of microscopic techniques
[3]. It is also well known that the interior structure of plastic parts can be
influenced by a proper choice of processing conditions (wall and melt
temperatures, extrusion or injection speeds, draw down ratio, packing pressure
etc.). It is also possible to influence those structures by additives (nucleation
agents, lubricants etc.) or by the choice of the proper polymer grade (tacticity,
degree of copolymerization, molar mass distribution). This whole package of
knowledge, however, is based on practical experience, by trial and error. The
same holds for the consequences of certain textures for the end use properties of
the products. * Dedicated to Professor Paul Urban, in gratitude for his decisive
early support of the first author. Polypropylene: Structure, blends and
composites. Edited by J. Karger-Kocsis. Published in 1995 by Chapman & Hall,
London. ISBN 0 412 58430 1

296 Processing-induced structure formation For three-dimensional parts internal


296 Processing-induced structure formation For three-dimensional parts internal
anisotropies are obviously undesir¬ able. In particular, they are characterized by
anisotropic thermal expansion, which leads to warping, or by the tendency to
crack formation in certain directions. For crystalline polymers a desirable feature
is the occurrence of fine grains (spherulites), because these tiny structures are
interconnected by a great number of tie-molecules, whereas well developed large
spherulites do not adhere very strongly to each other. Traditionally, three fields
of investigation have been developed to improve our knowledge of crystalline
polymers: 1. The characterization of textures. This can be achieved with the aid
of a great variety of techniques such as X-ray methods, electron microscopy and
visual light optics. 2. The correlation between texture and end use properties. For
this purpose the mechanical properties (including micro-mechanical properties
of sections) are investigated as well as shrinkage, reactivity to solvents, warping
and, last but not least, surface appearance. 3. Investigation of the texture
formation during processing. So far, mainly the systematic variation of process
conditions has been carried out. In the following we will refer to this as the
phenomenological approach. This approach is in line with the ideas prevailing in
research and development laboratories. Existing commercial products can be
im¬ proved in this way. Interestingly enough, it appears that few people have
dared to extend their investigations into the area of processing par¬ ameters that
no longer lead to acceptable products. However, very often the textures found
under these conditions more clearly reveal the underlying mechanisms. A more
satisfying approach to the mechanisms of texture formation is, of course, given
by fundamental research, where the interaction between crystallization kinetics
and relevant transport phenomena (heat transfer and flow) is investigated. In this
chapter, both research paths are described. So far, however, the gap between
obtainable results has been disappointingly wide. Conclusions, which can be
drawn from the variation of processing parameters, appear to be very restricted,
mainly because of the complexity of the processes. On the other hand,
fundamental research has always been restricted to experiments that can be
interpreted unambiguously. Some years ago, one of the present authors (H. J.-K.)
found out that the above mentioned interaction between crystallization kinetics
and relevant transport phenomena had not received sufficient attention. This, of
course, formed a welcome starting point for renewed efforts. With the aid of the
‘Fonds zur Forderung der wissenschaf- tlichen Forschung’ in Vienna he
succeeded in establishing the national working party S33 on the influence of
molecular structure and processing parameters on the properties of moulded
plastic parts [4]. Seven Austrian

Phenomenological approach 297 research laboratories cooperated for five years


Phenomenological approach 297 research laboratories cooperated for five years
in this working party. Subsequent projects were generously granted by the Fonds
zur Forderung der wissenschaftlichen Forschung. In this chapter, a review of the
results achieved by these projects is presented. Polypropylene was chosen as the
subject of investigation because of its suitable time scale in crystallization
experiments and also because it is a major product of Petrochemie Danubia, one
of the interested Austrian industries. The participants and projects are listed in
Table 10.1. Table 10.1 Projects and participants in working party S33 Project
Participants 01 Theory 02 Model experiments 03 Processing 04 Morphology 05
Texture and mechanical properties 06 Crystallization kinetics 07 Molecular
characterization Prof. W. Schneider, Dr J. Berger and Dr A. Koppl, Vienna
University of Technology Prof. H. Janeschitz-Kriegl, Dr G. Eder, Dr G. Krobath,
Dr C. H. Wu, Dr E. Ratajski, S. Liedauer, Dr H. Wippel, Linz University Prof.
W. Knappe, W. Friesenbichler, Ch. Kukla, Leoben University of Min¬ ing Prof.
W. Geymayer, Dr P. Zipper, Dr E. Ingolic, Dr A. Janosi, E. Wren- tschur, Graz
University and University of Technology Prof. J. Koppelmann, Dr E. Fleis-
chmann, Leoben University of Mining Dr H. Dragaun, Dr H. Muschitz, J.
Moitzi, A. Schmidt, Vienna University of Technology Prof. K. Lederer, Dr J.
Billiani, Dr I. Amtmann, Leoben University of Min¬ ing, Dr A. Schausberger,
Linz Univer¬ sity 10.2 PHENOMENOLOGICAL APPROACH: RESULTS OF
INJECTION MOULDING EXPERIMENT 10.2.1 Processing Technology Two
types of mould cavities were used: a strip mould of orthorhombic symmetry 230
mm x 70 mm x 2 mm with a line gate on one of the smallest side walls, and a
disk mould with central gate (diameter 280 mm, thickness 2 mm or 4 mm). The
injection moulding machines were of the reciprocating

298 Processing-induced structure formation screw type (Engel, Schwertberg,


Austria), equipped with microprocessors for control purposes. Large axial
velocities of the screw could be achieved with the aid of a hydraulic
accumulator. The forward screw speed could be measured with the aid of a
special device linked to a microprocessor. Several pressure gauges were
mounted flush into the large walls of the moulds at various distances from the
gates. Three types of experiments were carried out: 1. Conventional mould
filling with packing stage and sealing by normal freeze-off or mechanical means.
2. Volumetric mould filling. In this process the screw was stopped when the
mould was just filled. No packing was applied. 3. Mould filling with mechanical
sealing, where the pressure was raised to a peak value at the end of the filling in
order to compress the melt quickly before sealing. This compression was meant
to compensate for the shrinkage with cooling. In this process, any additional
flow of the melt during a prolonged packing stage was avoided. This process has
been called mould filling without separate packing stage [5], not to be confused
been called mould filling without separate packing stage [5], not to be confused
with process 2. To indicate the kind of filling process applied, the following
notation was used, (appendix 3 in [4]): Tm/Tw/us/Pn/c where TM is the
temperaure of the injected melt, Tw is the wall temperature of the mould, vs is
the axial speed of the screw during injection and PN is the pressure in the mould
under various conditions indicated separately under c. These conditions are:
holding time with mechanical sealing; th, thermal sealing (normal freeze-off); f,
continously decreasing holding pressure; vf, volumetric filling; and nf, no
separate packing stage (i$ = 0, PN = Ppeak)· In the cases of and th, PN is the
holding pressure. Temperatures are in Centigrade, screw speeds in mm/s,
pressures in MPa and times in seconds. 10.2.2 Polypropylene grades used
Sample specification Two commercial grades were investigated: Daplen KS 10
(lot no. 4095 D II) and Daplen PT 55 (lot no. 41207 TD). Both grades were
provided by Petrochemie Danubia in sufficient quantities for an extended series
of measurements. The first grade is a reactor product, the second is a ‘rheology
controlled’ grade manufactured in a degradation step with peroxide from a high
molar mass reactor product. Some data for these polymers are given in Table
10.2 [6]. The data of Mw (weight average molar mass) and Mn (number average
molar mass) were obtained from gel permeation chromatography (GPC)

Phenomenological approach 299 Table 10.2 Some data for the polypropylene
used (molar masses are in g/mol with no correction for peak broadening; see
chapter 2 in [4]) Polypropylene Mz M„ M„/M„ MFIa 230 °C/2.16 kg KS 10
1360 ±14% 322 ±3% 47 + 8% 6.8 8.0g/10min PT 55 522 ±4% 205 ±5% 59 ±8%
3.5 18.0 g/10 min a MFI, melt flow index in one of the laboratories cooperating
in the working party. As a reactor product KS 10 has a much broader molar mass
distribution than PT 55. Some remarks on the importance of the molar mass
distribution It turns out that, in reactor products, tiny fractions of very high
molar masses (no more than 0.5% in total) dominate the rheological properties
of the melts in the low shear rate range (zero shear viscosity, normal stresses,
shear and tensile compliances) [7]. This is responsible for the difference between
KS 10 and PT 55, which have nearly the same non-Newtonian shear viscosities
in the high shear rate range of capillary viscometry. (Values of the properties
mentioned above are much higher for KS 10). It appears also that the sensitivity
of reactor products to flow-induced crystallization is much higher than that of
rheology controlled grades. But this statement is qualitative. This is the reason
why, in this chapter, the rheology is also treated only qualitatively. The tiny
fraction of high molar mass was found when the molar mass distributions, as
obtained with the aid of GPC, were extrapolated in accordance with a log-
normal distribution into the range of high molecular weight fractions where
normal distribution into the range of high molecular weight fractions where
these fractions are too small for detection in the GPC apparatus. In Figure 10.1
the molar mass distributions of KS 10 and PT 55 are shown, as obtained by
GPC. The added fractions go up to about log M = 8. If the detectable fractions
only are used in a model calculation, one obtains quite unrealistic curves for the
storage and loss moduli (G', G") of these melts in the low frequency range. If
successively more and more extrapolated fractions are added, however, these
curves converge exactly into the experimental curves. For the storage modulus
G' this means an increase by a whole decade in the low frequency range [7]!
10.2.3 Means of investigation Sample preparation From the injection moulded
parts samples must be prepared for the various types of experiments described
below. The positions and directions of the

300 Processing-induced structure formation Figure 10.1 Differential molar mass


distributions of polypropylenes Daplen KS 10 and PT 55 and the standard
deviations ( ) calculated from 6 runs (M in g/mol), according to [6]. (Courtesy of
Marcel Dekker.) sections of the samples are described later; here, only the
methods of preparation will be summarized. It is obvious that the required shape
and size of the samples varies from technique to technique: rather thick samples
(up to 3 mm) are needed for X-ray experiments; the thinnest samples are used in
transmission electron microscopy (TEM). For extended samples (up to 30 mm)
and thicknesses down to 40 μιη an internal-hole-saw Leitz microtome was used
(Leitz Saegemikrotom 1600). Down to a thickness of 0.1 μιη a Reichart
ultramic¬ rotome (Ultracut E) was used. Samples of 3-6 μιη and of 40 μιη were
investigated using polarized light microscopy. A cryo-ultramicrotome was used
for cutting the samples that had to be etched and/or stained for electron
microscopy. Some remarks on the measuring techniques applied The normal
microscopic staining techniques for polyethylene cannot be applied to
polypropylene unrestrictedly. A particularly mild and useful technique for
polypropylene is staining by Ru04 vapour. A mapping technique was developed
in order to produce series of electron micrographs of an enlargement of about 80
000 down from the edge of the sample over considerable distances. Sometimes,
laser light scattering was applied to the obtained photographs in order to analyse
their patterns by Fraunhofer diffraction of areas of 3 μπι in diameter [8]. Several
different techniques were applied for X-ray measurements (CuKa radiation was
used throughout). A pin-hole camera was used to obtain

Phenomenological approach 301 small-angle and wide-angle X-ray scattering


(SAXS and WAXS) patterns on a photographic film; these patterns were needed
to establish the types of textures of the samples. However, the irradiated area on
to establish the types of textures of the samples. However, the irradiated area on
the samples was too big for scanning of details of the texture with this camera
(diameter 0.2 mm to 0.4 mm; the beam diameter cannot be reduced ad libitum
because the concomitant loss of X-ray flux has to be compensated for by a
prolonged exposure time). Fortunately, in cross-sections perpendicular to the
contact surface of the moulded part with the mould wall the texture does not
change very much along planes parallel to this surface, whereas in the direction
perpendicular to this surface drastic and abrupt changes may occur. Therefore, a
rather high spatial resolution is required only in the latter direction. This renders
possible compensation for the loss of flux which results from narrowing the
beam in the direction perpendicular to the surface by widening it along a line
parallel to the surface. Such a beam with a line-shaped cross-section of only 0.06
mm width and about 2 mm length can be created by the well- known Kratky
camera [9], a device developed for SAXS investigation at extremely small
scattering angles. A diagram of this collimation system and the positioning of
the beam on the sample is shown in Figure 10.2 (drawing not to scale) [10]. The
direction of the movement of the sample during scanning (in steps of 0.01 mm)
and the method of detecting the scattering under a fixed angle by means of a
proportional counter are indicated. It should be noted that sacrificing the high
axial symmetry of a pin-hole beam in favour of a high-flux line-shaped beam
certainly imposes some restrictions on the geometry and the information content
of the scattering experiment, but these restrictions do not seem too serious. This
holds especially for scanning the cross-section of the sample at selected
scattering angles in the wide-angle region (the chosen angles correspond to the
strongest reflections of polypropylene crystallites), although ‘scattering profiles’
can be obtained Detector Figure 10.2 Kratky collimation system [10] (Courtesy
of Verlag Lorenz.)

302 Processing-induced structure formation for any scattering angle. (Recently


the information content of this technique was significantly enhanced when the
proportional counter was replaced with a linear position-sensitive detector
(PSD), which enables the registration of scattering profiles at hundreds of
different angles at the same time.) Another quite unusual technique has been
developed for (manual) scann¬ ing of the layer structure of the moulded parts.
For this purpose, layers of 40 μηι thickness were cut from the parts in a direction
parallel to the contact surface with the mould wall. These layers were separately
investigated in a microtesting machine. In this way the elongation at break and
the tensile strength could be determined for each layer [11]. In addition, the
density and the degree of crystallinity by differential scanning calorimetry
(DSC) were determined for these layers. At the same time, tensile tests were
carried out on test bars of the thickness of the whole sample. The influence of
carried out on test bars of the thickness of the whole sample. The influence of
mild stretching of these bars on the SAXS profile was investigated. It appeared
that microvoids of certain shape and orientation were created in certain areas,
depending on the conditions of stretching [12]. The degree of (anisotropic)
shrinkage was also investigated. 10.2.4 Morphology Experiments with normal
injection moulding [12] As an example, results obtained from a disk of Daplen
KS 10 (diameter 280 mm, thickness 2 mm) are reproduced. Moulding conditions
were as follows: melt temperature TM = 230 °C, mould wall temperature Tw =
30 °C, injection time 1.0 s, packing pressure in the cavity 34 MPa, packing time
10 s. Some disks were also obtained by volumetric filling of the cavity. Figure
10.3 shows the positions of the specimens taken from the disks. The centres of
specimens 2, 3, AZ2 and AZ3 are 40 mm from the central gate. The centres of
the other specimens are 100 mm from the gate. The samples coded AZ did not
receive any mechanical treatment. The contours of the test bars, as shown in
Figure 10.3, indicate that these samples were loaded until whitening became
observable. In this case, samples of 15 mm x 3 mm x 2 mm (hatched areas) were
cut out from the test bars after loading. Layers for the micromechanical tests
were sliced from these test bars before loading. An idea of the layered structures
is obtained from Figures 10.4-10.6. Figure 10.4 shows the features of specimen 3
(40 mm from the gate, cut in the radial direction, with packing). The micrograph
at the bottom shows the layers as seen by polarizing microscope. The upper
graph shows the birefringence distribution and the elongation at break (as
obtained by the micromechanical test). The lower graph shows the SAXS
profiles at an angle of 1 mrad as obtained by slit collimation before and after the
application of

Phenomenological approach 303 Figure 10.3 Locations of sampling in the


injection moulded disk. AZ1-AZ4, Samples for measurements in the initial state;
1-4, samples for measurements after loading, where in a first step the indicated
testbars are cut from the disk, according to [12]. Distances are mm from the gate.
(Courtesy of Wiley.) a load on the sample. Layer structures of injection moulded
polypropylene parts have been reported elsewhere, but this seems the most
complete presentation so far. In Figure 10.5 results for a tangential cut are shown
(40 mm from the gate, with packing). In this cross-section the birefringence is
very weak. Nevertheless, the elongation at break indicates the transition between
highly oriented surface layer and core. With loading, however, there is only a
slight increase in the peak of the SAXS profile. In Figure 10.6 results for a disk
obtained by mere volumetric mould filling are shown. This figure is from a
radial cut 40 mm from the gate. The results shown here are not obtained from a
practicable process. The surface of the disk is not perfect and its dimensions
practicable process. The surface of the disk is not perfect and its dimensions
deviate from those of the cavity. However, from this figure one can conclude
that the highly oriented surface layer has been formed already during the filling
procedure (or at least is a consequence of this procedure). Because of the bumpy
surface of this disk no cuts could be made parallel to this surface. Therefore, no
micromechanical tests could be conducted. A comparison with Figure 10.4
shows that the packing stage has caused the birefringence to increase in the core.
For some important experiments, one has to resort to the pin-hole camera. This
is shown in Figures 10.7 and 10.8. In Figure 10.7 the small angle pattern is
shown for specimen AZ3 at 0.1 mm below the surface. The

304 Processing-induced structure formation 0 0,2 0,4 0,6 mm 1,0 Distance from
the surface Figure 10.4 Profiles of birefringence and elongation at break (top),
small angle X-ray scattering (centre) (—) before and ( ) after loading, and
polarized light micrograph (bottom) for sample 3 in Figure 10.3 [12]. (Courtesy
of Wiley.) Distance from the surface Figure 10.5 Profiles of birefringence and
elongation at break (top), small angle X-ray scattering (centre) (—) before and (
) after loading, and polarized light micrograph (bottom) for sample 2 in Figure
10.3 [12]. (Courtesy of Wiley.)

Phenomenological approach 305 Distance from the surface Figure 10.6 Profile
of birefringence (top), small angle X-ray scattering (centre) (—) before and ( )
after loading, and polarized light micrograph (bottom) for a sample in position 3
(figure 10.3) from a disk obtained by volumetric mould filling (courtesy of John
Wiley). Figure 10.7 Small angle X-ray pattern obtained using a pin-hole camera
for an (unloaded) sample from position AZ3 at a distance of 0.1 mm below the
sur¬ face. The previous flow direction is in the horizontal direction [12].
(Courtesy of Wiley.)

306 Processing-induced structure formation Figure 10.8 Wide angle X-ray


pattern obtained using a pin-hole camera for the sample in Fig. 10.7 at a distance
of 0.15 mm below the surface. The previously flow direction is in the horizontal
direction [12]. (Courtesy of Wiley.) main lobes are in the direction of previous
flow. Since the second order maxima are also noticeable, there must be a quite
pronounced periodicity in this direction, pointing to a long period of about 15
nm. The degree of orientation decreases in layers descending to the mid-plane.
With no packing stage, the range of very low orientation was much more
extended. In Figure 10.8 a wide angle diagram, obtained using the pin-hole
camera at 0.15 mm below the surface of specimen 3, is shown. The strongest
reflections of a- and /^-modifications of polypropylene are shown with their
reflections of a- and /^-modifications of polypropylene are shown with their
Miller indices. The degree of orientation is very high. The appearance of arcs of
the 1 1 0-reflection both on the equator and around the meridian indicates the
existence of crystallites of which the crystallographic c-axis points in the
direction of flow, as well as of crystallites with their a*-axis in this direction. In
specimens that had been loaded cracks could be observed under the microscope.
With radial loading these cracks, which were perpendicular to the direction of
stretching, were found exclusively in the spherulitic core. Despite the fact that
the highly oriented surface layers seemed undamaged under the microscope, the
enhanced SAXS in these layers points to the formation of anisometric
microvoids, the smallest extension of which lies in the radial direction. With
tangential loading cracks perpendicular to the stretch direction are also found in
the surface layers. In addition, some other properties of the highly oriented
surface layers and the directly adjacent inner layers should be mentioned. In
Figure 10.9 high resolution TEM micrographs are shown for locations in a radial
cross-section between 0 pm and 78 μτη from the surface. The very instructive
laser-light scattering of these micrographs is also included. It can be clearly

Phenomenological approach 307 Figure 10.9 TEM micrographs of a radial


ultramicrotome section taken perpen¬ dicular to the surface of a disk of KS 10
(229/32/55/614/10), showing the different orientation of supra-molecular
lamellae in their dependence on the distance from the surface and analysis of
these micrographs by laser light scattering (analysed regions 3 pm) [8]. seen that
near the surface the lamellae are nearly perpendicular to the mould wall. Because
in the corresponding birefringence the larger refractive index is in the direction
parallel to the surface, indicating the direction of the chain molecules, one can
conclude that these lamellae are formed by chain folded crystals. In lower layers,
a second orientation of lamellae occurs, so that at 78 pm the lamellae are
preferably parallel to the mould wall. However, the birefringence does not
change sign with the change of orientation of the lamellae. A reason for this may
be the larger field of observation in light microscopy. The local distribution of
the /^-modification of polypropylene in the mouldings was also explored (it is
well known that /?-spherulites are preferentially formed under certain conditions
[13]). Generally, jS-poly- propylene was found to be concentrated in the
transition zone between the surface layers and the core region. In mouldings of
KS 10 the amount of jS-phase was observed to be higher than in mouldings of
PT 55. With both types of polypropylene halo-rings are found by visual
inspection of moulded disks. For PT 55 such a ring was found rather close to the
gate. For a disk of KS 10, however, such a ring occurred at a greater distance
from the central gate. Moulding conditions (230/70/105/381/20) were applied in
both cases.
both cases.

308 Processing-induced structure formation Figure 10.10 shows a perspective in


three dimensions of the scattering profiles of the β-3 0 0 reflection and of the a-0
4 0 reflection as functions of the distance from the gate. The turbidity of the disk
as a whole as a function of the distance from the gate is also shown. It can be
clearly seen that the maximum of the turbidity (the centre of the halo) occurs in
layers close to the surface where the β-3 0 0 reflection is particularly strong.
Microscopic observations revealed that a great many /?-spherulites were
immersed in these layers. It can therefore be concluded that the halo is caused by
the light scattering from these /?-spherulites. (As is well known, /?-spherulites
appear bright between crossed polars, whereas α-spherulites - because of
epitactic growth - show only a weak birefringence [14]. Also, the average
refractive indices seem to be different!) Haloes due to jS-spherulites were found
also in rectangular plates and in the side walls of boxes of KS 10. Another
interesting observation was made in tangential cross-sections (cf. samples 1 and
2 in Figure 10.3). When a scattering profile for the a-040 reflection was taken on
such a cross-section, the outcome depended on whether the beam was directed
outward or inward. In both cases one obtained an asymmetric profile with
respect to the mid-plane of the sample. The two asymmetric profiles turned out
to be mirror images with respect to the mid-plane of the sample. This is shown in
Figure 10.11. On the other hand, profiles for the a-0 4 0 reflection taken on
radial cross-sections with Figure 10.10 Perspective in three dimensions of the
scattering profiles of the β-3 0 0 and a-0 4 0 reflections as functions of the
distance from the gate for a KS 10 disk (230/70/105/381/20), including the radial
turbidity distribution of the disk as a whole (the optical density) [10]. (Courtesy
of Verlag Lorenz.)

Phenomenological approach 309 Distance from the surface (mm) Figure 10.11
Scattering profiles of a-0 4 0 in a tangential section of a disk of KS 10, 89 mm
from the gate: X-ray beam ( ) in the outward direction and (—) in the inward
direction [10]. (Courtesy of Verlag Lorenz.) the beam directed perpendicular to
the previous flow direction were found to be nearly symmetric. An explanation
for this is given by the assumption that the α-crystallites must be inclined with
respect to the previous flow direction. These effects were found qualitatively for
varying mould condi¬ tions where an influence of the holding time on the
inclination could be established. In samples moulded with holding times of 20 s
the sign of the inclination angle was the same from the surface down to the mid-
plane of the sample, whereas samples moulded with shorter holding times
exhibited a change of sign of inclination somewhere between the surface and the
exhibited a change of sign of inclination somewhere between the surface and the
mid-plane. A procedure was developed for analysing the a-0 4 0 profiles, as
taken with three different beam directions. In addition, an investigation into the
spatial variation of crystallite sizes was initiated. While the above mentioned
first results did not allow definite conclusions regarding correla¬ tions between
crystallite size and position in the cross-section, very recent investigations, in
which the PSD was used, showed such correlations convincingly. Secondary
crystallization seems to be important only in the highly oriented surface layers.
The degree of crystallinity in samples of KS 10 was always slightly higher than
that in samples of PT 55 (Figure 10.12). Experiments with unusual injection
moulding As already mentioned in Sections 10.1 and 10.2.1 not only
conventional injection moulding experiments were carried out (i.e. within the
framework of the normal parameter variation). ‘Unusual experiments’ means
those that did not produce injection moulded parts of reasonable quality or those
that required special alterations to the equipment. These experiments were aimed

310 Processing-induced structure formation Figure 10.12 Crystallinity


distribution over the cross-sections of test-bar specimens of KS 10 and PT 55
(227/42/20/500th), as measured by DSC at different times after injection
moulding. at enlarging the range of experiences, in the hope that some
fundamental insights could be gained. Volumetric mould filling was mentioned
in the previous section, where it was noted that useful results could not be
obtained (see Figure 10.6). Volumetric mould filling, however, rightly belongs
to this section. A special application, discussed here, was aimed at estimating the
temperature range in which surface layers, as obviously caused by shear induced
crystallization, are formed during the primary stage of mould filling, where a
packing stage cannot cause a superposition of a second effect [15]. During the
late 1970s it was shown for an amorphous polymer that the oriented layers
which solidify during the filling process cannot have a uniform thickness [16,
17]. This was explained in simple terms for a strip mould with line gate.
Obviously, just behind the gate the hot melt becomes contiguous with the cold
mould wall at a very early stage of injection. But the hot melt aimed at deeper
zones of the mould cavity flows through this region behind the gate for rather a
long time. During this time convective heat transport hampers the growth of the
oriented layer. At the far end of the mould the situation is quite different.
Because the surface layer is marked by the flow - that is, by flow-induced
birefringence in amorphous polymers and by flow-induced, highly oriented
crystallization in crystallizing polymers - the contact time is counted only until
flow stops. As a consequence, the layer will be thinner and less distinct near the
end of the mould length. There will be a maximum of the layer thickness at a
end of the mould length. There will be a maximum of the layer thickness at a
distance from the gate of about 1/3 of the length of the total mould length. There,
the layer will also be most pronounced. Such layer characteristics were also
found for the investigated polypropylene grades. A spatial temperature
distribution over the mould cavity was calculated

Phenomenological approach 311 for the moment of complete mould filling with
the aid of a numerical program. This program took into account heat conduction
in a transverse direction, heat convection in the direction of flow and the
evolution of frictional heat, but not solidification. The temperature dependence
of the viscosity, however, was taken into account in this program. Subsequently,
from this spatial temperature distribution a temperature distribution was
calculated for a line which follows the boundary of the experimentally observed
highly oriented surface layer. In this way it was found that the oriented layers of
KS 10 crystallized between 155 and 190 °C, whereas those of PT 55 solidified at
much lower temperatures, between 80 and 140 °C. The strip mould (230 mm x
70 mm x 2 mm) was used for mouldings at Tm = 200°C, Tw = 20 °C and at TM
= 280°C, Tw = 50 °C with various flow front speeds (100, 800 and 2200 mm/s,
no packing). This result is also interesting in connection with a remark made in
Section 10.2.2 with respect to the influence of the high molar mass tail of the
reactor grade KS 10, apparently causing an enhanced sensitivity to shear-
induced crystallization. The experimentally found layer thickness distribution
was used for these estimates, because of our poor knowledge of the kinetics of
shear-induced crystallization (see Section 10.3). Another, still quite unusual,
moulding process is obtained when no separate packing stage is applied
(experiment type 3 in Section 10.2.1) [5, 15, 18]. With amorphous polymers
great successes could be achieved with this process because during the packing
stage additional amounts of birefringence and other anisotropies are created.
However, with crystallizable polymers like polypropylene the advantages are not
so obvious. A drawback is that one needs particularly high clamping forces for
the mould to compensate for the high degree of shrinkage of the material during
crystallization: the melt must be com¬ pressed to a much higher degree than
with an amorphous polymer. On the other hand, this technique can be used for
special research purposes. In Figure 10.13 for the strip mould two variables
characterizing the injection stage - the axial displacement of the screw and the
pressure reading in the mould close to the gate - are plotted as functions of time
for KS 10 at TM = 200 °C, Tw = 20 °C and a speed of 100 mm/s for the flow
front in the mould. After the mould was filled it was mechanically sealed, when
the peak pressure had risen to various values, namely 29.3, 70.2 and 155.9 MPa.
Apparently, in order to reach the highest pressure, an extra full second is needed.
A most interesting observation is made even with PT 55. This is shown in Figure
A most interesting observation is made even with PT 55. This is shown in Figure
10.14. Whereas by the final increase of the pressure to 156 MPa the
birefringence close to the surface, that is, over a distance of 0.23 mm from the
surface, is not influenced, a clear rise is found between 0.23 and 0.55 mm. On
one hand this seems to mean that the oriented layer between 0 and 0.23 mm has
already been formed during volumetric filling. On the other hand, the additional
birefringence in the deeper layers is caused by the extra pressure rise. Such a
high pressure may certainly cause

312 Processing-induced structure formation Figure 10.13 Pressures in the


orthorhombic cavity, as measured close to the gate, and screw displacement as a
function of time for various ultimate pressures PN for KS 10 (200/20/100 PN/nf)
[19] (1 bar = 0.1 MPa). (Courtesy of Carl Hanser Verlag.) Figure 10.14
Birefringence profiles (100 mm from the gate) under the same condi¬ tions as in
Figure 10.13 for PT 55 and their dependence on the ultimate peak pressure [19].
(Courtesy of Hanser Verlag.)

Phenomenological approach 313 an increase of the equilibrium melting point. In


our case this probably meant that in the indicated area precursors were already
present from previous flow, which gave rise to this additional oriented
crystallization (see Section 10.3.2, for a discussion of the formation 0f
precursors at elevated temperatures). A most interesting and acute analysis of the
processes occurring during and after injection of KS 10 into the said strip mould
was given by Koppelmann and Fleischmann [19-21] on the basis of pressure
readings. These researchers have shown that uncritical use of a diagram like that
shown in Figure 10.15, can lead to a completely wrong prediction of the
maximum pressure rise needed for injection moulding without a separate
packing stage. This figure shows the specific volume plotted against tem¬
perature for KS 10, for various hydrostatic pressures. The comparatively high
value of 0.4 K/s was used as a cooling rate for all measurements. These
measurements were carried out on behalf of Petrochemie Danubia at the Aachen
Institute of Polymer Processing. This graph shows that with Tm = 200 °C a
pressure rise up to 190 MPa will be needed, if the material is to be ejected from
the mould at 35 °C and 0.1 MPa (postulating constant specific volume during
cooling). The upper part of Figure 10.16 shows pressure readings obtained with
two pressure gauges, one positioned close to the gate and the other near the
Figure 10.15 PVT diagrams for KS 10 at several pressures, for an average
cooling speed of 0.4 K/s (1 bar = 0.1 MPa).

314 Processing-induced structure formation end of the cavity. The pressure rise
314 Processing-induced structure formation end of the cavity. The pressure rise
on the first gauge, as occurring during mould filling, is recorded within the first
second. When the pressure on this gauge rises to the maximum, the gauge at the
end of the cavity also starts to indicate a sharp rise. The injection conditions
were TM = 200 °C, Tw = 20 °C, filling time 1 s, no seperate packing. The peak
pressure of 123 MPa was chosen by trial and error so that a pressure of nearly
zero was reached at the end of the experiment. In the lower part of Figure 10.16
the results of several attempts of simulating this process are given. Curve 0 was
obtained when the assumption was made that the cooling process could be
described with the aid of the declining average temperature in the sample. The
pressure decline was obtained with the postulation that the specific volume
should be constant. A first improvement was obtained when the sample was
divided into three zones, a central zone of 1 mm thickness and two zones of 0.5
mm thickness on either side. In this case, the average temperatures over these
zones were used. As the outer zones cool down more rapidly, at any pressure
Figure 10.16 Pressure readings as functions of time from two pressure gauges,
one positioned close to the gate, the other near the end of the orthorhombic
cavity (top). Pressure versus time for various simulation models, as described in
the text [20] (1 bar = 0.1 MPa). (Courtesy of Carl Hanser Verlag.)

Fundamental approach 315 the specific volume of the outer zones is smaller than
that of the central zone. At any instant the pressure was selected so that the
average of the specific volumes of the zones was constant and equal to 1.115 cm
3/g ( = specific volume at 35 °C). In this way Curve 1 was obtained. If the
extreme assumption was made that the sample stayed amorphous during cooling,
Curve 4 was obtained. This curve does not approach zero. It can be seen,
therefore, that the shrinkage by crystallization is essential. Further progress in
the simulation can only be made, however, if Figure 10.15 is modified. In
particular, a shift of the inflection points of this graph to lower temperatures with
increasing cooling rates (larger supercooling ef¬ fects) is essential for a better
simulation, as shown successively by Curves 2 and 3 in Figure 10.16. On the
other hand, crystallization induced by shearing will shift the inflexion points in
the reverse direction, that is to higher temperatures. So, it seems that shear
induction counteracts the effects shown. However, there is another interesting
point, namely, the question why much lower pressure rises are necessary in
practice than those predicted by a naive use of Figure 10.15. In fact, instead of
the predicted 190 MPa, 123 MPa was sufficient in the experiment described. If a
major contribu¬ tion to the total shrinkage is ascribed to the crystallization
process, the explanation is quite simple: if part of the melt was already solidified
during the filling stage, say by shear-induced crystallization, the specific volume
during the filling stage, say by shear-induced crystallization, the specific volume
of this part was reduced to the value of the crystallized material before the
application of 190 MPa. So, its specific volume will be further reduced by this
pressure, whereas the specific volume of the still fluid part will be reduced to its
value at 0.1 MPa and 35 °C. This means that the material as a whole will be
compressed too much. This explains why a lower peak pressure will suffice. The
thicker the shear-induced layer, the more pronounced the reduction of the
necessary peak pressure will be. As a more practical conclusion the statement
was added that diagrams like Figure 10.15 are not useful for the prediction of the
necessary maxi¬ mum pressure. From Figure 10.16 one can also see that the
core of the sample started crystallizing after 7 seconds (see the bend in the
curves in the upper part of Figure 10.16). So, after some refinements this type of
measurement will probably be a source of information on crystalliza¬ tion
kinetics with rapid cooling. There do not seem to be many other possibilities.
Injection moulding without a separate packing stage also has other advantages.
The mould shrinkage also depends on the maximum pressure, as shown in
Figure 10.17. At about 135 MPa this shrinkage is zero independent of the other
parameters of the moulding process. Also the difference in shrinkage between
the flow direction and a direction normal to the flow direction decreases with
increasing maximum pressure [15].

316 Processing-induced structure formation Figure 10.17 Mould shrinkage as a


function of the peak pressure in the orthor¬ hombic mould cavity after
mechanical sealing (without a separate packing stage) for KS 10 for two melt
temperatures and three melt front speeds [15]: (O ·), vs= 100mm/s; (Δ A), u9 =
800mm/s; (□ ■), us = 2200 mm/s; open symbols, TM = 200 °C; closed symbols
TM = 280 °C (1 bar = 0.1 MPa). 10.3 FUNDAMENTAL APPROACH:
MODEL THEORIES AND EXPERIMENTS 10.3.1 Quiescent relaxed melts A
classification of the expected processes [22] It goes without saying that the
textures described in Section 10.2 are related to the processing technology. Some
examples have already been given in Section 10.2.4, under the heading
‘Experiments with unusual injection moulding’. However, the question arises as
to how these special experiments can be understood from a more general point
of view. When researching this matter, we quickly realized that no general
classification of processes has ever been attempted: different researchers have all
been working on their own particular problems. If the thickness D of an
infinitely extended slab is taken as a characteristic dimension, one obtains for the
time τ of thermal equilibrium [23]: TD=^(l + St) (1) where a is the heat
diffusivity of the polymer and St is the Stefan number defined as: S _ h c{T\ —
Tw) (2) mould shrinkage [%] maximum pressure [bar]
Fundamental approach 317 where h is the latent heat of crystallization per unit
mass of the polymer and c is its heat capacity. For this rough estimate no
difference is made between the thermal properties c and a for the melt and for
the crystalline matter. T{ is the initial melt temperature and Tw is the
temperature of the cold mould wall. The Stefan number is simply the relative
increase of the amount of heat to be removed from the sample if there is a phase
transition in the sample. Without crystallization the amount of heat to be
removed is c(T{ — Tw) per unit mass. A second time scale is given by the speed
of crystallization. As is well known, the speed of crystallization is a strong
function of temperature. It possesses a maximum about half-way between the
glass transition tempera¬ ture and the equilibrium melting point [24]. This is a
consequence of crystallization kinetics. On descending from the melting point,
increasing nucleation activity may be observed due to increasing supercooling;
but as the glass transition temperature is approached the transport processes
necessary for nucleation become increasingly hampered. Using the mini¬ mum
half time of conversion τ1/2, which is related to the maximum in the
crystallization speed, the following ratio R can be defined [25]: R=— (3) τ1/2
The magnitude of this ratio determines the character of the process. For a large
value of R the limiting case of a process controlled by the rate of thermal
diffusion is obtained. This case applies to relatively thick samples and high
speeds of crystallization. On the other hand, a small value of Λ, which is
obtained for thin samples of low crystallization speed, characterizes a process
controlled by the nucleation rate. The ratio R is also known as the Janeschitz-
Kriegl number [26]. It is tempting to assume that metals behave according to the
limiting case of the diffusion rate controlled processes, whereas polymers
behave accord¬ ing to the nucleation rate controlled processes. However, reality
always seems to lie between these limiting cases. The diffusion rate controlled
processes are characterized by the advance¬ ment of a crystallization front
starting at the cold wall of the mould. For this front two boundary conditions are
relevant: (i) a heat balance between the heat conducted into the front, augmented
by the released heat of crystalliza¬ tion (being proportional to the speed of the
front), and the heat conducted from the front into the crystallized phase; and (ii)
the assumption that the temperature of the front is equal to the thermodynamic
melting point Tm. This assumption is equivalent to the assumption of a local
equilibrium at the boundary, which can only hold approximately for a
sufficiently high rate of crystallization. Whereas condition (i) is of a very general
nature, condition (ii) is the origin of the well-known ‘square root law’.
According to this law, the distance of the front from the wall is proportional to
the square root of
318 Processing-induced structure formation the time, as counted from the
moment when the contact is made between the hot melt and the cold wall (of
invariable temperature). In fact, because at the boundary the condition T = Tm =
const. must be obeyed for all the times and all distances, the similarity solution
of the equation of heat conduction with parameter ζ = χ/£1/2 has to hold for both
areas, the fluid and the solid. In this way, the speed of the advance of the front is
forced to obey the square root law. Because, by definition, the melt is always at
or above the melting point with this model, no nuclei can occur in the melt. So,
the solidified matter should be structureless. Another point is that local
equilibrium can only hold for sufficiently slow processes. The square root law,
however, predicts a very fast process initially: the square root of time has an
infinite slope at t = 0. So, this theory is inconsistent. It can never be valid at short
times. In the other limiting case, the sample should cool down first and
afterwards crystallize homogeneously at the temperature of the cold wall. The
conse¬ quence would be a very homogeneous texture governed by isothermal
crystallization kinetics. But everyone knows that this never happens in polymer
processing. One always finds, besides the influence of the wall, a gradient in the
size of grains from the wall to the centre, where the largest grains (spherulites)
are found. This points to the fact that in the core the cooling rate has its lowest
value. Figure 10.18 shows the general situation encountered when a melt of
uniform temperature, as contained in the half infinite space, is quenched at time
zero at the plane x = 0. The distance x at which crystallization occurs is plotted
against time. The uppermost ‘classical’ curve gives the square root law. The
almost straight curve, labelled ‘front’, gives a correct treatment of the advancing
front [27]. In this treatment only boundary condition (ii) is changed. The
temperature at the phase boundary is no longer kept constant. It is considered as
part of the solution, whereas the speed of the front is given as a unique function
of this temperature (like the speed of the growth of spherulites) [24]. This
solution can be obtained by an iterative (quickly Figure 10.18 Schematic
presentation of the advance of crystallization from a wall quench at time zero
(one-dimensional heat conduction problem).

Fundamental approach 319 converging) calculation. The initial slope of the


obtained curve is simply equal to the growth speed of the spherulites at the
temperature of the quenched wall. But a fundamental change of the situation is
characteristic for this solution. Because the temperature at the front is permitted
to be supercooled, a zone of the melt ahead of this front will also already be
supercooled. So, one cannot exclude diffuse nucleation in this zone, which
means that it is reasonable to expect a diffuse crystallization zone [25]. Such a
zone moves into the melt at a speed higher than that of the front. This difference
zone moves into the melt at a speed higher than that of the front. This difference
in speed can be understood by a comparison between a normal running race and
a relay race, where every time the baton is handed over, a length of two arms is
gained. (A spherulite coming up at a more distant place will grow also in the
opposite direction to the average advancement of crystallinity until it impinges
on another spherulite which has started growing more close to the wall. But the
first spherulite also grows in the direction of average advancement. The two
arm’s lengths can be compared with the diameter of the spherulite at the moment
of the impingement.) As a consequence, the front starting at the wall will always
be superseded at a certain distance from the wall by the diffuse zone. This is
shown by the hatched area in Figure 10.18. At this point, the nature of the front
growth must also be examined. Actually, a front is only obtained if the average
distance of nuclei at the wall is much smaller than that in the adjacent melt. The
existence of these nuclei at the wall can be traced back from the texture in the
cross-section of the solidified sample: initially, every nucleus at the wall forms
half a spherulite. As soon as these spherulites impinge on each other on the
surface of the wall, what metallurgists call a ‘columnar’ growth occurs, which is
finally stopped by the diffuse zone. If the melt is ‘inoculated’ with the aid of an
insoluble powder, the columnar phase can disappear [28]. In such a case the
distance between the nuclei in the melt becomes smaller than that on the wall
surface. At the same time, the diffuse crystallization zone becomes more narrow
and shifts towards the classical curve: the speed of crystallization is enhanced. In
polymer science the traces of the columnar phase are called the traces of
‘transcrystallization’ [29]. Inoculants are called nucleation agents. Nobody will
doubt the intertwinement of the heat transfer and the nu¬ cleation mechanisms,
when looking at cross-sections of quenched materials. From the existence of a
columnar phase one can conclude that the crystallization proceeded from the
wall surface with a finite speed, in contrast to the classical solution. What is
known as the induction time at the mould surface appears to be nothing but the
time when the thermally more active crystallization zone supersedes the front
coming from the wall. (Measurements in solidifying ingots of metals are usually
made with inserted thermocouples [30]). This time does not even depend on the
surface quality of the wall. It is related only to the heat transfer and to the
nucleation in the bulk of the melt. If the wall is completely smooth so that no
nuclei are found on it, the zone will still come up at about the same time. With
polymer melts

320 Processing-induced structure formation two complications can be observed:


(i) flow enhances nucleation activities tremendously; and (ii) the memory of a
melt for previous flow is remarkable even if this flow has happened at a
melt for previous flow is remarkable even if this flow has happened at a
temperature well above the usual melting point [31, 32]. As a consequence, the
traces of transcrystallization normally do not play any role in moulded or
extruded parts. Under the surface of these parts we always find the remainder of
previous flow rates, a subject that also deserves intensive study. Mathematical
formulation of the problems The next step must be the development of the
mathematical tools necessary for examining the situation outlined above. The
classical theory of crystallization kinetics was formulated by Kolmogoroff [33].
For spherulitic growth, as characteristic for quiescent relaxed melts,
KolmogorofFs equations read: and £(f) = l-expj-y J* dsa(s) JdzG(z) j 00 -J
Nc(oo)= ds a(s) [1 — <^(s) (4) (5) where ξ is the fraction of space covered by
freshly grown spherulites at time f, α(ί) is the rate of nucleation at time t and G(t)
is the linear growth speed of the spherulites at time t and Nc(oo) is the ultimate
number of spherulites per unit volume, if the contraction of the material as a
consequence of crystalli¬ zation is disregarded. The notation of ‘freshly’ grown
spherulites is introduc¬ ed together with the assumption that the time scale of
secondary crystalliza¬ tion is much larger than that of primary crystallization. In
fact, with polypropylene it could be shown that this assumption is quite realistic
(see Figure 10.12). The introduced time dependencies of a and G seem to be of a
quite general nature. This means that Equations 4 and 5 can also be applied to
non-isothermal situations. Usually, G can be assumed to be a unique function of
temperature T and, indirectly, a function of time, if this temperature becomes a
function of time. For a we have to assume an activation time spectrum as the
most general approach [22]: α(ί, T)=Σ ^ exp ( - (6) where Nj is the number of
sites per unit volume for activation mechanism j and Tj is the corresponding
activation time as a unique function of temperature. Also, a becomes a unique
function of time if the temperature

Fundamental approach 321 becomes a function of time. But time t also occurs
explicitly in Equation 6. When introducing the influence of flow, we will
discover a second reason for an explicit influence of time. For the moment,
however, only a quiescent melt will be treated. However, even for this case
Equations 4 and 5 are only valid if the temperature change occurs
homogeneously throughout the whole melt, which is not the case under heat
transfer situations. But these equations can still be considered as a reasonable
approximation as long as the temperature gradients are small enough. This
means that, within the size of a formed spherulite, its growth speed was still
practically constant (use of the local temperature history at the site of the nucleus
for its development). Whether this condition was fulfilled can be checked from
the calculated result of this simplified treatment. Fortunately, the bigger
the calculated result of this simplified treatment. Fortunately, the bigger
spherulites always occur where the temperature gradient is smaller. The situation
in a fixed temperature gradient has been treated recently [34]. To integrate
Equations 4 and 5 together with the equation of heat conduction: fcT,-k^T+phft
(7) which is clearly a differential equation, Equations 4 and 5 have first to be
transformed into a system of differential equations [35, 36]. (In Equation 7 p is
the density, c is the heat capacity, k is the heat conductivity and h is the latent
heat of crystallization of the freshly crystallized material). Following an idea of
Schneider, Equation 4 is successively differentiated with respect to time. With
every step an auxiliary function of interesting physical meaning is introduced. In
a later formulation by Eder the obtained set of ‘rate equations’ reads: <8a| with
the ‘initial’ conditions: φ.(—οο) = 0 for i = 0, 1, 2, 3 (8b) In particular, one has
φ0(ή = 1η(\-ξ(ή) (9) where φ0 is the volume of all spherulites at time t per unit
volume, if the consequences of the impingement are disregarded (‘unrestricted
volume fraction’), cpi(t) is the total surface of unrestrictedly grown spherulites,
φ2(ή is some less interesting scaling parameters, φ3(ή is 8π times the number of
nuclei, unrestrictedly nucleated up to time t per unit volume. Finally, the time
derivative of φ3(ή is 8π times the rate of nucleation a as introduced by
Kolmogoroff. With respect to the physical meaning οϊ φ1 it should be noted that
cp0 = G ψι (according to Equation 8a). Here we clearly see that the rate

322 Processing-induced structure formation of increase of the degree of


crystallinity is G times the available surface of crystallites. According to
Kolmogoroff’s exact statistical treatment all im¬ pingement effects are covered
by Equation 9. So, this is the first explicit demonstration of the correct
description of crystal growth under changing conditions of temperature and/or
flow field etc. The rate of creation of new spherulites is given by: Nc(t) = oc(t)
exp [ - <po(0] (10) In principle, any problem can be solved for a quiescent melt
by numerical integration with this set of equations, if the temperature history is
given for the boundary of the (polymer) sample. However, the question remains
as to how the parameters for the material’s behaviour, in particular those con¬
tained in Equation 6, can be obtained. The growth rate of spherulites Finding the
function G(T) is a less difficult task. In fact, one can determine G(T) by a direct
observation of growing spherulites at various chosen temp¬ eratures on the
heated table of a microscope [37-39]. Additional measure¬ ments at lower
temperatures, where G(T) becomes too high for such an observation, show that
transcrystallization against a steep temperature gra¬ dient is successfully
applied. This temperature gradient aims at a suppression of the diffuse
crystallization zone [31, 40]. In practice, rather thin tablets of about 1 mm
thickness are quenched only on one side to the desired growth tem¬ perature.
After what we called the contact time, the whole sample is quenched to a much
After what we called the contact time, the whole sample is quenched to a much
lower temperature. The thickness of the surface layer, as observed in the cross-
sections, is plotted against the contact time. The initial slope of such a curve
gives the desired growth speed at the temperature of the contact surface.
Although this description is correct in principle, a more refined treatment, which
improves the quality of the results, can be found in [41]. In Figure 10.19 the
logarithm of G(T) is plotted for polypropylene against temperature for the a- and
/^-modifications. In the high temperature range data of Lovinger [39] are plotted
for both modifications. In the lower temperature range the data for the α-
modification were determined with the aid of transcrystallization by Krobath,
Liedauer and Janeschitz-Kriegl [40], those for the /^-modification were recently
determined by Ratajski [41] in the same way after preparation of the contact
surface with a nucleation agent for the /?-nucleation. In agreement with the
general experience of Magill, the molar mass (distribution) has no influence on
both curves for G(T). Choice of models for primary nucleation It will be
impossible to find a method by which the activation time spectrum, as defined in
Equation 6, can be evaluated reasonably well. In

Fundamental approach 323 log G (m/sec) Figure 10.19 Logarithm of growth


speed for the a- and ^-modifications of poly¬ propylene against temperature
[41]. (Recently we have found that the three highest points for the α-
modification must also lie below the points for the ^-modification, which are
correct.) some ways, the relaxation time spectrum for the elastico-viscous
behaviour of a polymer melt forms an analogous case, but mechanical
measurements of loss and storage moduli as functions of circular frequency,
from which the latter spectrum can be deduced (with some difficulty), are of a
much higher quality than any thermophysical measurement, from which the
parameters of Equation 6 may be deduced. Admittedly, there is some advantage
with the latter measurements: additional information from the textures in the
cross-sections of the samples after the special thermal treatment can be obtained.
Curiously, not many investigators have taken advantage of this possibility. It
seems reasonable to look for suitable simplifications of Equation 6. It is clear
that all activation times Tj must be infinite at the melting point. One possibility
is to assume that, with decreasing temperature, every activation time Tj suddenly
jumps to zero when the activation temperature 7] of the underlying nucleation
mechanism is reached. In this way the integration step, by which one obtains
φ3/$π from α(ί), becomes unnecessary. One immediately obtains for φ3/8π [22]:
Ν(Τ)= Σ Wj (11) T&T According to Equation 11, N(T) has become a unique
function of tempera¬ ture like G(T). This formulation goes back to Van
Krevelen [42]. One may observe that in this model all terms of the sum in
Krevelen [42]. One may observe that in this model all terms of the sum in
Equation 6 degenerate into

delta functions at times tj at which the respective activation temperatures Tj are


reached during cooling. Van Krevelen’s formulation was favoured in the
working group because of the experience that the ultimate number of spherulites
per unit volume increases tremendously with the degree of supercooling.
Moreover, we know from the literature that with dilatometric measurements at
mild super¬ coolings an Avrami index of three has been found for many
polymers [43] and for, polypropylenes in particular [38]. In fact, with an
isothermal process (G = const., N = const.) and a delta function for a at time zero
(moment of the quench) one obtains from Equations 9 and 4: 477" φ0 = — ΝΟΨ
(12) In Equation 12 the exponent of time t is called the Avrami index [44-46]. In
practice, such a situation is found if: τ«(8π N G3)“1/3 (13) This means that the
activation time τ is short compared with the time needed for the radius of a
growing spherulite to become of the order of the average distance between the
nucleation sites. Under these circumstances the number of sites is practically
equal to the number of spherulites per unit volume. Practically no potential sites
are swallowed by spherulites which started growing at an earlier time. If there
would be, as Avrami always assumed, a temperature independent number of
nucleation sites, the above mentioned experience of an Avrami index of three, as
occurring at mild degrees of supercooling, would lead to the conclusion that with
stronger supercooling the number of spherulites should not increase. In fact, all
available sites would already have given rise to the formation of spherulites. But
this is contrary to our experience. With increasing cooling rates the number of
spherulites increases by several decades, as will be shown below. However, it
should be admitted that Avrami’s postulate, namely, a constant number of
nucleation sites per unit volume, in combination with a single activation time of
strong but continuous temperature dependence, can be sustained if the
correctness of the Avrami index of three at low degrees of supercooling is
challenged. The authors do not know of any investigation in which the texture of
the samples, as revealed by dilatometry, has been investigated. If it were proved
that part of the crystallinity in these samples was due to transcrystallization it
could be postulated that the Avrami exponent is too low. In fact,
transcrystallization represents a one-dimen¬ sional growth, for which an Avrami
index of one is predicted. This effect could cause the lowering of an Avrami
index which would otherwise be, say, four. An index of four is expected if, for
the isothermal crystallization in the bulk, a constant value of a is assumed
instead of a delta function. A constant value of a, however, is in accordance with
a reversed inequality of Equation Processing-induced structure formation 324
a reversed inequality of Equation Processing-induced structure formation 324

Fundamental approach 325 13. Under this condition a very different situation is
envisaged: almost all sites are swallowed by previously nucleated spherulites.
This can explain the relatively low number of spherulites at mild supercooling. If
with increasing supercooling the linear growth rate G(T) increases more slowly
than the nucleation frequency (reciprocal value of τ), this will result in an
increase of the ultimate number of spherulites with decreasing temperature of
crystalli¬ zation, in at least qualitative accordance with experience. This
discussion is included because several authors in the field of model calculations
believe that they can use Avrami’s model. However, only coarse effects have
been calculated so far, such as the mere progress of cooling. No spatial distribu¬
tions of spherulite sizes have ever been reported and compared with
experimental findings. In an effort to solve these problems, three experimental
methods have been applied: (i) DCS at widely varying cooling rates, (ii) visual
observation of an advancing crystallization zone in a rather thick sample (~7
mm) after quenching of one side [41], and (iii) X-ray investigations with
synchrotron radiation [4, 47]. In this context the DSC method was thoroughly
tested. It was observed that with increasing cooling rate the DSC peak shifts to
lower temperatures, which means that crystallization occurs at higher degrees of
supercooling. However, the correct temperature of crystallization always
remains higher than the temperature at the DSC peak. The correction increases
with cooling rate and can only be calculated if the (effective) heat transfer
coefficient between the mid-plane of the polymer sample and the furnace is
known. We discovered that this effective heat transfer coefficient can be derived
from the exponential return of the heat flow curve to its baseline. This can easily
be recognized if the product of the cooling rate and the logarithm of the heat
flow is plotted against the furnace temperature (which is proportional to the
time, the relevant proportionality factor being the cooling rate!). If this plot is
carried out for several cooling rates, a series of straight lines of equal slope is
obtained in the low temperature range, as shown in Figure 10.20 [48].
Unfortunately, with increasing cooling rates, these corrections become so large
that the accuracy of the measurements becomes a critical point. An evaluation of
this is presented in [49-51]. Evaluation of the advancing crystallization fronts is
also in progress. In principle, all details of the progress of crystallization during
cool¬ ing can be obtained from synchrotron diffraction patterns, as shown in
Figure 10.21. The time axis and the axis for the diffraction angle at a wavelength
of 154 pm are in the horizontal plane; the scattering intensity is plotted in the
vertical direction. Results are shown for PT 55 at a cooling rate of 20 K/min,
according to [4, 47]. The interval at which patterns were taken was 20 s.
according to [4, 47]. The interval at which patterns were taken was 20 s.
Omitting much detail, the degree of crystallinity, as obtained from the peak
areas, can be plotted against temperature, as shown in Figure 10.22 for KS 10
[47]. Unfortunately, there is a heat transfer problem here,

326 Processing-induced structure formation Figure 10.20 Unusual plot of DSC


curves (cooling rate times log Capp against furnace temperature), showing the
influence of the effective heat transfer coefficient between the mid-plane of the
sample and the furnace for KS 10 (22.9 mg in Al-pan of 33.5 mg) [48].
(Courtesy of Technomic.) too, as in the case of DSC measurements. The
thickness of the sample was 0.5 mm. This foil sample was clamped between
metal plates which were subjected to the temperature program of cooling.
However, in both plate there was a rather large hole for the X-ray beam. So,
apparently, the local temperature history of the sample in the beam area was not
sufficiently well Figure 10.21 Three-dimensional perspective presentation of
wide-angle synchro¬ tron diffraction patterns as functions of time during a
constant cooling rate of 20 °C/min for PT 55 (T{ = 210 °C, At = 20 s) [47].

Fundamental approach 327 Figure 10.22 Degrees of crystallinity as functions of


temperature for KS 10 at various cooling rates [47]. defined. In fact, the
corresponding curve of crystallized fraction versus temperature, as obtained for a
cooling rate of 50K/min in the DSC apparatus, coincides reasonably with the one
for 40K/min, as shown in Figure 10.22. However we calculated a correction of
no less than 18 K, for the DSC curve, by which the real crystallization
temperature is higher than the transition temperature of this graph. Moreover, the
correction was greater near ξ = 1 and smaller near ξ = 0, so the corrected curve
was much steeper, looking more like the curve for T= 5 K/min, and
crystallization occurred at virtually one temperature. After the correction the
curves of Figure 10.22 will not only become more similar in shape, they will
also be closer together. Unfortunately, the conclusion must be that the main
problem with all these measurements is the heat transfer problem. Only if we are
able to draw curves like those shown in Figure 10.22 against the real
crystallization temperature, will we be able to evaluate our data. Only then can
conclusions be drawn with respect to the kinetics of primary nucleation, when
use is made of our knowledge of ‘secondary nucleatin’, that is, of G(T). Results
of rapid quenching So far, we have assumed that spherulites are distinguishable
in the cross- sections of the products of our quenching experiments. However,
with more rapid quenches, no spherulites can be distinguished. Consequently, in
these situations only the overall rate of crystallization can be of interest.

328 Processing-induced structure formation In an early stage of the working


328 Processing-induced structure formation In an early stage of the working
party, Koppelmann, Fleischmann and Leitner [52] embedded a very thin
thermocouple between sandwiching polymer strips. Thin metal plates were
clamped to the sample on both sides in order to prevent it from deforming when
heated and when subsequently immersed in a liquid coolant. Cooling rates were
influenced by the tempera¬ ture of the coolant and by the chosen thickness of the
sample. In Figure 10.23 the result obtained by this quench is shown
schematically. The curve showing the temperature in the sample as a function of
time has a bump, indicating the evolution of the heat of crystallization. When a
tangent is drawn to the lower side of the curve, touching this curve in two points,
T1 and ti are obtained as the approximate temperature and time when
crystallization starts, and T2 and t2 as the temperature and time when the
crystallization is assumed to be over. The rate of cooling is approximated by
v^iTo-TJ/h (14) where T0 is 200 °C and t = 0 is the moment when this
temperature is passed during the cooling process. The temperature T3 in Figure
10.23 is the coolant temperature. If Tx and T2 are plotted against vk, Figure
10.24, is obtained which seems to hold for KS 10 as well as for PT 55. Looking
at the cooling rate of 0.67 K/s (which corresponds to 40 K/min), it can be seen
that crystallization is meant to start at 110°C can be over at 82 °C. This would be
in agreement with Figure 10.22. This is surprising in view of the reservations
mentioned in connection with this figure. It is also important to Figure 10.23
Temperature as a function of time in the mid-plane of a strip of PP quenched in a
coolant. The bump indicates the time and temperature spans of crystallization
[52]. (Courtesy of Steinkopff Verlag.)

Fundamental approach 329 remember that an evaluation of curves like the one
shown in Figure 10.23 with the aid of a drawn tangent is far from exact. Even
when the difficulties with outside heat transfer are formally circumvented with
these measurements, where the thermocouple is in the centre of the sample, the
influence of a possible non-ideal configuration of the set-up cannot be excluded.
In fact, some explanation must be found for the temporal delay of the bump in
Figure 10.23, which is obvious from the comparison between Figures 10.22 and
10.24. So, Figure 10.24 should be considered as a first overview on what will
happen at ever increasing cooling rates. It is unfortunate that results, of this type,
which are of overwhelming interest on the time scale of injection moulding, will
never be obtained with the aid of DSC or any other commercially available
equipment. In this connection, we should not forget the tremendous difficulties
encountered with any type of thermophysical measurement. A valid prediction
of the course of an injection moulding process in which crystallizable polymers
are involved must still be considered beyond the present scope of our
are involved must still be considered beyond the present scope of our
understanding. Another series of difficulties is discussed in the following
sections. Figure 10.24 Temperatures of onset and ending of crystallization of PP
as a function of cooling speed [52]. (Courtesy of Steinkopff Verlag.) 10.3.2
Shear-induced crystallization Some crucial experiments In the literature
experiments have been described in which the polymer melts are sheared to the
very moment that crystallization sets in. It always appeared that with constant
shear rate and temperature it took some time before measurable effects evolved
quite suddenly, such as decrease of free volume, multiplication of the number of
countable nuclei or shear stress increase. In some cases it could even be shown
clearly that steady flow properties were established before the onset of
crystallization phenomena. A

330 Processing-induced structure formation review of this subject was presented


recently [22]. This review also included a discussion of the drawbacks of
elongational flow for a groping investi¬ gation of flow-induced crystallization.
Furthermore, Larsen, H&nde and Gedde [53] have recently pointed out that it
will be rather difficult to maintain truely isothermal conditions during a shearing
experiment of the type mentioned, because of the quite sudden release of the
heat of crystallization. This effect is particularly serious with polymers, for
which shear induced crystallization occurs quite close to the equilibrium melting
point, and where a constant supercooling must be observed quite carefully. In the
working group the decision was made at an early stage to try to separate
nucleation from growth. It was felt that in an accumulation of effects nothing
would become clear. Two approaches to this goal were used. In the first
approach the shear treatment was carried out at a relatively high temperature
above the usual melting point, where, in addition, pro¬ longed shearing did not
lead to crystallization [22, 32]. After cessation of flow the sample was quenched
immediately or after some delay. These experiments were carried out in a duct of
rectangular cross-section of an aspect ratio 1:10. In the cross-sections of the
quenched samples layers of shear-induced crystallization were found along the
long side walls of the duct. When the time span of the delay was prolonged, the
birefringence Δη in this layer showed an exponential decay with increasing time
t of delay. From a half logarithmic plot of Δη versus £, a relaxation time could
be derived by Wippel. In Figure 10.25 relaxation times of this type are plotted
against the shear viscosities corresponding to the shearing temperatures for KS
10. From left to right, the experimental points represent temperatures between
210 °C and 190 °C in equal intervals. The importance of this double logarithmic
plot lies in the fact that its inclination is surprisingly low. If the Figure 10.25 A
double logarithmic plot of zero shear melt viscosity against the relaxation time
double logarithmic plot of zero shear melt viscosity against the relaxation time
of the precursors for KS 10, [22]. (Courtesy of Pergamon Press.)

Fundamental approach 331 orientation of separate molecules would relax, the


slope of this line should be unity. So, one may see here indirect evidence of the
presence of precursor threads, which disintegrate much more slowly as a
consequence of thermal motion after cessation of flow. In the second approach,
which can be used at temperatures below the usual melting point, short-term
shearing was applied. In these experiments shearing was so short that observable
crystallization only developed after¬ wards. For KS 10 a temperature range from
157 °C to 143 °C could be covered by these experiments. In this temperature
range ordinary crystalli¬ zation in quiescent melts takes a tremendous time, so
the effect of previous shearing is undisturbed and can be observed easily. As for
the quenching experiments, experimental details will be omitted here. Again,
flow is created in a rectangular duct. At the end of the duct glass windows are
mounted in both long side walls for optical observation [22, 54, 56]. Figures
10.26 and 10.27 show cross-sections of samples extracted from this duct after
the following treatments at 150 °C: shearing time fs = 10s, shear rate at long duct
walls yw = 43s_1; and is = 15s, yw = 68s_1. Waiting time before extraction was
about 600 s in both cases. In this time the samples solidified completely at 150
°C. After the milder treatment two fine grained surface layers on either side of
the cross-section were observed (cutting the surface in the flow direction and
perpendicular to the long side walls, taken in the middle of the sample, far from
the edges). With the more intensive shearing two types of layers were formed.
This can be seen in Figure 10.27. Close to the long side walls highly oriented
layers were obtained, but these layers do not completely cover the fine-grained
texture of the layers also found with mild shearing. Figure 10.26 Cross-section
of a strip of KS 10 from the duct of the short-term shearing apparatus after
‘mild’ shearing: T=150°C, ts=10s, yw = 43s_1 [54,55]. (Courtesy of Steinkopff
Verlag.)

332 Processing-induced structure formation In Figure 10.28 the development of


the highly oriented and birefringent layer is monitored with the aid of the optical
retardation for a shear treatment of KS 10 at 150 °C and a wall shear rate yw =
169s_1. Several applied shearing times are indicated near the curves. Using the
assumption that the degree of highly oriented crystallization is proportional to
the optical retardation, it can be concluded that there was one-dimensional
growth expressed by an Avrami index of one (linear start of the curves). Even
the longest shearing time (5 s) is so short that it cannot be observed on the time
scale of the monitoring. Another interesting result is shown in Figure 10.29. In
this double logarithmic plot a special monitoring time ί(λ/2) is plotted against the
this double logarithmic plot a special monitoring time ί(λ/2) is plotted against the
Figure 10.28 Development of optical retardation versus monitoring time for KS
10 after various shearing times: T=150°C, yw=169s_1 [54,55]. (Courtesy of
Stein¬ kopff Verlag). Figure 10.27 Cross-section of a sample after more severe
shearing: T=150°C, £s= 15 s, ^ = 68 s"1 [54, 55]. (Courtesy of Steinkopff
Verlag.)

Fundamental approach 333 shearing time ts, where ί(λ/2) means the time when
the retardation reaches half the wavelength (Γ = λ/2). As the wavelength was 550
nm, one can see from Figure 10.28, that these times were always found in the
linear range of the curves Γ versus t, representing their reciprocal initial slopes.
In Figure 10.29 results are gathered for three shear rates at the wall. Two
features may be observed: the slope of the drawn curves is —2 and the distance
between these curves corresponds with A = 41og(yi/yi_1). This pattern of lines
was superimposed on the experimental points in a best fit. Similar results were
obtained for T= 143 °C and T= 157 °C. Figure 10.29 A double logarithmic plot
of the monitoring time, at which retardation reached half the wavelength, against
shearing time, for KS 10 at T= 150 °C. Shear rates at the duct wall are indicated
near the curves. [47]. (Courtesy of Technomic.) In addition to the optical
retardation, the increasing turbidity of the samples was measured simultaneously
as a function of time. It seems that this turbidity is mainly due to the
development of the fine-grained layers. However, there is no quantitative
method of interpretation. A welcome alternative to these turbidity measurements
is formed by the measurement of the synchrotron scattering of such samples
[47]. As an example, Figure 10.30 shows the degree of crystallinity of KS 10 at
T=130°C after shear treatments at y=4s_1 and 12s_1 with shearing times
indicated on the time axis. It can be observed that, with higher shear rate,
notwithstanding the shorter shearing time, subsequent crystallization is much
faster.

334 Processing-induced structure formation o Figure 10.30 Development of


crystallinity, as determined by synchrotron scatter¬ ing, for KS 10 after shearing
at shear rates of 4 s"1 and 12s"1 at 130 °C [48]. Shearing times were 46 s and 16
s, respectively. Theoretical considerations In a previous review [22] a probability
function was introduced, on which primary and secondary nucleation should
depend. For this function a differential equation containing a growth and a decay
term was introduced. On the basis of the experiments discussed above, however,
a more direct interpretation can be attempted. The most simple case is the
‘sporadic’ creation of a great number of new nuclei by the influence of shear
flow. For this purpose the following differential equation is proposed [56]:
flow. For this purpose the following differential equation is proposed [56]:
where N is the number of nuclei per unit volume, N can be identified with the
rate of nucleation a in Kolmogoroff’s equation (Equation 4), y&tIi is a shear rate
of activation, gn a factor and τη a decay time. The quadratic dependence on γ is
introduced on the basis of an old argument of rheologists: the rate of creation
should not depend on the direction of shearing (γ2 being the second invariant of
the rate of deformation tensor). In view of the well known difficulties in the
description of the flow behaviour of polymer melts in the high shear rate range,
Equation 15 seems incredibly simple. Our excuse can only be that we believe
that our solution will be of the kinematic type and not of a dynamic type, shear
rates rather than shear stresses being of importance. If τη and ya,n are kept
constant (isothermal (15)

Fundamental approach 335 situation), a solution of Equation 15 can be obtained


easily. If flow starts at time zero with a constant rate of shear, one has: N = A κ
t„g(„[l-e ,/tn] (16) Because of our expectation that τη will be rather large in
supercooled melts, only the linear increase of N at rather short times will be of
importance for our argument: N*gn^Jt (17) An equation of the same type as
Equation 15 is obtained with the rather arbitrary assumption that threads start to
grow from a number of spots in the melt. The number of nuclei N has simply to
be replaced by the total length L of the fibrils per unit volume, and the subscripts
n have to be replaced by subscripts 1. However, a much more interesting
situation is given if the creation of starting points is sporadic. In this case one
obtains for the total length Ltot of fibrils per unit volume: Ltol = 2 N (s) ds o or,
for short-term shearing: t t L (z) dz = 2 J N (s) L(t — s) ds o ^tot-01 0n *2 *2 ^
7a,n 7a,1 (18) (19) This model was chosen on the basis of electron microscopic
evidence. According to this method threads are formed during shear flow (see
below). Another finding of electron microscopy is that lateral growth on these
threads does not show ordinary shish-kebab growth. With increasing distance
from the threads the fraction of material crystallization in an orderly manner,
which means parallel to the fibril, decreases. An increasing amount of randomly
oriented stacks of lamellae, which are stained more easly by Ru04 than the
orderly oriented crystals is found. This will be shown below. Here, the proper
model will be developed. If the material crystallizes around the threads, for the
accretion of crystallized material per unit volume (with Lt = Ltot): 2nrLt dr (20)
or, if r = Gt and b/r is the fraction of orderly oriented material (b being a
constant): 2n LtbG dt with dr = G dt (21)

336 Processing-induced structure formation where G can be assumed to be the


growth rate for a quiescent melt. This means that if the optical retardation Γ is
proportional to the fraction of orderly oriented crystals, one finds for the initial
proportional to the fraction of orderly oriented crystals, one finds for the initial
increase of Γ as a function of the monitoring time t (linear ranges in Figure
10.28): HI 2 Γ = 2n(tG)Ancb Lt(x, is)dx (22) -HI 2 where Anc is the
birefringence of the orderly crystallized material and Lt(x, is) is the total length
of the fibrils per unit volume at distance x from the mid-plane of the duct and H
is the duct height (its width being unimportant because of the large aspect ratio).
In fact, Lt depends on x, because the shear rate depends on x, whereas the
shearing time ts is the same for all values of x. For the duct flow the power law
rheology is assumed for simplicity, which means that: /2x\1/w y(*)=ywi¥J (23)
where n is the power law index and yw is the shear rate at the duct wall.
Introducing Equation 23 into Equation 22, using Equation 19, one obtains: H/2
Γ=4π(iG)Anc*>Lt,w J 0^dx (24) 0 where w again stands for the duct wall
(■ywjA.w)· In this equation there are still two unknown parameters: the
birefringence Anc of the orderly crystal¬ lized material and the factor b of
Equation 21. These parameters, however, can easily be eliminated because their
product is contained in the actually measured birefringence. In fact, the total
amount of the orderly oriented crystals in a cylinder of radius D/2 and unity
length, is approximately: D/2 2π J - r dr = nbD (25) o where D is the average
distance between the threads. The average fraction of orderly oriented crystals
thus becomes: 4nbD _ 4b nD2 = ~D (26) With this equation one has: 4b

Fundamental approach 337 where An is the actually measured birefringence


under the assumption that D is small compared with the wavelength of the light.
Before inserting Equation 27 into Equation 24 we have to realize that Lt is also
related to D. For the cylinders of diameter D, as formed around the threads, we
assume hexagonal closed packing. (For simplicity we assume a regular
arrangement of these threads.) One hexagon contains one thread in its centre and
six threads along its edges, which belong only for one-third to this hexagon. So,
in total there are three fibrils per hexagon. If this number is divided by the
surface of the hexagon, i.e. by 3(31/2 )D2, one obtains the number of threads
piercing through the unit surface, but this number is identical to the total length
of threads per unit volume. So, for D = DW: Lt,v/= ji/2 q2 Observing that the
integral of Equation 24 is equal to Hn/2(4 + n), and introducing Equations 27
and 28 into Equation 24, one finally obtains for the zone close to the duct wall: Γ
= n(tG) 4v/3Z>„ AnwH 4 + n (29) In Figure 10.28 Γ is plotted against t. This
means that we have used the monitoring time t, when Γ has reached λ/2. So, for
Dw from Equation 29: Dw = 2(^jT4^i(ί(Α/2)0)Δη» (3°) All quantities on the
right-hand side can be measured directly. So, the value of Dw calculated can be
compared directly with the one derived from the electron micrograph.
Comparison with experiments [56] The experiment used for this comparison is
Comparison with experiments [56] The experiment used for this comparison is
characterized by the following parameters: yw = 72s“1, is = 7.0s, i(A/2) = 316s,
H = 0.7mm, T=150°C. For this temperature G= 1.32 x 10_9/ms. The wavelength
of the light was A = 550nm. The birefringence near the edge of the cross-section
of the solidified sample was Aww = 0.016. (Actually, with monitoring of Γ the
light beam was in a direction perpendicular to the one in the microscope, but this
does not matter, because the birefringence in a cross-section perpendicular to the
flow direction turns out to be zero, in agreement with the model considerations
given above.) The power law index of polypropylene is n = 0.33. With these
data, from Equation 30: Dw = 592 x 10"7 m«0.6 pm (31)

338 Processing-induced structure formation The corresponding electron


micrographs (obtained after staining with Ru04) are given in Figures 10.31 and
10.32. Figure 10.31, which shows a cross-section parallel to the flow direction,
clearly shows streaks at distances less than 1 pm. Sometimes the distance is also
a little bigger. However, it is important to appreciate the statistical nature of
these micrographs and to realize that, in principle, distances should be greater in
this cross-section because of the cutting surface being parallel to the threads. On
the left-hand side, the rim of the sample, which was in touch with the wall
surface, can be seen. In Figure 10.32 a cross-section perpendicular to the flow
direction is shown. In this cross-section the places where the threads are cut
through can be seen. The cross-sections of the threads and of their orderly
crystallized overgrowth are stained less than the remaining areas. The average
distance between the threads is in fair agreement with our predictions. Finally, in
Figure 10.33 a cross-section of the sample, as photographed in the polarizing
microscope, is shown. Also in this cross-section, taken parallel to the flow
direction, streaks parallel to the duct wall can be seen. Interestingly, the much
bigger mutual distances observed are in agreement with the fact that the
distances should increase with the ratio (yw/yM)4/n· Furthermore electron
micrographs like the one shown in Figure 10.32 (section perpendicular to the
flow direction), but at a greater distance from Figure 10.31 Electron micrograph
of an ultramicrotome section parallel to the flow direction and perpendicular to
the duct wall of a sample of KS 10 after a shear for is = 7.05 s at yw = 72s_1 and
150 °C, showing the texture near the surface [56] (Courtesy of Hanser).

Fundamental approach 339 Figure 10.32 Electron micrograph for a section


similar to that in Figure 10.31, but perpendicular to the flow direction [56]
(Courtesy of Hanser). Figure 10.33 Mircrograph from a polarizing microscope
for a section similar to that in Figure 10.31 [56] (Courtesy of Hanser).
340 Processing-induced structure formation the surface of the sample,
thoroughly corroborate this expectation. Here the average distance should be
proportional to (yw/y(x))2,n> in accordance with Equation 28. More details are
given in [56]. This seems to be the first example of an investigation in which a
direct bridge has been built between the observed kinetics and the obtained
texture for the case of flow-induced crystallization. But this means that the
kinetic parameters which can be deduced from Figure 10.28 in connection with
Equation 19, will be useful for numerical simulations. Unfortunately, it will not
be easy to widen the experimental window, which is characterized for the
moment by the limits listed in Table 10.3. 10.4 ACKNOWLEDGEMENTS The
authors are very much indebted to the Fonds zur Forderung der
wissenschaftlichen Forschung (Vienna) for having sponsored this work through
many years in the course of national working party S33 and subsequent projects.
The work quoted in [56] was partly sponsored by the Christian Doppler Society
(Senior Fellowship of the first author). Figure 10.19 is from work sponsored by
the Austrian Forschungsforderungsfonds fur die gewerbliche Wirtschaft and
Petrochemie Danubia, project 5/594. The latter company, as well as Ludwig
Engel KG, Schwertberg (injection moulding machines), were also actively
engaged in the national working party. The authors are also indebted to Mr H.
Raberger for his aid in the preparation of the manuscript and of the figures. The
authors would also like to express their thanks to Mrs E. Ratajski, Mr S.
Liedauer and Mr J. Moitzi for providing figures and material from their doctoral
theses prior to publication. Finally, mention should be made of the great
readiness of all participants of the working party to cooperate in the
accomplishment of this manuscript. Table 10.3 Experimental window for shear
induced crystallization of PP Daplen KS 10 Temperature Shear rate Shearing
time 143-157°C 30-200 s'1 30-1 s 10.5 APPENDIX: SYMBOLS AND
ABBREVIATIONS a Heat diffusivity (m2/s) b Factor determining the
orderliness of crystallization (Equation 21) c Specific heat (J/kg/K)

Appendix 341 Thickness of sample Average distance between thread-like


precursors Differential scanning calorimetry Continously decreasing holding
pressure Growth rate of spherulites (m/s) Growth factor (Equation 19; m/s)
Growth factor (Equation 15; m-3/s) Gel permeation chromatography Duct height
(m) Latent heat of crystallization (J/kg) Heat conductivity (J/m/kg/s) Length per
unit volume of thread-like precursors (m-2) Melt flow index Number average
molar mass Weight average molar mass Centrifuge average molar mass Number
of nuclei per unit volume (m“3) Power law index (Equation 23) Birefringence of
oriented surface layer Birefringence of orderly crystallized material (Equation
22) The omission of a separate packing stage Pressure in the mould (Pa)
22) The omission of a separate packing stage Pressure in the mould (Pa)
Position-sensitive detector Ratio of thermal equilibration time and time of half
conversion, also called Janeschitz-Kriegl number Radius of growing cylinder
(Equation 20; m) Small-angle X-ray scattering Stefan number (Equation 2)
Temperature (K) Time (s) Transmission electron microscopy Thermodynamic
melting point Melt temperature Mechanical sealing time Thermal sealing Mould
wall temperature Volumetric mould filling Rate of cooling (Equation 14; K/s)
Axial speed of the screw during mould filling (m/s) Wide-angle X-ray scattering
Coordinate axis perpendicular to the large duct wall (rectangular cross-section
with large aspect ratio, x = 0... mid-plane) (Equation 22; m) Rate of nucleation
(m_3/s)

342 Processing-induced structure formation P τΌ τη τ 1/2 Ψ\ Γ y 7a, 1 7a,η λ ξ


Optical retardation (m) Shear rate (s-1) Shear rate of activation (Equation 19; s-
1) Shear rate of activation (Equation 15; s-1) Wavelength of light (m) Degree of
crystallinity Density (kg/m3) Time of thermal equilibration (Equation 1; s)
Decay time (Equation 15; s) Half time of conversion (Equation 3; s) Auxilliary
functions (Equations 8a, b; i = 0, 1, 2, 3) 10.6 REFERENCES 1. Wales, J. L. S.,
Van Leuwen, J. and Van der Vijgh, R. (1972) Polymer Engineering and Science,
12, 358. 2. Isayev, A. I. and Hariharan, T. (1985) Polymer Engineering and
Science, 25, 271. 3. Kantz, M. R., Newman, H. D. and Stigale F. H., (1972)
Journal of Applied Polymer Science, 16, 1249. 4. Janeschitz-Kriegl, H. (ed.)
(1991) Influence of Molecular Structure and Process¬ ing Parameters on the
Properties of Moulded Plastic Parts, Final Report of the National Working Party
S33, Linz. 5. Friesenbichler, W., Knappe, W. and Rabitsch, R. (1988)
International Polymer Proceedings, III, 191. 6. Billiani, J. and Lederer, K. (1990)
Journal of Liquid Chromatography, 13, 3013. 7. Eder, G., Janeschitz-Kriegl, H.,
Liedauer, S. et al (1989) Journal of Rheology, 33, 805. 8. Geymayer, W. and
Ingolic, E. (1988) The staining of polyolefines for TEM- investigations, in
EUREM 88, Vol. 2, Institute of Physics Conference Series No. 93, York, p. 453.
9. Glatter, O. and Kratky, O. (1982) Small Angle X-Ray Scattering, Academic
Press, London. 10. Zipper, P., Janosi, A. and Wrentschur, E. (1990)
0sterreichische Kunststoffe Zeitschrift, 21, 54. 11. Fleischmann, E. (1989)
International Polymer Proceedings, IV, 158. 12. Fleischmann, E., Zipper, P.,
Janosi, A. et al. (1989) Polymer Engineering and Science, 29, 835. 13. Janosi,
A., Zipper, P. and Wrentschur, E. (1989) The distribution of the ^-modification
in injection-moulded polypropylene discs and the ‘halo-effect’, in XIII
Hungarian Diffraction Conference, Balatonf fired, Abstract P2-38. 14. Turner-
Jones, A., Aizlewood, J. M. and Beckett, D. R. (1964) Makromolekulare
Chemie, 75, 134. 15. Fleischmann, E. (1990) Investigation of injection moulded
Chemie, 75, 134. 15. Fleischmann, E. (1990) Investigation of injection moulded
parts-of semicrystal¬ line polypropylene, PhD Thesis, Leoben University of
Mining, (in German). 16. Janeschitz-Kriegl, H. (1977) Rheologica Acta, 16, 327.
17. Janeschitz-Kriegl, H. (1979) Rheologica Acta, 18, 693. 18. Friesenbichler,
W., Knappe, W. and Pfleger, W. (1991) Kunststoffe, 81, 211. 19. Fleischmann,
E. and Koppelmann, J. (1988) Kunststoffe, 78, 453. 20. Kotmelmann, J. and
Fleischmann, E. (1988) Kunststoffe, 78, 312.

References 343 21. Fleischmann, E. and Koppelmann, J. (1990) Journal of


Applied Polymer Science, 41, 115. 22. Eder, G., Janeschitz-Kriegl, H. and
Liedauer, S. (1990) Progress in Polymer Science, 15, 629-714. / 23. Janeschitz-
Kriegl, H. (1989) Changing view on the classical Stefan problem, in One
Hundred Years of Chemical Engineering (ed. N. A. Peppas), Kluwer, Dordrecht,
p. 111. 24. Magill, H. J. (1964) Journal of Applied Physics, 35, 3249. 25. Berger,
J. and Schneider, W. (1986) Plastics and Rubber Processing and Ap¬ plications,
6, 127. 26. Astarita, G. and Kenny, J. M. (1987) Chemical Engineering
Communications, 53, 69. 27. Eder, G. and Janeschitz-Kriegl, H. (1984) Polymer
Bulletin, 11, 93. 28. Rappaz, M. (1989) International Materials Review, 34, 93.
29. Wunderlich, B. (1973) Macromolecular Physics, Vol. 1, Academic Press,
New York, pp. 282, 336. 30. Ruddle, R. W. (1957) The Solidification of
Castings, 2nd edn, The Institute of Metals, London. 31. Janeschitz-Kriegl, H.,
Eder, G., Krobath, G. and Liedauer, S. (1987) J. Non- Newtonian Fluid
Mechanics, 23, 107. 32. Janeschitz-Kriegl, H., Wimberger-Friedl, R., Krobath,
G. and Liedauer, S. (1987) Kautschuk und Gummi, Kunststojfe, 40, 301. 33.
Kolmogoroff, A. W. (1937) Isvestiya Akademii Nauk SSSR, Ser. Math. 1, 355
(in Russian). 34. Schulze, G. E. W. and Naujeck, T. R. (1991) Colloid and
Polymer Science, 269, 689, 695. 35. Schneider, W., Koppl, A. and Berger, J.
(1988) International Polymer Pro¬ ceedings, II, 151. 36. Berger, J. (1988) PhD
Thesis, Vienna University of Technology. 37. Padden, F. J. and Keith, H. D.
(1959) Journal of Applied Physics, 30, 1479. 38. von Falkai, B. (1960)
Makromolekulare Chemie, 41, 86. 39. Lovinger, A. J., Chua, J. O. and Gryte, C.
C. (1977) Journal of Polymer Science, 15, 641. 40. Krobath, G., Liedauer, S. and
Janeschitz-Kriegl, H. (1985) Polymer Bulletin, 14, 1. 41. Ratajski, E. (1993)
PhD thesis, Linz University (in German). 42. Van Krevelen, D. W. (1978)
Chimia, 32, 279. 43. Mandelkern, L. (1964) Crystallization of Polymers,
McGraw-Hill, New York. 44. Avrami, M. (1939) Journal of Chemical Physics,
7, 1103. 45. Avrami, M. (1940) Journal of Chemical Physics, 8, 212. 46.
Avrami, M. (1941) Journal of Chemical Physics, 9, 177. 47. Moitzi, J. (1992)
Crystallization kinetic investigations on isotactic polypropylene by means of
synchrotron scattering, PhD Thesis, Vienna University of Technol¬ ogy (in
synchrotron scattering, PhD Thesis, Vienna University of Technol¬ ogy (in
German). 48. Eder, G. and Janeschitz-Kriegl, H. (1993) Heat transfer and flow:
Transport phenomena controlling the crystallization processes in polymers, in
1st Interna¬ tional Conference on Transport Phenomena, 22-26 March 1992,
Honolulu, Hawaii, Technomic, pp. 1031. 49. Janeschitz-Kriegl, H., Wippel, H.,
Paulik, C. and Eder, G. (1993) Colloid and Polymer Science, 271, 1107. 50. Wu,
C. H., Eder, G. and Janeschitz-Kriegl, H. (1993) Colloid and Polymer Science,
271, 1116.

344 Processing-induced structure formation 51. Eder, G. and Wu, C. H. (1994)


Colloid and Polymer Science, in press. 52. Koppelmann, J., Fleischmann, E. and
Leitner, G. (1987) Rheologica Acta, 26, 548. 53. Larsen, A. H£nde, O. and
Gedde, U. (1991) Experiments in shear-induced crystallization under isothermal
boundary conditions, in Proceedings of the 7th Annual Meeting of the Polymer
Processing Society, Hamilton, Ontario, April. 54. Liedauer, S. PhD Thesis, Linz
University (in preperation). 55. Janeschitz-Kriegl, H. (1992) Progress in Colloid
and Polymer Science, 87, 117. 56. Liedauer, S., Eder, G., Janeschitz-Kriegl, H.
et al, (1993) International Polymer Processingy 8, 236.

to figures and page numbers appearing in Calorimetric analysis crystalline


phases and their stability 22-3 Calorimetry 62 Charpy impact test 280
Copolymer crystal structure 45-50 different compositions 45-9 different
tacticities 48-50 definition 4 molecular structure 4-24 PP 32 Creep behaviour of
self-reinforced PP 285 Critical entanglement molecular weight 247-8 PE 247-8
PP 248 theoretical and experimental values 247 Cross-hatching 89-90, 141, 142,
144, 145 Crystalline structures of polypropylene homo- and copolymers 31-55
crystallization kinetics theory 320-22 crystal structure of IPP 34-41 hexagonal
lattice of /?-form 37-9 quenched forms of IPP crystals 40-41 triclinic lattice of y-
form 39-40 crystal structure of PP copolymers 45-50 crystal structure of SPP 41-
5 soace erouns of IPP 35

Index Crystalline structures of polypropylene contd Crystallization of /?-IPP 74-


9 kinetics of growth 75-7 nucleations 75 overall crystallization 77-9 Cylindritic
supermolecular structure in IPP 91-8 Decoration effect 152, 153 Degree of
polymerization 4, 10 Descartes’ oval 86 Differential scanning calorimetry (DSC)
7, 281, 302, 325-6 injection-moulded PP 171 self-reinforced PP 290, 289
Disentanglement and fracture of interfaces 251-3 molecular weight dependence
of fracture 253 welding at interfaces 251-2 Double crystallization 62 Drawing
274-5 die drawing 274-5 drawing of gels 275 simple drawing 274 zone drawing
275 DSC, see Differential scanning calorimetry (DSC) Dynamic mechanical
analysis 23, 281 changes of conformation 23-4 Enantiotropy 100, 102
analysis 23, 281 changes of conformation 23-4 Enantiotropy 100, 102
Entanglement 243-51 critical entanglement molecular weight 247-8 deformation
and disentanglement 248-51 entanglement model for random coil polymers 243-
4 Environmental stress cracking 207 EPDM, see Ethylene-propylene diene
monomer (EPDM) Epitaxial growth 140-63 decorational effects on IPP 152-4
heteroepitaxies in IPP 145-51 heteroepitaxies in SPP 151 homepitaxy in IPP
141-5 influence on mechanical properties 154-61 TDD until UArtYiο 1
noroffinc 1^1 0 EPM, see Ethylene-propylene copolymer (EPM) EPR, see
Ethylene-propylene rubbers (EPR) Ethylene-propylene copolymer (EPM) IR
analysis 19-21 Raman spectroscopy 21-2 Ethylene-propylene diene monomer
(EPDM) 6, 13 Ethylene-propylene rubbers (EPR) 6 Extrusion effect of process
parameters 285-7 hydrostatic extrusion 276 process description 279 ram
extrusion 276 thermal and microscopic analysis 288-89 Flexural modulus
injection-moulded PP 195, 196, 197, 198 Flory-Huggins thermodynamics theory
9, 234, 257 Fold surface free energy 59 Fracture disentanglement and fracture
249-51, 251-3 fracture energy 227, 230, 260-61 fracture theory 242-3 molecular
weight dependence of fracture 253 y-form of IPP crystallization 123 triclinic
lattice 39-40, 141 Gel 275 Gel permeation chromatography (GPC) 5, 8
measurement of molecular weight distribution 11-13 Glass fiber injection-
moulded reinforced PP 188-9 GPC, see Gel permeation chromatography (GPC)
Graphoepitaxy 140 Heat of fusion 63 Heteroepitaxy IPP 145-51 μαΙιιιυιογο i ςι
346

Index 347 Homoepitaxy 141 IPP 141-5 Homopolymer 4 Huggins coefficient 9-


11 Hydrostatic extrusion 276 Incompatible semicrystalline interfaces 253-8
interface preparation and thermal history 254-5 PP/PE interfaces 258-68
isothermally crystallized PP/ HDPE and PP/LLDPE 258-64 non-isothermally
crystallized PP/ HDPE 264 non-isothermally crystallized PP/ LLDPE 264-8
structure development 255-8 crystallization 258 interdiffusion 257 solidification
258 surface rearrangement 255-6 wetting 256-7 Infra-red (IR) spectroscopy 6
analysis of EPM and PP 19-21 Injection-moulded polypropylene addition of
crystallization nucleators 184-5 crystalline structure 168-75 effect of
copolymerization 181-4 glass fiber filling 188-9 higher order structure 167-204
effect of molecular mass 176-81 effect of moulding conditions 175-6 molecular
orientation process 189-94 moulding conditions 175-6 particulate filling 185-8
process description 278 properties of injection-moulded PP 194-7 self-
reinforcement 281-4 skin-core structure 168-72, 176, 183 structure-property
relationships 198-201 thermal and microscopic analysis 289-92 unusual injection
moulding 309-15 Interdiffusion 257 Interfaces classification of interfaces 232-
42, 233 amorphous structural asymmetry 234-6 asymmetric polymer-polymer
42, 233 amorphous structural asymmetry 234-6 asymmetric polymer-polymer
interfaces 234 incompatible amorphous interfaces 236-9 multicomponent
polymer interfaces 239-40 symmetric polymer-polymer interfaces 232-4 two-
component blends with particles 240-42 disentanglement and fracture of
interfaces 251-3 molecular weight dependence of fracture 253 welding at
interfaces 251-2 incompatible semicrystalline interfaces 253-69 interface
preparation and thermal history 254-5 PP/PE interfaces 258-68 surface
development at semicrystalline interfaces 255-8 interdiffusion of random coil
changes 228 polymer interfaces 228-42 polymer welding 228-9 strength
development at interfaces 229-31 structure and strength of interfaces 231-2 IPP,
see Isotactic polypropylene (IPP) Isomerism 5, 5-6 stereo-isomerisms in
polypropylene chains 6-7 Isotactic polypropylene (IPP) 6-7, 7, 15, 31-6, 34, 56-7
α-modification of IPP 32, 34, 35-7, 57-68 crystallization of a-IPP 57-62 melting
characteristics 72, 73, 74, 62-8 ^-modification of IPP 32, 34, 37-9, 68-85
crystallization 74-9 melting characteristics 72, 73, 74, 79-85, 105, 106

348 Index Isotactic polypropylene (IPP) contd preparation of 68-74


recrystallization 79-85 y-modification of IPP 32, 34, 39-40 crystal structure 34-
41 crystal unit cell 34 decorational effects on IPP 152-4 epitaxies in IPP with
normal paraffins 151-2 heteroepitaxies in IPP 145-51 homoepitaxy 141-5
nucleating agents for a-IPP 128-30 polymorphism in IPP 100-104 primary
nucleation in a-IPP 123-8, 133 primary nucleation in β-ΙΡΡ 130-31, 133
quenched form of IPP crystals 40-41 secondary nucleation in IPP 131-4 space
groups 35 supermolecular structure of IPP 85-99 cylindritic structure 91-8
spherulitic structure 86-91 transcrystalline structure 98-9 Izod impact strength
195, 197 Janeschitz-Kriegl number 317 Knit-line behaviour 205-27 formation
210-13, 212 knit-line factor 219-22, 220, 221, 222, 224 mechanical properties
219-23 morphology 213—9 mould filling simulation (isochrones) for a cup 207
origin 206-10 PP/EPDM blend 214, 215, 216, 217, 217, 223, 224 weakness 210
Kratky collimation system 301 Lamella spherulite 89-90 thickening 68
Laminates mechanical properties and epitaxial interfaces 154-61 preparation of
multilayer films 157 Linear growth rate 58, 59 Loss modulus injection-moulded
PP 172 Maltese cross 87 Mark-Houwink relationship 11, 12, 13 Mechanical
properties self-reinforcement 280-81, 283, 284, 287-8 Mechanically loaded
melts (preparation of /LIPP) 74 Melting memory effect 69 Miller indices 306
Molar mass distribution 299 Molecular structure of polypropylene homo- and
copolymers 3-30 characteristics of polymer chain 4-7 chemical structure 5-6
molecular size 4-5 stereo-isomerisms in polypropylene chains 6-7 chemical
composition 14-25 crystalline phases and their stability 22-3 dynamical changes
of conformation 23-4 IR analysis of EPM and polypropylene 19-21
of conformation 23-4 IR analysis of EPM and polypropylene 19-21
microstructure 14 NMR analysis of chain configuration 17-19 Raman
spectroscopy of polypropylene and EPM 21-2 molecular weight distribution of
polypropylenes 8-13 dissolution of polypropylene 8-9 molecular weight
selection 9-13 Molecular weight molecular weight distribution of
polypropylenes 8-13 dissolution of polypropylene 8-9 molecular weight
selection 9-13 number-average molecular weight 4 weight-average molecular
weight 5 z-average molecular weight 5 Monomer 4 Monotropy 100 Nuclear
magnetic resonance (NMR) spectroscopy 6 analysis of PP chain 17-19

Index 349 dynamic changes of conformation 23-4 Nucleation athermal 58


nucleants in injection-moulded PP 184-5 polymers 116-19 primary spherulite
nucleation in PP 119-30 α-form of IPP 123 general remarks 119-22 models for
324-8 nucleating agents for a-IPP 128-30 nucleation of /MPP 130-31 secondary
nucleation in IPP 131-4 syndiotactic PP 134-5, 135 thermal 58 Orientation
function in injection- moulded PP 177-9, 179, 188, 191, 192, 193 Oriented
polymer blends preparation technique 154 Oriented polypropylene 3 Particulate
filling injection-moulded PP 185-8 Polymer blends mechanical properties and
epitaxial interfaces 154-61 Polymer chain, characteristics 4-7 chemical structure
5-6 molecular size 4-5 Polymer fractionation 9-10 Polymolecularity index
(polydispersity) 11 Polymorphism in IPP 100-104 PP/PE interfaces 258-68
isothermally crystallized PP/HDPE and PP/LLDPE 258-64 non-isothermally
crystallized PP/HDPE 264 non-isothermally crystallized PP/LLDPE 264-8
Primary crystallization of IPP 57 Processing 167 processing technology 297-8
Processing-induced structure formation 295-344 model theories and experiments
quiescent relaxed melts 315-26 shear-induced crystallization 326-40
phenomenological approach 297-315 experimental method 299-302 molar mass
distribution 299 morphology 302-15 PP grades 298-9 processing technology
297-8 Quenched form of IPP crystals 40-1 Ram extrusion 276 Raman
spectroscopy PP and EPM 21-2 Random copolymers 45 different compositions
45-7 different tacticities 48-50 Rapid quenching 327-9 Recooling temperature 79
Rolling 276 SAXD, see Small angle X-ray diffraction (SAXD) SEC, see Size
exclusion chromatography (SEC) Self-nucleation 127 Self-reinforcement of PP
273-94 melt processes 276-88 experimental 278-81 properties 281-8 solid-state
processes 274-6 die drawing 274-5 drawing of gels 275 hydrostatic extrusion
276 ram extrusion 276 rolling 276 simple drawing 274 zone drawing 275
thermal and microscopic analysis 288-92 extrusion 288-9 injection moulding
289-92 Shear modulus PP extrudates 289 Shear-induced crystallization 310,
329-40 ‘Shish-kebab’ structure 174, 175 Size exclusion chromatography (SEC)
5 Small angle X-ray diffraction (SAXD) 302, 304, 305 Kratky collimation
system 301
system 301

Index Small angle X-ray diffraction (SAXD) contd injection-moulded PP 170,


171 Smetic 24, 56 Spherulite growth rate 322-3 injection-moulded PP 168
primary spherulite nucleation in PP 119-30 structure in IPP 86-91 anti-epitaxy
90-91 cross-hatching and autoepitaxy 89-90, 142 influence of thermal conditions
88-9 types 87-8 SPP, see Syndiotactic polypropylene (SPP) Stefan number 316
Stereo-isomerisms in polypropylene chains 6-7 atactic 6-7, 7 isotactic 6-7, 7
syndiotactic 6-7, 7 Stereo-isomerisms in polypropylene chains 6-7 Storage loss
injection-moulded PP 172 Strain hardening 248-9 Structure development at
semicrystalline interfaces 255-8 crystallization 258 interdiffusion 257
solidification 258 surface rearrangement 255-6 wetting 256-7 Supermolecular
structure of IPP 85-99 cylindritic structure 91-8 spherulitic structure 86-91
transcrystalline structure 98-9 Syndiotactic polypropylene (SPP) 6-7, 7, 15, 31-2
crystallization 135 crystal structure 41-5 heteroepitaxies 151 high temperature
orthorhombic form 41-3 low temperature, quenched orthorhombic form 43-4
nucleation 134-5, 135 triclinic form 44-5 Temperature gradient crystallization 70
Tensile yield strength injection-moulded PP 195, 200 self-reinforced PP 288
Thermal conduction 190-91, 321 Thermal expansion self-reinforced PP 290, 291
Thermo-mechanical analysis (TMA) 7 Thermodynamical equilibrium melting
point 63 TMA, see Thermo-mechanical analysis (TMA) Transcrystallization 319
transcrystalline structures in IPP 85, 98-9 WAXD, see Wide angle X-ray
diffraction (WAXD) Weld-line, see Knit-line behaviour Welding 227-72
disentanglement and fracture of interfaces 251-3 molecular weight dependence
of fracture 253 welding at interfaces 251-2 fracture theory 242-3 incompatible
semicrystalline interfaces 253-69 interface preparation and thermal history 254-2
PP/PE interfaces 258-68 structure development at semicrystalline interfaces 255-
8 polymer entanglements 243-51 critical entanglement molecular weight 247-8
deformation and disentanglement 248-51 entanglement model for random coil
polymers 243-4 polymer interfaces 228-42 classification of interfaces 232-42
polymer welding 228-9, 251-2 strength development at interfaces 229-31
structure and strength of interfaces 231-2 Wetting 256-7 Wide angle X-ray
diffraction (WAXD), 34, 42, 301, 306 injection moulded PP 169, 169, 174, 350

Index 351 β-ΙΡΡ 70 crystalline phases and their stability 22-3 Ziegler-Natta
catalysis 7, 18 Zone drawing 275

Potrebbero piacerti anche